() @ Appl. Gen. Topol. 19, no. 1 (2018), 21-53doi:10.4995/agt.2018.7146 c© AGT, UPV, 2018 Alternate product adjacencies in digital topology Laurence Boxer Department of Computer and Information Sciences, Niagara University, NY 14109, USA; and Department of Computer Science and Engineering, SUNY at Buffalo (boxer@niagara.edu) Communicated by S. Romaguera Abstract We study properties of Cartesian products of digital images, using a variety of adjacencies that have appeared in the literature. 2010 MSC: 54C99; 05C99. Keywords: digital topology; digital image; retraction; approximate fixed point property; continuous multivalued function; shy map. 1. Introduction We study various adjacency relations for Cartesian products of multiple dig- ital images. We are particularly interested in “product properties” - properties that are preserved by taking Cartesian products - and “factor properties” for which possession by a Cartesian product of digital images implies possession of the property by the factors. Many of the properties examined in this paper were considered in [9] for adjacencies based on the normal product adjacency. We consider other adjacencies in this paper, including the tensor product adja- cency, the Cartesian product adjacency, and the composition or lexicographic adjacency. Received 23 January 2017 – Accepted 05 October 2017 http://dx.doi.org/10.4995/agt.2018.7146 L. Boxer 2. Preliminaries Much of the material that appears in this section is quoted or paraphrased from [9, 12], and other papers cited in this section. We use N, Z, and R to represent the sets of natural numbers, integers, and real numbers, respectively, A digital image is a graph. Usually, we consider the vertex set of a digital image to be a subset of Zn for some n ∈ N. Further, we often, although not always, restrict our study of digital images to finite graphs. We will assume familiarity with the topological theory of digital images. See, e.g., [3] for many of the standard definitions. All digital images X are assumed to carry their own adjacency relations (which may differ from one image to another). When we wish to emphasize the particular adjacency relation we write the image as (X, κ), where κ represents the adjacency relation. 2.1. Common adjacencies. To denote that x and y are κ-adjacent points of some digital image, we use the notation x ↔κ y, or x ↔ y when κ can be understood. The cu-adjacencies are commonly used. Let x, y ∈ Z n, x 6= y. Let u be an integer, 1 ≤ u ≤ n. We say x and y are cu-adjacent, x ↔cu y, if • there are at most u indices i for which |xi − yi| = 1, and • for all indices j such that |xj − yj| 6= 1 we have xj = yj. A cu-adjacency is often denoted by the number of points adjacent to a given point in Zn using this adjacency. E.g., • In Z1, c1-adjacency is 2-adjacency. • In Z2, c1-adjacency is 4-adjacency and c2-adjacency is 8-adjacency. • In Z3, c1-adjacency is 6-adjacency, c2-adjacency is 18-adjacency, and c3-adjacency is 26-adjacency. For Cartesian products of digital images, the normal product adjacency (see Definitions 2.1 and 2.2) has been used in papers including [22, 6, 11, 9] (errors in [22] are corrected in [6]). The tensor product adjacency (see Definition 2.3), Cartesian product adjacency (see Definition 2.4), and the lexicographic adja- cency (see Definition 2.6) have not to our knowledge been studied in digital topology, so their respective roles in digital topology remain to be determined. Given digital images or graphs (X, κ) and (Y, λ), the normal product ad- jacency NP(κ, λ), also called the strong product adjacency (denoted κ∗(κ, λ) in [11]) generated by κ and λ on the Cartesian product X × Y is defined as follows. Definition 2.1 ([1, 28]). Let x, x′ ∈ X, y, y′ ∈ Y . Then (x, y) and (x′, y′) are NP(κ, λ)-adjacent in X × Y if and only if • x = x′ and y ↔λ y ′; or • x ↔κ x ′ and y = y′; or • x ↔κ x ′ and y ↔λ y ′. As a generalization of Definition 2.1, we have the following. c© AGT, UPV, 2018 Appl. Gen. Topol. 19, no. 1 22 Product adjacencies in digital topology Definition 2.2 ([9]). Let u and v be positive integers, 1 ≤ u ≤ v. Let {(Xi, κi)} v i=1 be digital images. Let NPu(κ1, . . . , κv) be the adjacency defined on the Cartesian product Πvi=1Xi as follows. For xi, x ′ i ∈ Xi, p = (x1, . . . , xv) and q = (x′1, . . . , x ′ v) are NPu(κ1, . . . , κv)-adjacent if and only if • for at least 1 and at most u indices i, xi ↔κi x ′ i, and • for all other indices i, xi = x ′ i. Definition 2.3 ([20]). The tensor product adjacency on the Cartesian product Πvi=1Xi of (Xi, κi), denoted T (κ1, . . . , κv), is as follows. Given xi, x ′ i ∈ Xi, we have (x1, . . . , xv) and (x ′ 1, . . . , x ′ v) are T (κ1, . . . , κv)-adjacent in Π v i=1Xi if and only if for all i, xi ↔κi x ′ i. Figure 1. A digital simple closed curve and its Cartesian product with [0, 1]Z. (a) shows the simple closed curve MSC8 ⊂ (Z 2, c2) [21]. (b) shows the set MSC8 × [0, 1]Z ⊂ Z 3 with either the c2 × c1- or the NP1(c2, c1)-adjacency. (c) shows the set MSC8 × [0, 1]Z ⊂ Z 3 with the T (c2, c1)- adjacency, where adjacencies are shown by the solid lines. If the points of MSC8 are circularly labeled p0, . . . , p5, then the T (c2, c1)-neighbors of (pi, t) are (p(i−1) mod 6, 1 − t) and (p(i+1) mod 6, 1 − t), t ∈ {0, 1}. Definition 2.4 ([26]). The Cartesian product adjacency on the Cartesian product Πvi=1Xi of (Xi, κi), denoted × v i=1κi or κ1 × . . . × κv, is as follows. Given xi, x ′ i ∈ Xi, we have (x1, . . . , xv) and (x ′ 1, . . . , x ′ v) are × v i=1κi-adjacent in Πvi=1Xi if and only if for some i, xi ↔κi x ′ i, and for all indices j 6= i, xj = x ′ j. The following has an elementary proof. Proposition 2.5. For Πvi=1(Xi, κi), × v i=1κi = NP1(κ1, . . . , κv). Definition 2.6 ([19]). Let (Xi, κi) be digital images, 1 ≤ i ≤ v. Let xi, x ′ i ∈ Xi. Let p = (x1, . . . , xv), p ′ = (x′1, . . . , x ′ v). We say p and p ′ are adjacent in the composition or lexicographic adjacency on Πvi=1Xi if x1 ↔κ1 x ′ 1, or if for some c© AGT, UPV, 2018 Appl. Gen. Topol. 19, no. 1 23 L. Boxer index j, 1 ≤ j < v, we have (x1, . . . , xj) = (x ′ 1, . . . , x ′ j) and xj+1 ↔κj+1 x ′ j+1. The adjacency is denoted L(κ1, . . . , κv). Figure 2. An illustration of lexicographic adjacency. This is [0, 1]Z × {−2, 0, 2}, with both factors regarded as subsets of (Z, c1), and the L(c1, c1) adjacency. Remark 2.7. Notice that for p and p′ to be L(κ1, . . . , κv)-adjacent with xk and x′k κk-adjacent, for indices m > k we do not require that xm and x ′ m be either equal or adjacent. See, e.g., Figure 2, where (0, 0) and (1, 2) are L(c1, c1)-adjacent. This is unlike other adjacencies discussed above. 2.2. Connectedness. A subset Y of a digital image (X, κ) is κ-connected [25], or connected when κ is understood, if for every pair of points a, b ∈ Y there exists a sequence {yi} m i=0 ⊂ Y such that a = y0, b = ym, and yi ↔κ yi+1 for 0 ≤ i < m. For two subsets A, B ⊂ X, we will say that A and B are adjacent when there exist points a ∈ A and b ∈ B such that a and b are equal or adjacent. Thus sets with nonempty intersection are automatically adjacent, while disjoint sets may or may not be adjacent. It is easy to see that a finite union of connected adjacent sets is connected. 2.3. Continuous functions. The following generalizes a definition of [25]. Definition 2.8 ([4]). Let (X, κ) and (Y, λ) be digital images. A function f : X → Y is (κ, λ)-continuous if for every κ-connected A of X we have that f(A) is a λ-connected subset of Y . When the adjacency relations are understood, we will simply say that f is continuous. Continuity can be reformulated in terms of adjacency of points: Theorem 2.9 ([25, 4]). A function f : X → Y is continuous if and only if, for any adjacent points x, x′ ∈ X, the points f(x) and f(x′) are equal or adjacent. Note that similar notions appear in [14, 15] under the names immersion, gradually varied operator, and gradually varied mapping. c© AGT, UPV, 2018 Appl. Gen. Topol. 19, no. 1 24 Product adjacencies in digital topology Theorem 2.10 ([3, 4]). If f : (A, κ) → (B, λ) and g : (B, λ) → (C, µ) are continuous, then g ◦ f : (A, κ) → (C, µ) is continuous. Example 2.11 ([25]). A constant function between digital images is continu- ous. Example 2.12. The identity function 1X : (X, κ) → (X, κ) is continuous. Definition 2.13. Let (X, κ) be a digital image in Zn. Let x, y ∈ X. A κ-path of length m from x to y is a set {xi} m i=0 ⊂ X such that x = x0, xm = y, and xi−1 and xi are equal or κ-adjacent for 1 ≤ i ≤ m. If x = y, we say {x} is a path of length 0 from x to x. Notice that for a path from x to y as described above, the function f : [0, m]Z → X defined by f(i) = xi is (c1, κ)-continuous. Such a function is also called a κ-path of length m from x to y. 2.4. Digital homotopy. A homotopy between continuous functions may be thought of as a continuous deformation of one of the functions into the other over a finite time period. Definition 2.14 ([4]; see also [23]). Let (X, κ) and (Y, κ′) be digital images. Let f, g : X → Y be (κ, κ′)-continuous functions. Suppose there is a positive integer m and a function F : X × [0, m]Z → Y such that • for all x ∈ X, F(x, 0) = f(x) and F(x, m) = g(x); • for all x ∈ X, the induced function Fx : [0, m]Z → Y defined by Fx(t) = F(x, t) for all t ∈ [0, m]Z is (2, κ′)−continuous. That is, Fx(t) is a path in Y . • for all t ∈ [0, m]Z, the induced function Ft : X → Y defined by Ft(x) = F(x, t) for all x ∈ X is (κ, κ′)−continuous. Then F is a digital (κ, κ′)−homotopy between f and g, and f and g are digitally (κ, κ′)−homotopic in Y . If for some x0 ∈ X we have F(x0, t) = F(x0, 0) for all t ∈ [0, m]Z, we say F holds x0 fixed, and F is a pointed homotopy. We denote a pair of homotopic functions as described above by f ≃κ,κ′ g. When the adjacency relations κ and κ′ are understood in context, we say f and g are digitally homotopic (or just homotopic) to abbreviate “digitally (κ, κ′)−homotopic in Y ,” and write f ≃ g. Proposition 2.15 ([23, 4]). Digital homotopy is an equivalence relation among digitally continuous functions f : X → Y . Definition 2.16 ([5]). Let f : X → Y be a (κ, κ′)-continuous function and let g : Y → X be a (κ′, κ)-continuous function such that f ◦ g ≃κ′,κ′ 1X and g ◦ f ≃κ,κ 1Y . c© AGT, UPV, 2018 Appl. Gen. Topol. 19, no. 1 25 L. Boxer Then we say X and Y have the same (κ, κ′)-homotopy type and that X and Y are (κ, κ′)-homotopy equivalent, denoted X ≃κ,κ′ Y or as X ≃ Y when κ and κ′ are understood. If for some x0 ∈ X and y0 ∈ Y we have f(x0) = y0, g(y0) = x0, and there exists a homotopy between f ◦ g and 1X that holds x0 fixed, and a homotopy between g ◦ f and 1Y that holds y0 fixed, we say (X, x0, κ) and (Y, y0, κ ′) are pointed homotopy equivalent and that (X, x0) and (Y, y0) have the same pointed homotopy type, denoted (X, x0) ≃κ,κ′ (Y, y0) or as (X, x0) ≃ (Y, y0) when κ and κ ′ are understood. It is easily seen, from Proposition 2.15, that having the same homotopy type (respectively, the same pointed homotopy type) is an equivalence relation among digital images (respectively, among pointed digital images). 2.5. Continuous and connectivity preserving multivalued functions. Given sets X and Y , a multivalued function f : X → Y assigns a subset of Y to each point of x. We will write f : X ⊸ Y . For A ⊂ X and a multivalued function f : X ⊸ Y , let f(A) = ⋃ x∈A f(x). Definition 2.17 ([24]). A multivalued function f : X ⊸ Y is connectivity preserving if f(A) ⊂ Y is connected whenever A ⊂ X is connected. As is the case with Definition 2.8, we can reformulate connectivity preser- vation in terms of adjacencies. Theorem 2.18 ([12]). A multivalued function f : X ⊸ Y is connectivity preserving if and only if the following are satisfied: • For every x ∈ X, f(x) is a connected subset of Y . • For any adjacent points x, x′ ∈ X, the sets f(x) and f(x′) are adjacent. Definition 2.17 is related to a definition of multivalued continuity for subsets of Zn given and explored by Escribano, Giraldo, and Sastre in [16, 17] based on subdivisions. (These papers make a small error with respect to compositions, that is corrected in [18].) Their definitions are as follows: Definition 2.19. For any positive integer r, the r-th subdivision of Zn is Z n r = {(z1/r, . . . , zn/r) | zi ∈ Z}. An adjacency relation κ on Zn naturally induces an adjacency relation (which we also call κ) on Znr as follows: (z1/r, . . . , zn/r), (z ′ 1/r, . . . , z ′ n/r) are adjacent in Znr if and only if (z1, . . . , zn) and (z ′ 1, . . . , z ′ n) are adjacent in Z n. Given a digital image (X, κ) ⊂ (Zn, κ), the r-th subdivision of X is S(X, r) = {(x1, . . . , xn) ∈ Z n r | (⌊x1⌋, . . . , ⌊xn⌋) ∈ X}. Let Er : S(X, r) → X be the natural map sending (x1, . . . , xn) ∈ S(X, r) to (⌊x1⌋, . . . , ⌊xn⌋). Definition 2.20. For a digital image (X, κ) ⊂ (Zn, κ), a function f : S(X, r) → Y induces a multivalued function F : X ⊸ Y if x ∈ X implies F(x) = ⋃ x′∈E −1 r (x) {f(x′)}. c© AGT, UPV, 2018 Appl. Gen. Topol. 19, no. 1 26 Product adjacencies in digital topology Definition 2.21. A multivalued function F : X ⊸ Y is called continuous when there is some r such that F is induced by some single valued continuous function f : S(X, r) → Y . Figure 3. [12] Two images X and Y with their second sub- divisions. (Subdivisions are drawn at half-scale.) Note [12] that the subdivision construction (and thus the notion of continu- ity) depends on the particular embedding of X as a subset of Zn. In particular we may have X, Y ⊂ Zn with X isomorphic to Y but S(X, r) not isomorphic to S(Y, r). E.g., in Figure 3, when we use 8-adjacency for all images, X and Y are isomorphic, each being a set of two adjacent points, but S(X, 2) and S(Y, 2) are not isomorphic since S(X, 2) can be disconnected by removing a single point, while this is impossible in S(Y, 2). The definition of connectivity preservation makes no reference to X as being embedded inside of any particular integer lattice Zn. Proposition 2.22 ([16, 17]). Let F : X ⊸ Y be a continuous multivalued function between digital images. Then • for all x ∈ X, F(x) is connected; and • for all connected subsets A of X, F(A) is connected. Theorem 2.23 ([12]). For (X, κ) ⊂ (Zn, κ), if F : X ⊸ Y is a continuous multivalued function, then F is connectivity preserving. The subdivision machinery often makes it difficult to prove that a given multivalued function is continuous. By contrast, many maps can easily be shown to be connectivity preserving. 2.6. Other notions of multivalued continuity. Other notions of continu- ity have been given for multivalued functions between graphs (equivalently, between digital images). We have the following. Definition 2.24 ([27]). Let F : X ⊸ Y be a multivalued function between digital images. • F has weak continuity if for each pair of adjacent x, y ∈ X, f(x) and f(y) are adjacent subsets of Y . c© AGT, UPV, 2018 Appl. Gen. Topol. 19, no. 1 27 L. Boxer • F has strong continuity if for each pair of adjacent x, y ∈ X, every point of f(x) is adjacent or equal to some point of f(y) and every point of f(y) is adjacent or equal to some point of f(x). Proposition 2.25 ([12]). Let F : X ⊸ Y be a multivalued function between digital images. Then F is connectivity preserving if and only if F has weak continuity and for all x ∈ X, F(x) is connected. Example 2.26 ([12]). If F : [0, 1]Z ⊸ [0, 2]Z is defined by F(0) = {0, 2}, F(1) = {1}, then F has both weak and strong continuity. Thus a multivalued function between digital images that has weak or strong continuity need not have connected point-images. By Theorem 2.18 and Proposition 2.22 it follows that neither having weak continuity nor having strong continuity implies that a multivalued function is connectivity preserving or continuous. Example 2.27 ([12]). Let F : [0, 1]Z ⊸ [0, 2]Z be defined by F(0) = {0, 1}, F(1) = {2}. Then F is continuous and has weak continuity but does not have strong continuity. Proposition 2.28 ([12]). Let F : X ⊸ Y be a multivalued function between digital images. If F has strong continuity and for each x ∈ X, F(x) is con- nected, then F is connectivity preserving. The following shows that not requiring the image of a point F(p) to be connected can yield topologically unsatisfying consequences for weak and strong continuity. Example 2.29 ([12]). Let X and Y be nonempty digital images. Let the multivalued function f : X ⊸ Y be defined by f(x) = Y for all x ∈ X. • f has both weak and strong continuity. • f is connectivity preserving if and only if Y is connected. As a specific example [12] consider X = {0} ⊂ Z and Y = {0, 2}, all with c1 adjacency. Then the function F : X ⊸ Y with F(0) = Y has both weak and strong continuity, even though it maps a connected image surjectively onto a disconnected image. 2.7. Shy maps and their inverses. Definition 2.30 ([5]). Let f : X → Y be a continuous surjection of digital images. We say f is shy if • for each y ∈ Y , f−1(y) is connected, and • for every y0, y1 ∈ Y such that y0 and y1 are adjacent, f −1({y0, y1}) is connected. Shy maps induce surjections on fundamental groups [5]. Some relation- ships between shy maps f and their inverses f−1 as multivalued functions were studied in [7, 12, 8]. Shyness as a factor or product property for the normal product adjacency was studied in [9]. We have the following. c© AGT, UPV, 2018 Appl. Gen. Topol. 19, no. 1 28 Product adjacencies in digital topology Theorem 2.31 ([12, 8]). Let f : X → Y be a continuous surjection between digital images. Then the following are equivalent. • f is a shy map. • For every connected Y0 ⊂ Y , f −1(Y0) is a connected subset of X. • f−1 : Y ⊸ X is a connectivity preserving multi-valued function. • f−1 : Y ⊸ X is a multi-valued function with weak continuity such that for all y ∈ Y , f−1(y) is a connected subset of X. 2.8. Other tools. Other terminology we use includes the following. Given a digital image (X, κ) ⊂ Zn and x ∈ X, the set of points adjacent to x ∈ Zn and the neighborhood of x in Zn are, respectively, Nκ(x) = {y ∈ Z n | y is κ-adjacent to x}, N∗κ(x) = Nκ(x) ∪ {x}. 3. Maps on products In this section, we consider various product adjacencies with respect to con- tinuity of functions. 3.1. General properties. Definition 3.1. Let κ1 and κ2 be adjacency relations on a set X. We say κ1 dominates κ2, κ1 ≥d κ2, or κ2 is dominated by κ1, κ2 ≤d κ1, if for x, x ′ ∈ X, if x and x′ are κ1-adjacent then x and x ′ are κ2-adjacent. Example 3.2. We have the following comparisons of adjacencies. • For X ⊂ Zn and 1 ≤ u ≤ v ≤ n, cu ≥d cv. • For Πvi=1(Xi, κi) and 1 ≤ u ≤ v ≤ n, NPu(κ1, . . . κv) ≥d NPv(κ1, . . . κv). • For Πvi=1(Xi, κi), T (κ1, . . . κv) ≥d NPv(κ1, . . . κv). • For Πvi=1(Xi, κi), we have: – NPu(κ1, . . . , κv) ≥d L(κ1, . . . , κv) for 1 ≤ u ≤ v; – T (κ1, . . . , κv) ≥d L(κ1, . . . , κv); – ×vi=1κi ≥d L(κ1, . . . , κv). Proof. These follow immediately from the definitions of these adjacencies. � The next example shows that there are adjacencies that can be applied to the same set X such that neither dominates the other. Example 3.3. In X = Z6 = Z3 × Z3, neither of T (c2, c2) nor T (c1, c3) domi- nates the other. Proof. Consider the points p = (0, 0, 0, 0, 0, 0) and q = (1, 1, 0, 1, 1, 0). We have p ↔T (c2,c2) q but p and q are not T (c1, c3)-adjacent. Therefore T (c2, c2) does not dominate T (c1, c3). Now consider r = (1, 0, 0, 1, 1, 1). We have p ↔T (c1,c3) r but p and r are not T (c2, c2)-adjacent. Therefore T (c1, c3) does not dominate T (c2, c2). � c© AGT, UPV, 2018 Appl. Gen. Topol. 19, no. 1 29 L. Boxer Domination, and being dominated, are transitive relations among the adja- cencies of a graph. I.e., we have the following. Proposition 3.4. Given adjacencies κ, λ, µ for a graph, if κ ≤d λ and λ ≤d µ, then κ ≤d µ. Proof. Elementary, and left to the reader. � Proposition 3.5. Let f : X → Y be a function. • Let λ1 and λ2 be adjacency relations on Y . If f is (κ, λ1) continuous and λ1 ≥d λ2, then f is (κ, λ2) continuous. • Let κ1 and κ2 be adjacency relations on X. If f is (κ1, λ)-continuous and κ1 ≤d κ2, then f is (κ2, λ)-continuous. Proof. The assertions follows from the definitions of continuity and the ≥d relation. � Given functions fi : (Xi, κi) → (Yi, λi), 1 < i ≤ v, the function Πvi=1fi : Π v i=1Xi → Π v i=1Yi is defined by (Πvi=1fi)(x1, . . . , xv) = (f1(x1), . . . , fv(xv)), where xi ∈ Xi. 3.2. Normal product. Here, we recall continuity properties of the normal product adjacency. Theorem 3.6 ([9]). Let fi : (Xi, κi) → (Yi, λi), 1 < i ≤ v. Then the product map f = Πvi=1fi : (Π v i=1Xi, NPv(κ1, . . . , κv)) → (Π v i=1Yi, NPv(λ1, . . . , λv)) is continuous if and only if each fi is continuous. Theorem 3.7 ([9]). Let X = Πvi=1Xi. Let fi : (Xi, κi) → (Yi λi), 1 ≤ i ≤ v. • For 1 ≤ u ≤ v, if the product map f = Πvi=1fi : (X, NPu(κ1, . . . , κv)) → (Πvi=1Yi, NPu(λ1, . . . , κv)) is an isomorphism, then for 1 ≤ i ≤ v, fi is an isomorphism. • If fi is an isomorphism for all i, then the product map f = Π v i=1fi : (X, NPv(κ1, . . . , κv)) → (Π v i=1Yi, NPv(λ1, . . . , κv)) is an isomorphism. Theorem 3.8 ([22, 9]). The projection maps pi : (Π v j=1Xj, NPu(κ1, . . . , κv)) → (Xi, κi) defined by pi(x1, . . . , xv) = xi for xi ∈ (Xi, κi), are all continuous, for 1 ≤ u ≤ v. 3.3. Tensor product. For the tensor product adjacency, we have the follow- ing. Proposition 3.9. Suppose X = Πvi=1Xi has a pair of T (κ1, . . . , κv)-adjacent points. Then • each Xi has 2 κi-adjacent points; and c© AGT, UPV, 2018 Appl. Gen. Topol. 19, no. 1 30 Product adjacencies in digital topology • If f : (X, T (κ1, . . . , κv)) → (Π w j=1Yj, T (λ1, . . . , λw)) is continuous and not constant on some component of X, then for every j, Yj has 2 λj- adjacent points. Proof. Let p = (x1, . . . , xv) and p ′ = (x′1, . . . , x ′ v) be T (κ1, . . . , κv)-adjacent in X. Then for each i, xi and x ′ i are κi-adjacent in Xi, which establishes the first assertion. Further, if f is as hypothesized, the continuity of f implies there are T (κ1, . . . , κv)-adjacent p, p ′ such that f(p) = (y1, . . . , yw) and f(p ′) = (y′1, . . . , y ′ w) are unequal, hence T (λ1, . . . , λw)-adjacent. Therefore, for all j, yj and y′j are λj-adjacent. � It is easy to construct examples showing that the assertions obtained from Proposition 3.9 by substituting the normal product adjacency NPv for T are false. Theorem 3.10. Let X = Πvi=1Xi, Y = Π v i=1Yi. If the product map f = Πvi=1fi : (X, T (κ1, . . . , κv)) → (Y, T (λ1, . . . , λv)) is continuous, then for each i, fi : (Xi, κi) → (Yi, λi) is continuous. Proof. If xi, x ′ i are κi-adjacent in Xi, then p = (x1, . . . , xv) and p ′ = (x′1, . . . , x ′ v) are T (κ1, . . . , κv)-adjacent in X. Thus f(p) and f(p ′) are equal or T (λ1, . . . , λv)- adjacent in Y . This implies fi(xi) and fi(x ′ i) are equal or λi-adjacent in Yi. Thus fi is continuous. � However, the converse to Theorem 3.10 is not generally true, as shown in the following. Example 3.11. Let f : [0, 1]Z → [0, 1]Z be the identity function. Let g : [0, 1]Z → [0, 1]Z be the constant function g(x) = 0. Then, using Examples 2.12 and 2.11, f and g are each (c1, c1)-continuous, but f × g : [0, 1]Z × [0, 1]Z → [0, 1]Z × [0, 1]Z is not (T (c1, c1), T (c1, c1))-continuous. Proof. This follows from the observations that (0, 0) and (1, 1) are T (c1, c1)- adjacent, but (f × g)(0, 0) = (0, 0) and (f × g)(1, 1) = (1, 0) are neither equal nor T (c1, c1)-adjacent. � A partial converse to Theorem 3.10 is obtained by using the following notion. Definition 3.12. A continuous function f : (X, κ) → (Y, λ) is locally one-to- one if f|N∗ κ (x,1) is one-to-one for all x ∈ X. Note any function between digital images that is one-to-one must be locally one-to-one. Theorem 3.13. Suppose fi : (Xi, κi) → (Yi, λi) is continuous and locally one- to-one for 1 ≤ i ≤ v. Then the product function f = Πvi=1fi : Π v i=1Xi → Π v i=1Yi is (T (κ1, . . . , κv), T (λ1, . . . , λv))-continuous and locally one-to-one. c© AGT, UPV, 2018 Appl. Gen. Topol. 19, no. 1 31 L. Boxer Proof. Suppose fi : (Xi, κi) → (Yi, λi) is continuous and locally one-to-one for 1 ≤ i ≤ v. Let p = (x1, . . . , xv) and p ′ = (x′1, . . . , x ′ v) be T (κ1, . . . , κv)- adjacent, where xi and x ′ i are κi-adjacent in Xi. Since fi is continuous and locally one-to-one, we must have that fi(xi) and fi(x ′ i) are λi-adjacent in Yi. Thus, f(p) and f(p′) are T (λ1, . . . , λv)-adjacent, so f is continuous and locally one-to-one. � Theorem 3.14. Let X = Πvi=1Xi, Y = Π v i=1Yi. Then the product map f = Πvi=1fi : (X, T (κ1, . . . , κv)) → (Y, T (λ1, . . . , λv)) is an isomorphism if and only if each fi is an isomorphism. Proof. If f is an isomorphism, each fi must be one-to-one and onto. Therefore, f−1i : Yi → Xi is a single-valued function. By Theorem 3.10, each fi is continuous. Since f −1 = Πvi=1f −1 i , it follows from Theorem 3.10 that each f−1i is continuous. Hence fi is an isomorphism. Conversely, if each fi is an isomorphism, then f is one-to-one and onto, so f−1 = Πvi=1f −1 i is a single-valued function. By Theorem 3.13, f is continuous. Similarly, f−1 is continuous. Therefore, f is an isomorphism. � Theorem 3.15. The projection maps pi : (Π v i=1Xi, T (κ1, . . . , κv)) → (Xi, κi) defined by pi(x1, . . . , xv) = xi for xi ∈ Xi are all continuous. Proof. Let p = (x1, . . . , xv) and p ′ = (x′1, . . . , x ′ v) be T (κ1, . . . , κv)-adjacent in Πvi=1Xi, where xi, x ′ i ∈ Xi. Then for all indices i, xi = pi(p) and x ′ i = pi(p ′) are κi-adjacent. Thus, pi is continuous. � A seeming oddity is that a common method of injection that is often contin- uous, is not continuous when the tensor product adjacency is used, as shown in the following. Proposition 3.16. Let (X, κ) and (Y, λ) be digital images. Let y ∈ Y . If X has a pair of κ-adjacent points, then the function f : X → (X × Y, T (κ, λ)) defined by f(x) = (x, y) is not continuous. Proof. This is because given κ-adjacent x, x′ ∈ X, f(x) = (x, y) and f(x′) = (x′, y) are not T (κ, λ)-adjacent. � 3.4. Cartesian product. Theorem 3.17. Let fi : (Xi, κi) → (Yi, λi) be functions between digital im- ages, 1 ≤ i ≤ v. Let X = Πvi=1Xi, Y = Π v i=1Yi. Then the product function f = Πvi=1fi : X → Y is (× v i=1κi, × v i=1λi)-continuous if and only if each fi is continuous. Proof. Suppose f is continuous. Let xi ↔κi x ′ i in Xi. Let p = (x1, . . . , xv), p′ = (x1, . . . , xi−1, x ′ i, xi+1, . . . , xv). Then p ↔×vi=1κi p ′, so either f(p) = f(p′) or f(p) ↔×v i=1 λi f(p ′). The former case implies fi(xi) = fi(x ′ i) and the latter case implies fi(xi) ↔λi fi(x ′ i). Hence fi is continuous. c© AGT, UPV, 2018 Appl. Gen. Topol. 19, no. 1 32 Product adjacencies in digital topology Suppose each fi is continuous. Let p and p ′ be ×vi=1κi-adjacent points of X. Then there is only one index k in which p and p′ differ, i.e., for some xi ∈ Xi and x ′ k ∈ Xk, p = (x1, . . . , xv), p ′ = (x1, . . . , xk−1, x ′ k , xk+1, . . . , xv), and xk ↔κk x ′ k. Then f(p) and f(p ′) have the same ith coordinate for i 6= k, and have kth coordinates of fk(xk) and fk(x ′ k), respectively. Continuity of fk implies either fk(xk) = fk(x ′ k) or fk(xk) ↔κk fk(x ′ k). Therefore, f is continuous. � Theorem 3.18. The projection maps pi : (Π v i=1Xi, × v i=1κi) → (Xi, κi) defined by pi(x1, . . . , xv) = xi for xi ∈ Xi are all continuous. Proof. This follows from Proposition 2.5 and Theorem 3.8. � By contrast with Proposition 3.16, we have the following. Proposition 3.19. Let (Xi, κi) be digital images, 1 ≤ i ≤ v. Let xi ∈ Xi. The functions Ii : Xi → (Π v i=1Xi, × v i=1κi) defined by Ii(x) =    (x, x2, . . . , xv) for i = 1; (x1, . . . , xi−1, x, xi+1 . . . , xv) for 1 < i < v; (x1, . . . , xv−1, x) for i = v, are continuous. Proof. This follows immediately from Definition 2.4. � Theorem 3.20. Let X = Πvi=1Xi, Y = Π v i=1Yi. Then the product map f = Πvi=1fi : (X, × v i=1κi) → (Y, × v i=1λi) is an isomorphism if and only if each fi is an isomorphism,. Proof. Suppose f is an isomorphism. Then it follows from Proposition 2.5 and Theorem 3.7 that fi is an isomorphism. Suppose each fi is an isomorphism. Then f must be one-to-one and onto, and by Theorem 3.17, f is continuous. Similarly, f−1 = Πvi=1f −1 i is continuous. Therefore, f is an isomorphism. � 3.5. Lexicographic adjacency. Theorem 3.21. Suppose fi : (Xi, κi) → (Yi, λi) is a function between digital images, 1 ≤ i ≤ v. Let f = Πvi=1fi : Π v i=1Xi → Π v i=1Yi be the product function. • If f is (L(κ1, . . . , κv), L(λ1, . . . , λv))-continuous, then each fi is (κi, λi)- continuous. Further, if f is locally one-to-one, then each fi is locally one-to-one. • If each fi is a continuous function that is locally one-to-one, then f is (L(κ1, . . . , κv), L(λ1, . . . , λv))-continuous. Proof. Suppose f is (L(κ1, . . . , κv), L(λ1, . . . , λv))-continuous. Let xi, x ′ i ∈ Xi such that xi ↔κi x ′ i. Let p0 = (x1, x2, . . . , xv) and let pi =    (x′1, x2, . . . , xv) for i = 1; (x1, . . . , xi−1, x ′ i, xi+1, . . . , xv) for 1 < i < v; (x1, . . . , xv−1, x ′ v) for i = v. c© AGT, UPV, 2018 Appl. Gen. Topol. 19, no. 1 33 L. Boxer Notice (3.1) p0 and pi differ only at index i and p0 ↔L(κ1,...,κv) pi for 1 ≤ i ≤ v. Therefore, f(p0) and f(pi) are L(λ1, . . . , λv)-adjacent or equal. It follows from statement (3.1) that fi(xi) and fi(x ′ i) are λi-adjacent or equal. Since {xi, x ′ i} is an arbitrary set of κi-adjacent members of Xi, fi is (κi, λi)-continuous. Further if f is locally one-to-one, then from statement (3.1), fi(xi) and fi(x ′ i) are not equal, so fi is locally one-to-one. Suppose each fi is continuous and locally one-to-one. Let p, p ′ ∈ X = Πvi=1Xi, where p = (x1, . . . , xv), p ′ = (x′1, . . . , x ′ v), for xi, x ′ i ∈ Xi. Assume p ↔L(κ1,...,κv) p ′. Let k be the smallest index such that xk ↔κk x ′ k. Since fk is locally one-to-one, (3.2) fk(xk) ↔λk fk(x ′ k). • If k = 1, it follows from Definition 2.6 that f(p) ↔L(λ1,...,λv) f(p ′). • Otherwise, i < k implies xi = x ′ i, hence fi(xi) = fi(x ′ i). Together with statement (3.2), this implies f(p) ↔L(λ1,...,λv) f(p ′). Then f is (L(κ1, . . . , κv), L(λ1, . . . , λv))-continuous, since p and p ′ were arbi- trarily chosen. � The following example illustrates the importance of the locally one-to-one hypothesis in Theorem 3.21. Example 3.22. Let Xi = [0, i]Z for i ∈ {1, 2}. Let f : X1 → X2 be the con- stant function with value 0. Then f and 1X2 are (c1, c1) continuous. However, f × 1X2 : X1 × X2 → X 2 2 is not (L(c1, c1), L(c1, c1))-continuous. Proof. Consider the points p = (0, 0) and p′ = (1, 2). These points are L(c1, c1)- adjacent in X1 × X2. However, (f × 1X2)(p) = (0, 0) and (f × 1X2)(p ′) = (0, 2) are neither equal nor L(c1, c1)-adjacent in X 2 2. � Theorem 3.23. Suppose fi : (Xi, κi) → (Yi, λi) is a function between digital images, 1 ≤ i ≤ v. Let f = Πvi=1fi : Π v i=1Xi → Π v i=1Yi be the product function. Then f is an (L(κ1, . . . , κv), L(λ1, . . . , λv))-isomorphism if and only if each fi is a (κi, λi)-isomorphism. Proof. This follows easily from Theorem 3.21. � Proposition 3.24. The projection map p1 : (Π v i=1Xi, L(κ1, . . . , κv)) → (X1, κ1) is continuous. Proof. Let p ↔L(κ1,...,κv) p ′ in Πvi=1Xi. Then p = (x1, . . . , xv), p ′ = (x′1, . . . , x ′ v) for some xi, x ′ i ∈ Xi, where either x1 = x ′ 1 or x1 ↔κ1 x ′ 1. Since p1(p) = x1 and p1(p ′) = x′1, it follows that p1 is continuous. � By contrast, we have the following. Example 3.25. Let v > 1. The projection maps pi : ([0, 2] v Z , L(c1, . . . , c1)) → ([0, 2]Z, c1) are not continuous for 1 < i ≤ v. c© AGT, UPV, 2018 Appl. Gen. Topol. 19, no. 1 34 Product adjacencies in digital topology Proof. Let x = (0, 0, . . . , 0), y = (1, 2, . . . , 2). Then x ↔L(c1,...,c1) y in [0, 2] v Z , but i > 1 implies pi(x) = 0 and pi(y) = 2, which are not c1-adjacent in [0, 2]Z. The assertion follows. � 3.6. More on isomorphisms. We have the following. Theorem 3.26. Let σ : {i}vi=1 → {i} v i=1 be a permutation. Let fi : (Xi, κi) → (Yσ(i), λσ(i)) be an isomorphism of digital images, 1 ≤ i ≤ v. Let 1 ≤ u ≤ v. Let (κ, λ) be any of (NPu(κ1, . . . , κv), NPu(λσ(1), . . . , λσ(v))), (T (κ1, . . . , κv), T (λσ(1), . . . , λσ(v))), or (×vi=1κi, × v i=1λσ(i)). Let X = Πvi=1Xi, Y = Π v i=1Yσ(i). Then the function f : X → Y defined by f(x1, . . . , xv) = (f1(x1), . . . , (fv(xv)) is an isomorphism. Proof. It is easy to see that f is one-to-one and onto. Continuity of f and of f−1 follows easily from the definitions of the adjacencies under discussion. Thus, f is an isomorphism. � The following example shows that the lexicographic adjacency does not yield a conclusion analogous to that of Theorem 3.26. Example 3.27. Let X1 = {0, 1} ⊂ (Z, c1). Let X2 = {0, 2} ⊂ (Z, c1). Then X = (X1 × X2, L(c1, c1)) and Y = (X2 × X1, L(c1, c1)) are not isomorphic. Proof. Observe that X is connected, since the 4 points of X form a path in the sequence (0, 0), (1, 0), (0, 2), (1, 2) (see Figure 2). However, Y is not connected, as there is no path in Y from (0, 0) to (2, 0). The assertion follows. � 4. Connectedness In this section, we compare product adjacencies with respect to the property of connectedness. Theorem 4.1 ([9]). Let (Xi, κi) be digital images, i ∈ {1, 2, . . . , v}. Then (Xi, κi) is connected for all i if and only (Π v i=1Xi, NPv(κ1, . . . , κv)) is con- nected. Theorem 4.2. Let (Xi, κi) be digital images, i ∈ {1, 2, . . . , v}. If Π v i=1Xi is T (κ1, . . . , κv)-connected, then Xi is κi-connected for all i. Proof. These assertions follow from Definition 2.8 and Theorem 3.15. � However, the converse to Theorem 4.2 is not generally true, as shown by the following. c© AGT, UPV, 2018 Appl. Gen. Topol. 19, no. 1 35 L. Boxer Example 4.3. Let X = {0} ⊂ Z, Y = [0, 1]Z ⊂ Z. Then X and Y are each c1-connected. However: • X × Y = {(0, 0), (0, 1)} is not T (c1, c1)-connected. • Y × Y has two T (c1, c1)-components, {(0, 0), (1, 1)} and {(1, 0), (0, 1)}. See also Figure 1(c), which illustrates that MSC8 × [0, 1]Z is not T (c2, c1)- connected, although MSC8 is c2-connected and [0, 1]Z is c1-connected. For the Cartesian product adjacency, we have the following. Theorem 4.4. Let (Xi, κi) be digital images, i ∈ {1, 2, . . . , v}. Then Π v i=1Xi is ×vi=1κi-connected if and only if Xi is κi-connected for all i. Proof. Suppose X = Πvi=1Xi is × v i=1κi-connected. It follows from Proposi- tion 3.18 that each Xi is κi-connected. Suppose each Xi is κi-connected. Let p = (x1, . . . , xv) and p ′ = (x′1, . . . , x ′ v) be points of X such that xi, x ′ i ∈ Xi. There are κi-paths Pi in Xi from xi to x′i. If the functions Ii are as in Proposition 3.19, then it is easily seen that ⋃v i=1 Ii(Pi) is a × v i=1κi-path in X from p to p ′. Since p and p′ were arbitrarily chosen, it follows that X is ×vi=1κi-connected. � Proposition 4.5. Let (X, κ) and (Y, λ) be digital images, such that |X| > 1. Then (X × Y, L(κ, λ)) is connected if and only if (X, κ) is connected. Proof. Suppose (X, κ) is connected. Let p = (x, y), p′ = (x′, y′), with x, x′ ∈ X, y, y′ ∈ Y . • If x = x′ then, since |X| > 1 and X is connected, there exists x0 ∈ X such that x ↔κ x0. Therefore, p ↔L(κ,λ) (x0, y) ↔L(κ,λ) (x, y ′) = p′. • Suppose x 6= x′. Since X is connected, there is a path in X, P = {xi} n i=0, such that x = x0 ↔κ x1 ↔κ . . . ↔κ xn−1 ↔κ xn = x ′. Therefore, p = (x0, y) ↔L(κ,λ) (x1, y ′) ↔L(κ,λ) (x2, y ′) ↔L(κ,λ) . . . ↔L(κ,λ) (xn, y ′) = p′. Therefore, (X × Y, L(κ, λ)) is connected. Suppose (X, κ) is not connected. Then there exist x, x′ ∈ X such that x and x′ are in distinct components of X. Let y, y′ ∈ Y . By Definition 2.6, there is no path in (X × Y, L(κ, λ)) from (x, y) to (x′, y′). Therefore, (X × Y, L(κ, λ)) is not connected. � An argument similar to that used for the proof of Proposition 4.5 yields the following. Theorem 4.6. Let (Xi, κi) be digital images, 1 ≤ i ≤ v. Suppose k is the smallest index for which |Xk| > 1. Then (Π v i=1Xi, L(κ1, . . . , κv)) is connected if and only if (Xk, κk) is connected. c© AGT, UPV, 2018 Appl. Gen. Topol. 19, no. 1 36 Product adjacencies in digital topology 5. Homotopy 5.1. Tensor product. In [9], it is shown that many homotopy properties are preserved by Cartesian products with the NPv adjacency. We show that we cannot make analogous claims for the tensor product adjacency. Example 5.1. There are digital images (Xi, κi) and (Yi, λi) and continuous functions fi, gi : Xi → Yi, i ∈ {1, 2}, such that fi ≃ gi but f1 × f2 6≃T (κ1,κ2),T (λ1,λ2) g1 × g2. Proof. We can use Example 4.3. E.g., if X1 = X2 = Y1 = Y2 = [0, 1]Z, f1 = f2 : X1 → Y1 is the identity function, and g1 = g2 : X2 → Y2 is the constant function taking the value 0, we have f1 ≃c1,c1 g1 and f2 ≃c1,c1 g2. As we saw in Example 4.3, [0, 1]2 Z is not T (c1, c1)-connected, so its identity function f1 × f2 is not homotopic to the constant function g1 × g2. � Example 5.2. There are digital images (Xi, κi) and (Yi, λi) for i ∈ {1, 2}, such that Xi and Yi have the same homotopy type, but (X1 × X2, T (κ1, λ1)) and (Y1 × Y2, T (κ2, λ2)) do not have the same homotopy type. Proof. We saw in Example 4.3 that [0, 1]2 Z is not T (c1, c1)-connected; however, it is trivial that {0}2 = {(0, 0)} is T (c1, c1)-connected. Therefore, we can take X1 = X2 = [0, 1]Z ⊂ (Z, c1), Y1 = Y2 = {0} ⊂ (Z, c1). � 5.2. Cartesian product adjacency. Theorem 5.3. Let fi, gi : (Xi, κi) → (Yi, λi) be continuous functions between digital images, 1 ≤ i ≤ v. Let X = Πvi=1Xi, Y = Π v i=1Yi, f = Π v i=1fi : X → Y , g = Πvi=1gi : X → Y . Then f ≃×vi=1κi,× v i=1 λi g if and only if for all i, fi ≃κi,λi gi. Further, f and g are pointed homotopic if and only if for each i, fi and gi are pointed homotopic. Proof. Suppose f ≃×v i=1 κi,× v i=1 λi g. Then there is a homotopy H : Πvi=1Xi × [0, m]Z → Π v i=1Xi such that H(p, 0) = f(p) and H(p, m) = g(p) for all p ∈ X. Let xi ∈ Xi and let Hi : Xi × [0, m]Z → Yi be defined by Hi(x, t) = pi(H(Ii(x), t)), where Ii is the continuous injection of Proposition 3.19 corresponding to the point (x1, . . . , xv) ∈ X and pi is the continuous projection map of Theo- rem 3.18. Then Hi(x, 0) = pi(f(Ii(x)) = fi(x) and Hi(x, m) = pi(g(Ii(x)) = gi(x). Since the composition of continuous functions is continuous (Theorem 2.10), it follows that Hi is a homotopy from fi to gi. Further, if H holds some point p0 of X fixed, then we can take p0 = (x1, . . . , xv) to be the point of X used in Proposition 3.19, and we can conclude that Hi holds pi(p) = xi fixed. Suppose for all i, fi ≃κi,λi gi. Let Hi : Xi × [0, mi]Z → Yi be a (κi, λi)- homotopy from fi to gi. We execute these homotopies “one coordinate at a c© AGT, UPV, 2018 Appl. Gen. Topol. 19, no. 1 37 L. Boxer time,” as follows. For x = (x1, . . . , xv) ∈ X such that xi ∈ Xi, let Mi = ∑i k=1 mi for all i and let H : X×[0, Mv]Z → Y be defined by H(x1, . . . , xv, t) = • (H1(x1, t), f2(x2), . . . , fv(xv)) if 0 ≤ t ≤ m1; • (g1(x1) . . . , gj−1(xj−1), Hj(xj, t−Mj−1), fj+1(xj+1), . . . , fv(xv)) if Mj−1 ≤ t ≤ Mj; • (g1(x1) . . . , gv−1(xv−1), Hv(xv, t − Mv−1)) if Mj−1 ≤ t ≤ Mj. It is easily seen that H is well defined and is a homotopy from f to g. Further, if Hi holds xi fixed, then H holds x fixed. � Corollary 5.4. Let (Xi, κi) and (Yi, λi) be digital images, 1 ≤ i ≤ v. Then X = Πvi=1Xi and Y = Π v i=1Yi are (× v i=1κi, × v i=1λi)-(pointed) homotopy equiv- alent if and only if for each i, (Xi, κi) and (Yi, λi) are (pointed) homotopy equivalent. Proof. This follows from Theorem 5.3. � 5.3. Lexicographic adjacency. Theorem 5.5. Let (Xi, κi) be digital images for 1 ≤ i ≤ v. Let X = Π v i=1Xi. If there is a smallest index k such that |Xk| > 1, then (X, L(κ1, . . . , κv)) and (Xk, κk) have the same pointed homotopy type. Proof. For each i 6= k, let xi ∈ Xi. Let Ik : Xk → X be the injection of Proposition 3.19. By choice of k, Ik is (κk, L(κ1, . . . , κv))-continuous. Also by choice of k, the projection map pk : (X, L(κ1, . . . , κv)) → (Xk, κk) is continu- ous. Notice pk ◦ Ik = 1Xk . Also, the function H : X × [0, 1]Z → X defined for p = (y1, . . . , yv) ∈ X with yi ∈ Xi by H(p, t) =        p if t = 0; (y1, x2, . . . , xv) if t = 1 and k = 1; (x1, . . . , xk−1, yk, xk+1, . . . , xv) if t = 1 and 1 < k < v; (x1, . . . , xv−1, yv) if t = 1 and k = v, is easily seen from the choice of k to be a homotopy from 1X to Ik ◦ pk that holds fixed the point (x1, . . . , xv). The assertion follows. � Corollary 5.6. Let (X, κ) and (Y, λ) be digital images of different pointed homotopy types. If |X| > 1 and |Y | > 1, then (X × Y, L(κ, λ)) and (Y × X, L(λ, κ)) have different pointed homotopy types. Proof. This follows immediately from Theorem 5.5. � Corollary 5.7. Let (Xi, κi) and (Yi, λi) be digital images, 1 ≤ i ≤ v. Let X = Πvi=1Xi, Y = Π v i=1Yi. Suppose there exist a smallest index j such that |Xj| > 1, and a smallest index k such that |Yk| > 1. If (Xj, κj) and (Yk, κk) have the same (pointed) homotopy type, then (X, L(κ1, . . . , κv)) and (Y, L(λ1, . . . , λv)) have the same (pointed) homotopy type. c© AGT, UPV, 2018 Appl. Gen. Topol. 19, no. 1 38 Product adjacencies in digital topology Proof. By Theorem 5.5, (X, L(κ1, . . . , κv)) and (Xj, κj) have the same pointed homotopy type, and (Yk, λk) and (Y, L(λ1, . . . , λv)) have the same pointed homotopy type. Since we also have assumed (Xj, κj) and (Yk, λk) have the same (pointed) homotopy type, the assertion follows from the transitivity of (pointed) homotopy type. � 6. Retractions Definition 6.1 ([2, 3]). Let Y ⊂ (X, κ). A (κ, κ)-continuous function r : X → Y is a retraction, and A is a retract of X, if r(y) = y for all y ∈ Y . Theorem 6.2 ([12]). Let Ai ⊂ (Xi, κi), i ∈ {1, . . . , v}. Then Ai is a retract of Xi for all i if and only if Π v i=1Ai is a retract of (Π v i=1Xi, NPv(κ1, . . . , κv)). 6.1. Tensor product adjacency. The following example shows that one of the assertions obtained by using the tensor product adjacency rather than NPv in Theorem 6.2 is not generally valid. Example 6.3. Let X = {(0, 0), (1, 0), (1, 1)} ⊂ (Z2, c2). Observe that X ′ = {(0, 0), (1, 0)} is a c2-retract of X, and {0} is a c1-retract of [0, 1]Z. However, X′ × {0} is not a T (c2, c1)-retract of X × [0, 1]Z. Proof. Note X × [0, 1]Z is T (c2, c1)-connected, since (0, 0, 0), (1, 0, 1), (1, 1, 0), (0, 0, 1), (1, 0, 0), (1, 1, 1) is a listing of its points in a T (c2, c1)-path; but X ′ × {0} = {(0, 0, 0), (1, 0, 0)} is not T (c2, c1)-connected. The assertion follows. � The question of whether Πvi=1Ai being a retract of (Π v i=1Xi, T (κ1, . . . , κv)) implies Ai is a κi-retract of Xi, for all i, is unknown at the current writing. 6.2. Cartesian product adjacency. For the Cartesian product adjacency, we have the following analog of Theorem 6.2. Theorem 6.4. Suppose Ai ⊂ (Xi, κi). Let X = Π v i=1Xi, A = Π v i=1Ai. Then there is a retraction ri : Xi → Ai, 1 ≤ i ≤ v if and only if there is a retraction r : (X, ×vi=1κi) → (A, × v i=1κi). Proof. Suppose there is a retraction ri : Xi → Ai, 1 ≤ i ≤ v. Let r = Π v i=1ri : X → A. Clearly r(x) ∈ A for all x ∈ X, and r(a) = a for all a ∈ A. By Theorem 3.17, r is continuous. Therefore, r is a retraction. Conversely, suppose there exists a retraction r : (X, ×vi=1κi) → (A, × v i=1κi). Let ri = pi ◦ r ◦ Ii : (Xi, κi) → (Ai, κi), where Ii is the injection of Proposi- tion 3.19 and the xi of Proposition 3.19 satisfies xi ∈ Ai. Since composition preserves continuity, Theorem 3.18 and Proposition 3.19 imply ri is continuous. Further, for ai ∈ Ai we clearly have ri(ai) = ai. Thus, ri is a retraction. � c© AGT, UPV, 2018 Appl. Gen. Topol. 19, no. 1 39 L. Boxer 6.3. Lexicographic adjacency. For the lexicographic adjacency, we do not have an analog of Theorem 6.2, as shown by the following example. Example 6.5. {0} is a c1-retract of [0, 1]Z and [1, 4]Z is a c1-retract of [0, 5]Z. However, A = {0} × [1, 4]Z is not an L(c1, c1)-retract of X = [0, 1]Z × [0, 5]Z. Proof. We give a proof by contradiction. Suppose there is an L(c1, c1)-retraction r : [0, 1]Z × [0, 5]Z → {0} × [1, 4]Z. Notice p = (0, 1) ↔L(c1,c1) (1, 5) = p ′. Since r(p) = p, the continuity of r requires that r(p′) = p or r(p′) ↔L(c1,c1) p, hence r(p′) ∈ {p, (0, 2)}. But also p′ ↔L(c1,c1) (0, 4) = q, and since r(q) = q, the continuity of r similarly requires that r(p′) ↔L(c1,c1) {q, (0, 3)}. Therefore, r(p′) ∈ {p, (0, 2)} ∩ {q, (0, 3)} = ∅. Since this is impossible, no such retraction r can exist. � 7. Approximate fixed point property Some material in this section is quoted or paraphrased from [9, 10]. In both topology and digital topology, • a fixed point of a continuous function f : X → X is a point x ∈ X satisfying f(x) = x; • if every continuous f : X → X has a fixed point, then X has the fixed point property (FPP). However, a digital image X has the FPP if and only if X has a single point [10]. Therefore, it turns out that the approximate fixed point property is more inter- esting for digital images. Definition 7.1 ([10]). A digital image (X, κ) has the approximate fixed point property (AFPP) if every continuous f : X → X has an approximate fixed point, i.e., a point x ∈ X such that f(x) is equal or κ-adjacent to x. The following is a minor generalization of Theorem 5.10 of [10]. Theorem 7.2 ([9]). Let (Xi, κi) be digital images, 1 ≤ i ≤ v. Then for any u ∈ Z such that 1 ≤ u ≤ v, if (Πvi=1Xi, NPu(κ1, . . . , κv)) has the AFPP then (Xi, κi) has the AFPP for all i. Determining whether analogs of Theorem 7.2 for the tensor product adja- cency, or for the Cartesian product adjacency, are generally true, appear to be difficult problems. The following examples show that the analogs of converses to Theorem 7.2 for the tensor product adjacency and for the Cartesian product adjacency are not generally true. Example 7.3. Although ([0, 1]Z, c1) has the AFPP [25], ([0, 1] 2 Z , T (c1, c1)) does not have the AFPP. c© AGT, UPV, 2018 Appl. Gen. Topol. 19, no. 1 40 Product adjacencies in digital topology Proof. Consider the function f : [0, 1]2 Z → [0, 1]2 Z defined by f(a, b) = (1 − a, b), i.e., f(0, 0) = (1, 0), f(0, 1) = (1, 1), f(1, 0) = (0, 0), f(1, 1) = (0, 1). One can easily check that f is continuous and has no approximate fixed point when the T (c1, c1) adjacency is used. � Example 7.4. Although ([0, 1]Z, c1) has the AFPP, ([0, 1] 2 Z , c1 × c1) does not have the AFPP. Proof. Consider the function f : [0, 1]2 Z → [0, 1]2 Z defined by f(a, b) = (1−a, 1− b), i.e., f(0, 0) = (1, 1), f(0, 1) = (1, 0), f(1, 0) = (0, 1), f(1, 1) = (0, 0). One can easily check that f is continuous and has no approximate fixed point when the c1 × c1 adjacency is used. � We have the following. Theorem 7.5. Let (Xi, κi) be digital images, 1 ≤ i ≤ v. Suppose there is a smallest index k such that Xk is κk-connected and |Xk| > 1. If the product (Πvi=1Xi, L(κ1, . . . , κv)) has the AFPP property, then (Xk, κk) has the AFPP property. Proof. Let X = Πvi=1Xi. Suppose the product (X, L(κ1, . . . , κv)) has the AFPP property. Let g : Xk → Xk be κ-continuous. Let xi ∈ Xi. Notice this means Xi = {xi} for i < k. Let X = Πvi=1Xi. Let G : X → X be defined by G(y1, . . . , yv) =    (g(y1), x2, . . . , xv) if k = 1; (x1, . . . , xk−1, g(yk), xk+1, . . . , xv) if 1 < k < v; (x1, . . . , xv−1, g(yv)) if k = v. Since g is κk-continuous, our choice of k implies G is L(κ1, . . . , κv)-continuous. By hypothesis, there is a p = (y′1, . . . , y ′ v) ∈ X with y ′ i ∈ Xi such that G(p) = p or G(p) ↔ p. Therefore, either g(y′k) = pk(G(p)) = pk(p) = y ′ k or g(y ′ k) ↔κk y ′ k. Thus, y′k is an approximate fixed point for g. � 8. Multivalued functions We study various product adjacencies with respect to properties of multi- valued functions. The following has an elementary proof. Proposition 8.1. Let f : (X, κ) → (Y, λ) be a single-valued function between digital images. Then the following are equivalent. • f is continuous. • As a multivalued function, f has weak continuity. • As a multivalued function, f has strong continuity. c© AGT, UPV, 2018 Appl. Gen. Topol. 19, no. 1 41 L. Boxer For multivalued functions Fi : Xi ⊸ Yi, 1 ≤ i ≤ v, define the product multivalued function Πvi=1Fi : Π v i=1Xi ⊸ Π v i=1Yi by (Πvi=1Fi)(x1, . . . , xv) = Π v i=1Fi(xi). 8.1. Weak continuity. For NPv, we have the following results. Theorem 8.2 ([9]). Let Fi : (Xi, κi) ⊸ (Yi, λi) be multivalued functions for 1 ≤ i ≤ v. Let X = Πvi=1Xi, Y = Π v i=1Yi, and F = Π v i=1Fi : (X, NPv(κ1, . . . , κv)) ⊸ (Y, NPv(λ1, . . . , λv)). Then F has weak continuity if and only if each Fi has weak continuity. For the tensor product, we have the following. Theorem 8.3. For each index i such that 1 ≤ i ≤ v, let fi : (Xi, κi) ⊸ (Yi, λi) be a multivalued map between digital images. Let X = Πvi=1Xi, Y = Π v i=1Yi. If the product multivalued map f = Πvi=1fi : (X, T (κ1, . . . , κv)) ⊸ (Y, T (λ1, . . . , λv)) has weak continuity, then for each i, fi has weak continuity. Proof. For all indices i, let xi ↔κi x ′ i in Xi. Then, in X, we have p = (x1, . . . , xv) ↔T (κ1,...,κv) p ′ = (x′1, . . . , x ′ v). The weak continuity of f implies f(p) and f(p′) are adjacent subsets of (Y, T (λ1, . . . , λv)). Therefore, there exist y ∈ f(p) and y′ ∈ f(p′) such that y = y′ or y ↔T (λ1,...,λv) y ′. Now, y = (y1, . . . , yv) where yi ∈ fi(xi), and y ′ = (y′1, . . . , y ′ v) where y ′ i ∈ fi(x ′ i). If y = y ′ then we have yi = y ′ i for all indices i. If y ↔T (λ1,...,λv) y ′ then we have yi ↔λi y ′ i for all indices i. In either case, we have for all i that fi(xi) and fi(x ′ i) are adjacent subsets of Yi. It follows that each fi has weak continuity. � The converse of Theorem 8.3 is not generally true, as shown by the following. Example 8.4. Let f and g be the single-valued functions of Example 3.11. By Proposition 8.1, f and g have weak continuity. However, Example 3.11 shows that f × g is not (T (c1, c1), T (c1, c1))-continuous, so by Proposition 8.1, f × g does not have (T (c1, c1), T (c1, c1))-weak continuity. For the Cartesian product adjacency, we have the following. Theorem 8.5. Let fi : (Xi, κi) ⊸ (Yi, λi) be multivalued maps between dig- ital images, 1 ≤ i ≤ v. Let X = Πvi=1Xi, Y = Π v i=1Yi. Then the product multivalued map f = Πvi=1fi : (X, × v i=1κi) ⊸ (Y, × v i=1λi) has weak continuity if and only if for each i, fi has weak continuity. c© AGT, UPV, 2018 Appl. Gen. Topol. 19, no. 1 42 Product adjacencies in digital topology Proof. Suppose f has weak continuity. Let xi ↔κi x ′ i in Xi. Let x = (x1, . . . , xv) ∈ X, x′ = (x1, . . . , xj−1, x ′ j, xj+1, . . . , xv) ∈ X for some index j. We have x ↔×v i=1 κi x ′. Therefore, there exist y = (y1, . . . , yv) ∈ f(x) = Π v i=1fi(xi), y′ = (y′1, . . . , y ′ v) ∈ f(x ′) = Π j−1 i=1 fi(xi) × fj(xj) × Π v i=j+1fi(xi) such that y ↔×v i=1 λi y ′. Therefore, we have yj ∈ fj(xj), y ′ j ∈ fj(x ′ j), and yj = y ′ j or yj ↔λj y ′ j. Thus, fj has weak continuity. Suppose each fi has weak continuity. Let p ↔×v i=1 κi p ′ in X, where p = (x1, . . . , xv), p ′ = (x′1, . . . , x ′ v), xi, x ′ i ∈ Xi, and, from the definition of the ×vi=1κi adjacency, there is one index j such that xj ↔κj x ′ j and for all in- dices i 6= j, xi = x ′ i and therefore fi(xi) = fi(x ′ i). Since fj has weak continuity, there exist yj ∈ fj(xj) and y ′ j ∈ fj(x ′ j) such that yj = y ′ j or yj ↔λj y ′ j. For i 6= j we can take yi ∈ fi(xi). Then y = (y1, . . . , yv) and y′ = (y1, . . . , yj−1, y ′ j, yj+1, . . . , yv) are equal or × v i=1λi-adjacent, and we have y ∈ f(p), y′ ∈ f(p′). Therefore, f has weak continuity. � For the lexicographic adjacency, Example 8.10 below shows there is no gen- eral product property for weak continuity, and Example 8.11 below shows there is not a general factor property for weak continuity. 8.2. Strong continuity. Theorem 8.6 ([9]). Let Fi : (Xi, κi) ⊸ (Yi, λi) be multivalued functions for 1 ≤ i ≤ v. Let X = Πvi=1Xi, Y = Π v i=1Yi, F = Π v i=1Fi : (X, NPv(κ1, . . . , κv)) ⊸ (Y, NPv(λ1, . . . , λv)). Then F has strong continuity if and only if each Fi has strong continuity. For the tensor product adjacency, we have the following. Theorem 8.7. Let fi : (Xi, κi) ⊸ (Yi, λi) be multivalued maps between digital images, 1 ≤ i ≤ v. Let X = Πvi=1Xi, Y = Π v i=1Yi. If the product multivalued map f = Πvi=1fi : (X, T (κ1, . . . , κv)) ⊸ (Y, T (λ1, . . . , λv)) has strong continuity, then for each i, fi has strong continuity. Proof. Let xi ↔κi x ′ i in Xi. Let p = (x1, . . . , xv) and p ′ = (x′1, . . . , x ′ v). Note p ↔T (κ1,...,κv) p ′ in X. Since f has strong continuity, for every q = (y1, . . . , yv) ∈ f(p) = Π v i=1fi(xi) where yi ∈ fi(xi), there exists q ′ = (y′1, . . . , y ′ v) ∈ f(p′) = Πvi=1fi(x ′ i) where y ′ i ∈ fi(x ′ i) such that either q = q ′ or q ↔T (λ1,...,λv) q′; and therefore yi = y ′ i for all i or yi ↔λi y ′ i for all i. Also, for every r′ = (r′1, . . . , r ′ v) ∈ f(p ′) where r′i ∈ fi(x ′ i), there exists r = (r1, . . . , rv) ∈ f(p) where ri ∈ fi(xi) such that either r = r ′ or r ↔T (λ1,...,λv) r ′; and therefore ri = r ′ i for all i or ri ↔λi r ′ i for all i. Thus fi has (κi, λi)-strong continuity. � The converse of Theorem 8.7 is not generally true, as shown by the following. c© AGT, UPV, 2018 Appl. Gen. Topol. 19, no. 1 43 L. Boxer Example 8.8. Let f1 : ([0, 1]Z, c1) ⊸ ([0, 1]Z, c1) be defined by f1(x) = {0}. Let f2 : ([0, 1]Z, c1) ⊸ ([0, 1]Z, c1) be defined by f2(x) = {x}. Then f1 and f2 both have strong continuity. However, f1×f2 does not have (T (c1, c1), T (c1, c1))- strong continuity. Proof. It is easily seen that f1 and f2 both have strong continuity. However, in Example 8.4, we showed that f1 × f2 does not have (T (c1, c1), T (c1, c1))- weak continuity. Therefore, f1 × f2 does not have (T (c1, c1), T (c1, c1))-strong continuity. � Theorem 8.9. Let fi : (Xi, κi) ⊸ (Yi, λi) be multivalued maps between dig- ital images, 1 ≤ i ≤ v. Let X = Πvi=1Xi, Y = Π v i=1Yi. Then the product multivalued map f = Πvi=1fi : (X, × v i=1κi) ⊸ (Y, × v i=1λi) has strong continuity if and only if for each i, fi has strong continuity. Proof. Suppose f has strong continuity. Let xi ↔κi x ′ i in Xi. Then p = (x1, . . . , xv) ↔×v i=1 κi (x1, . . . , xj−1, x ′ j, xj+1, . . . , xv) = p ′ in X, for some index j. Since f has strong continuity, we must have that for every q = (q1, . . . , qv) ∈ f(p) there exists q ′ = (q′1, . . . , q ′ v) ∈ f(p ′) such that q = q′ or q ↔×v i=1 λi q ′, so qi = q ′ i or qi ↔λi q ′ i; and for every r ′ = (r′1, . . . , r ′ v) ∈ f(p′) there exists r = (r1, . . . , rv) ∈ f(p) such that r = r ′ or r ↔×v i=1 λi r ′, so ri = r ′ i or ri ↔λi r ′ i. Therefore, fi has strong continuity. Suppose for each i, fi has strong continuity. Let p = (x1, . . . , xv) and p ′ = (x′1, . . . , x ′ v) with xi, x ′ i ∈ Xi be such that p ↔×vi=1κi p ′. Then for some index j, xj ↔κj x ′ j and for all indices i 6= j, xi = x ′ i. Therefore, i 6= j implies there exists qi ∈ fi(xi) = fi(x ′ i); and since fj has strong continuity, for every qj ∈ fj(xj) there exists q ′ j ∈ fj(x ′ j) such that qj = q ′ j or qj ↔λj q ′ j. Let q′ = (q1, . . . , qj−1, q ′ j, qj+1, . . . , qv). Then q = (q1, . . . , qv) = q ′ or q ↔×v i=1 λi q ′ with q ∈ f(p), q′ ∈ f(p′). Similarly, for every r′ ∈ f(p′) there exists r ∈ f(p) such that r = r′ or r ↔×v i=1 λi r ′. Thus, f has strong continuity. � For the lexicographic adjacency, the following shows there is not a general product property for weak or strong continuity. Example 8.10. Let f1 : ([0, 1]Z, c1) ⊸ ([0, 1]Z, c1) be the multivalued function f1(x) = {0}. Let f2 : ({0, 2}, c1) ⊸ ({0, 2}, c1) be the function f2(x) = {x}. Then f1 and f2 have weak continuity and strong continuity, but f1 × f2 lacks both (L(c1, c1), L(c1, c1))-weak continuity and (L(c1, c1), L(c1, c1))-strong con- tinuity. Proof. It is easy to see that f1 and f2 have weak continuity and strong conti- nuity, and that p = (0, 0) ↔L(c1,c1) (1, 2) = p ′. However (f1 × f2)(p) = {(0, 0)} and (f1 × f2)(p ′) = {(0, 2)}, are not L(c1, c1)-adjacent, so f1 ×f2 lacks (L(c1, c1), L(c1, c1))-weak continuity and therefore lacks (L(c1, c1), L(c1, c1))-strong continuity. � c© AGT, UPV, 2018 Appl. Gen. Topol. 19, no. 1 44 Product adjacencies in digital topology For the lexicographic adjacency, the following shows there is not a general factor property for weak or strong continuity. Example 8.11. Let f1 : ([0, 1]Z, c1) ⊸ ([0, 1]Z, c1) be the multivalued function f1(x) = [0, 1]Z. Let f2 : ([0, 1]Z, c1) ⊸ ({0, 2}, c1) be the multivalued function f2(x) = {2x}. Then f1 × f2 : [0, 1] 2 Z ⊸ [0, 1]Z × {0, 2} has (L(c1, c1), L(c1, c1))- weak and (L(c1, c1), L(c1, c1))-strong continuity, although f2 lacks both weak and strong continuity. Proof. It is easy to see that f2 lacks weak and strong continuity. Since (f1 × f2)(0, 0) = (f1 × f2)(1, 0) = {(0, 0), (1, 0)}, (f1 × f2)(0, 1) = (f1 × f2)(1, 1) = {(0, 2), (1, 2)}, it follows easily that f1 × f2 has both (L(c1, c1), L(c1, c1))-weak continuity and (L(c1, c1), L(c1, c1))-strong continuity. � 8.3. Continuous multifunctions. Lemma 8.12 ([9]). Let X ⊂ Zm, Y ⊂ Zn. Let F : (X, ca) ⊸ (Y, cb) be a continuous multivalued function. Let f : (S(X, r), ca) → (Y, cb) be a continuous function that induces F. Let s ∈ N. Then there is a continuous function fs : (S(X, rs), ca) → (Y, cb) that induces F. For the NPv adjacency, we have the following. Theorem 8.13 ([9]). Given multivalued functions Fi : (Xi, cai) ⊸ (Yi, cbi), 1 ≤ i ≤ v, each Fi is continuous if and only if the product multivalued function Πvi=1Fi : (Π v i=1Xi, NPv(ca1, . . . , cav)) ⊸ (Π v i=1Yi, NPv(cb1, . . . , cbv )) is continuous. For the tensor product, since a single-valued function can be considered as multivalued, Example 3.11 shows there is no general product rule for the continuity of multivalued functions. However, we have the following. Theorem 8.14. Let Fi : (Xi, cai) ⊸ (Yi, cbi) be a continuous multivalued function between digital images, 1 ≤ i ≤ v. Let X = Πvi=1Xi, Y = Π v i=1Yi, F = Πvi=1Fi : X ⊸ Y . If for some positive integer r and for all i there is a con- tinuous locally one-to-one function fi : (S(Xi, r), cai) → (Yi, cbi) that generates Fi, then F is (T (ca1, . . . , cav ), T (cb1, . . . , cbv ))-continuous and is generated by a function that is locally one-to-one. Proof. Let f = Πvi=1fi : Π v i=1S(Xi, r) → Y . It follows from Theorem 3.13 that f is (T (ca1, . . . , cav), T (cb1, . . . , cbv ))-continuous. Further, given q ∈ F(p) where p = (x1, . . . , xv) for xi ∈ Xi and q = (y1, . . . , yv) where yi ∈ Fi(xi), there exists x′i ∈ S({xi}, r) ⊂ S(Xi, r) such that fi(x ′ i) = yi. Therefore, f(x ′ 1, . . . , x ′ v) = q. For w ↔T (ca1 ,...,cav ) w ′ in S(X, r) = Πvi=1S(Xi, r), we have w = (w1, . . . , wv) and w′ = (w′1, . . . , w ′ v), where wi, w ′ i ∈ S(Xi, r) and wi ↔cai w ′ i. Since fi is locally one-to-one and continuous, we have fi(wi) ↔cbi fi(w ′ i). It follows that c© AGT, UPV, 2018 Appl. Gen. Topol. 19, no. 1 45 L. Boxer f(w1, . . . , wv) ↔T (cb1 ,...,cbv ) f(w ′ 1, . . . , w ′ v). This allows us to conclude that f is (T (ca1, . . . , cav), T (cb1, . . . , cbv ))-continuous. Thus, f generates F . Let p′ = (x′1, . . . , x ′ v) ↔T (κ1,...,κv) p in X, where x ′ i ∈ Xi. Since fi is locally one-to-one, fi(xi) ↔λi fi(x ′ i) for all i. Therefore, f(p) ↔T (λ1,...,λv) f(p ′), so f is locally one-to-one. � Deciding whether the converse of Theorem 8.14 is true appears to be a difficult problem. For the Cartesian product adjacency, we have the following. Theorem 8.15. Let Fi : (Xi, κi) ⊸ (Yi, λi) be a multivalued function between digital images, where κi = cai, λi = cbi, 1 ≤ i ≤ v. Let X = Π v i=1Xi, Y = Πvi=1Yi, F = Π v i=1Fi : X ⊸ Y . If each Fi is continuous, then F is (×vi=1κi, × v i=1λi)-continuous. Proof. Suppose each Fi is continuous. By Lemma 8.12, there exists r ∈ N and generating functions fi : S(Xi, r) → Yi of Fi. We wish to show that f = Πvi=1fi generates F . Suppose p ↔×vi=1κi p ′ in S(X, r). Then p = (x1, . . . , xv) and p ′ = (x′1, . . . , x ′ v) where xi, x ′ i ∈ S(Xi, r) and xi = x ′ i for all but one index j, with xj ↔κj x ′ j. Since each fi is (κi, λi)- continuous, we have fj(xj) = fj(x ′ j) or fj(xj) ↔λj fj(x ′ j) and for all indices i 6= j we have fi(xi) = fi(x ′ i). Thus we have f(p) = f(p ′) or f(p) ↔×v i=1 λi f(p′). Thus, f is (×vi=1κi, × v i=1λi)-continuous. Let y = (y1, . . . , yv) ∈ F(X), where yi ∈ Yi. Then there exists xi ∈ S(Xi, r) such that fi(xi) = yi. For p = (x1, . . . , xv), we have f(p) = (y1, . . . , yv). Thus, f generates F , so F is continuous. � Deciding whether the converse of Theorem 8.15 is true appears to be a difficult problem. For the lexicographic adjacency, there is no general product rule for the conti- nuity of multivalued functions, as shown in Example 3.22 (since a single-valued function can be regarded as multivalued). However, we have the following. Theorem 8.16. Let Fi : (Xi, κi) ⊸ (Yi, λi) be a continuous multivalued func- tion between digital images, 1 ≤ i ≤ v. Let X = Πvi=1Xi, Y = Π v i=1Yi, F = Πvi=1Fi : X ⊸ Y . If each Fi is generated by a function fi : (S(Xi, r), κi) → Yi that is locally one-to-one, then F is (L(κ1, . . . , κv), L(λ1, . . . , λv))-continuous. Proof. By Theorem 3.21, the single-valued function f = Πvi=1fi : Π v i=1S(Xi, r) → Y is (L(κ1, . . . , κv), L(λ1, . . . , λv))-continuous. Further, given y = (y1, . . . , yv) ∈ F(X) with yi ∈ Yi, there exist x ′ i ∈ S({xi}, r) ⊂ S(Xi, r) such that fi(x ′ i) = yi. Therefore, y = f(x′1, . . . , x ′ v) ∈ F(x1, . . . , xv). Therefore, f generates F , and the assertion follows. � The paper [16] has several results concerning the following notions. Definition 8.17 ([16]). Let (X, κ) ⊂ Zn be a digital image and Y ⊂ X. We say that Y is a κ-retract of X if there exists a κ-continuous multivalued function F : X ⊸ Y (a multivalued κ-retraction) such that F(y) = {y} if y ∈ Y . c© AGT, UPV, 2018 Appl. Gen. Topol. 19, no. 1 46 Product adjacencies in digital topology We generalize Theorem 6.2 as follows. Theorem 8.18 ([9]). For 1 ≤ i ≤ v, let Ai ⊂ (Xi, κi) ⊂ Z ni. Suppose Fi : Xi ⊸ Ai is a continuous multivalued function for all i. Then Fi is a multivalued κi-retraction for all i if and only if F = Π v i=1Fi : Π v i=1Xi ⊸ Π v i=1Ai is a multivalued NPv(κ1, . . . , κv)-retraction. For the Cartesian product adjacency, we have the following. Theorem 8.19. Let ri : Xi ⊸ Ai be multivalued κi-retractions, 1 ≤ i ≤ v. Let X = Πvi=1Xi, A = Π v i=1Ai, r = Π v i=1ri : X ⊸ A. Then r is a × v i=1κi- multivalued retraction. Proof. Since ri is a multivalued retraction, we must have that ri(Xi) = Ai and ri(ai) = {ai} for all ai ∈ Ai. Therefore, r(X) = A and r(a) = {a} for all a ∈ A. By Theorem 8.15, r is continuous, and therefore is a multivalued retraction. � 8.4. Connectivity preserving multifunctions. Theorem 8.20 ([9]). Let fi : (Xi, κi) ⊸ (Yi, λi) be a multivalued function between digital images, 1 ≤ i ≤ v. Then the product map Πvi=1fi : (Π v i=1Xi, NPv(κ1, . . . , κv)) ⊸ (Π v i=1Yi, NPv(λ1, . . . , λv)) is a connectivity preserving multifunction if and only if each fi is a connectivity preserving multifunction. The tensor product adjacency does not yield a similar result, as shown in the following. Example 8.21. Consider {0} ⊂ Z, [0, 1]Z ⊂ Z. The multivalued function f : ({0}, c1) ⊸ ([0, 1]Z, c1) defined by f(0) = [0, 1]Z is connectivity preserv- ing. However, f × f : {0}2 = {(0, 0)} ⊸ [0, 1]2 Z is not (T (c1, c1), T (c1, c1))- connectivity preserving. Proof. This follows from the observations that {(0, 0)} has a single point, hence must be T (c1, c1)-connected; but, by Example 4.3, (f × f)(0, 0) = [0, 1] 2 Z is not T (c1, c1)-connected. � However, we have the following. Theorem 8.22. Let fi : (Xi, κi) ⊸ (Yi, λi) be multivalued functions, 1 ≤ i ≤ v. Let X = Πvi=1Xi, Y = Π v i=1Yi. Suppose f = Π v i=1fi : X ⊸ Y is (T (κ1, . . . , κv), T (λ1, . . . , λv))-connectivity preserving. Then each fi is connec- tivity preserving. Proof. Let p = (x1, . . . , xv) ∈ X, where xi ∈ Xi. By assumption, f(p) = Πvi=1fi(xi) is T (λ1, . . . , λv)-connected. From Theorem 4.2, it follows that fi(xi) is λi-connected. Suppose x′i ↔κi xi in Xi. Then p ′ = (x′1, . . . , x ′ v) ↔T (κ1,...,κv) p. Since f is connectivity preserving, f(p′) and f(p) are T (λ1, . . . , λv)-adjacent subsets of c© AGT, UPV, 2018 Appl. Gen. Topol. 19, no. 1 47 L. Boxer Y . This implies there exist q′ = (y′1, . . . , y ′ v) ∈ f(p ′), q = (y1, . . . , yv) ∈ f(p) such that q′ ↔T (κ1,...,κv) q or q ′ = q. Therefore, for each index i, y′i ↔λi yi or y′i = yi. Since y ′ i ∈ fi(x ′ i) and yi ∈ fi(xi), we have that fi(x ′ i) and fi(xi) are λi-adjacent subsets of Yi. From Theorem 2.18, fi is connectivity preserving. � For the Cartesian product adjacency, we have the following. Theorem 8.23. Let (Xi, κi) and (Yi, λi) be digital images, for 1 ≤ i ≤ v. Let fi : Xi ⊸ Yi be a multivalued function. Let f = Π v i=1fi : X = Π v i=1Xi ⊸ Y = Πvi=1Yi be the product function. Then f is (× v i=1κi, × v i=1λi)-connectivity preserving if and only if each fi is connectivity preserving. Proof. Suppose f is connectivity preserving. Let p = (x1, . . . , xv) ∈ X, where xi ∈ Xi. Then f(p) = Π v i=1fi(xi) is × v i=1λi-connected. By Theorem 3.18, fi(xi) = pi(f(p)) is λi-connected. For any given index k, let xk ↔κk x ′ k in Xk. For all indices i 6= k, let xi ∈ Xi. Then p = (x1, . . . , xv) and p ′ = (x1, . . . , xk−1, x ′ k , xk+1, . . . , xv) are ×vi=1κi-adjacent. Since f is connectivity preserving, f(p) and f(p ′) are ×vi=1λi- adjacent subsets of Y . Therefore, Theorem 3.18 implies fk(xk) = pk(f(p)) and fk(x ′ k ) = pk(f(p ′)) are λk-adjacent subsets of Yk. It follows from Theorem 2.18 that fk is connectivity preserving. Since k was an arbitrarily selected index, fi is connectivity preserving for all i. Now suppose each fi is connectivity preserving. Let p = (x1, . . . , xv) ∈ X where xi ∈ Xi. Then f(p) = Π v i=1fi(xi) is, by Theorem 4.4, × v i=1λi-connected. Suppose p ↔×v i=1 λi p ′ in X. Then for some index k, xk ↔κi x ′ k in Xk and for i 6= k there exist xi ∈ Xi such that p = (x1, . . . , xv), p ′ = (x1, . . . , xk−1, x ′ k, xk+1, . . . , xv). Since fk is connectivity preserving, there exist yk ∈ fk(xk) and y ′ k ∈ fk(x ′ k ) such that yk ↔λk y ′ k or yk = y ′ k. For i 6= k, let yi ∈ fi(xi). Then q = (y1, . . . , yv) ∈ f(p) and q ′ = (y1, . . . , yk−1, y ′ k , yk+1, . . . , yv) ∈ f(p ′) are ×vi=1λi- adjacent or equal. Therefore, f(p) and f(q) are ×vi=1λi-adjacent subsets of Y . It follows from Theorem 2.18 that f is connectivity preserving. � For lexicographic adjacency, • Example 3.22 shows that there is no product property for connectivity preservation; and • there is no factor property for connectivity preservation, as the follow- ing example shows. Example 8.24. Let f1 : ({0}, c1) ⊸ ([0, 1]Z, c1) be the multivalued function f1(0) = [0, 1]Z. Let f2 : ({0}, c1) ⊸ ({0, 2}, c1) be the multivalued function f2(0) = {0, 2}. Then f = f1 × f2 : {0} 2 = {(0, 0)} ⊸ [0, 1]Z × {0, 2} is (L(c1, c1), L(c1, c1))-connectivity preserving, but f2 is not (c1, c1)-connectivity preserving. c© AGT, UPV, 2018 Appl. Gen. Topol. 19, no. 1 48 Product adjacencies in digital topology Proof. This follows from the observations that the single point (0, 0) is con- nected, and f(0, 0) = [0, 1]Z × {0, 2} is L(c1, c1)-connected. � 9. Shy maps We have the following. Theorem 9.1. Let f : (X, κ) → (Y, λ) be a shy map of digital images. Then f is an isomorphism if and only if f is locally one-to-one. Proof. It is obvious that if f is an isomorphism, then f is locally one-to-one. To show the converse, we argue as follows. Since f is shy, we know f is a continuous surjection. To show f is one-to-one, suppose there exist x, x′ ∈ X such that y = f(x) = f(x′) ∈ Y . Since f is shy, f−1(y) is κ-connected. Therefore, if x 6= x′ then there is a path of distinct points P = {xi} m i=1 ⊂ f −1(y) such that x = x1, xi ↔ xi+1 for 1 ≤ i < m, and xm = x ′. But since f is locally one-to-one, f|N∗κ(x) is one- to-one, so f(x2) 6= f(x), contrary to the assumption P ⊂ f −1(y). Therefore, we must have x = x′, so f is one-to-one. Since f is one-to-one, f−1 is one-to-one. Since f is shy, given y ↔ y′ in Y , f−1({y, y′}) is connected. Thus, f−1 is continuous. This completes the proof that f is an isomorphism. � The following generalizes a result of [8]. Theorem 9.2 ([9]). Let fi : (Xi, κi) → (Yi, λi) be a continuous surjection between digital images, 1 ≤ i ≤ v. Then the product map Πvi=1fi : (Π v i=1Xi, NPv(κ1, . . . , κv)) → (Π v i=1Yi, NPv(λ1, . . . , λv)) is shy if and only if each fi is a shy map. For the tensor product, we have the following. Theorem 9.3. Let fi : (Xi, κi) → (Yi, λi) be a surjection between digital images, 1 ≤ i ≤ v. Let X = Πvi=1Xi, Y = Π v i=1Yi. If the product function f = Πvi=1fi : (X, T (κ1, . . . , κv)) → (Y, T (λ1, . . . , λv)) is shy, then fi is shy for each i. Proof. Since f is shy, it is continuous, so by Theorem 3.10, each fi is continuous. Clearly, each fi is a surjection. Let yi ∈ Yi. Let y = (y1, . . . , yv) ∈ Y . Since f is shy, f −1(y) = Πvi=1f −1 i (yi) is T (κ1, . . . , κv)-connected. By Theorem 4.2, fi(yi) is κi-connected. Let y′i ↔λi yi in Yi. Then y ′ = (y′1, . . . , y ′ v) ↔T (λ1,...,λv) y. Since f is shy, f−1({y, y′}) = f−1({y}) ∪ f−1({y′}) = Πvi=1f −1 i (yi) ∪ Π v i=1f −1 i (y ′ i) is T (κ1, . . . , κv)-connected. By Theorem 3.15, pi(f −1({y, y′})) = f−1i (yi) ∪ f −1 i (y ′ i) is κi-connected. From Definition 2.30, we conclude that fi is a shy map. � c© AGT, UPV, 2018 Appl. Gen. Topol. 19, no. 1 49 L. Boxer The converse to Theorem 9.3 is not generally true, as shown by the following. Example 9.4. Let f1 : ([0, 1]Z, c1) → ({0}, c1) be the function f1(x) = 0. Let f2 : ([0, 1]Z, c1) → ([0, 1]Z, c1) be the function f2(x) = x. Then f1 and f2 are shy, but f1 × f2 : ([0, 1] 2 Z , T (c1, c1)) → ({0} × [0, 1]Z, T (c1, c1)) is not shy. Proof. That f1 and f2 are shy is easily seen. Further, f1 × f2 is a surjec- tion. Notice that (0, 0) ↔T (c1,c1) (1, 1), but (f1 × f2)(0, 0) = (0, 0) and (f1 × f2)(1, 1) = (0, 1) are neither equal nor T (c1, c1)-adjacent. Therefore, f1×f2 is not (T (c1, c1), T (c1, c1))-continuous, hence is not (T (c1, c1), T (c1, c1))- shy. � For the Cartesian product adjacency, we have the following. Theorem 9.5. Let fi : (Xi, κi) → (Yi, λi) be a surjection between digital images, 1 ≤ i ≤ v. Let X = Πvi=1Xi, Y = Π v i=1Yi. Then the product function f = Πvi=1fi : (X, × v i=1κi) → (Y, × v i=1λi) is shy if and only if fi is shy for each i. Proof. Suppose f is shy. Then clearly each fi is a surjection, and by Theo- rem 3.17, fi is continuous. Let yi ∈ Yi. Let y = (y1, . . . , yv) ∈ Y . Since f is shy, f −1(y) = Πvi=1f −1 i (yi) is ×vi=1κi-connected. By Theorem 3.18, the projection map pi is continuous, so pi(f −1(y)) = f−1 i (yi) is κi-connected. Let y′ ∈ Y be such that y′ ↔×v i=1 λi y. Then y ′ must be among the points qi = (y1, . . . , yi−1, y ′ i, yi+1, . . . , yv), where y ′ i ∈ Yi satisfies y ′ i ↔λi yi. Since f is shy, f−1({y, qi}) = f −1(y) ∪ f−1(qi) is × v i=1κi-connected. Since pi is continuous, pi(f −1({y, qi})) = pi(f −1(y) ∪ f−1(qi)) = f −1 i (yi) ∪ f −1 i (y ′ i) = f −1 i ({yi, y ′ i}) is κi-connected. This completes the proof that each fi is shy. Suppose each fi is shy. Then clearly f is a surjection, and by Theorem 3.17, f is continuous. Let yi ∈ Yi. Let y = (y1, . . . , yv) ∈ Y . Since fi is shy, f −1 i (yi) is κi- connected. By Theorem 4.4, (9.1) f−1(y) = Πvi=1f −1 i (yi) is × v i=1 κi-connected. Let y′ ∈ Y be such that y′ ↔×v i=1 λi y. Then for some index i, y ′ = (y1, . . . , yi−1, y ′ i, yi+1, . . . , yv), where y ′ i ∈ Yi satisfies y ′ i ↔λi yi. Similarly, (9.2) f−1(y′) is ×vi=1 κi-connected. Since fi is shy, f −1 i ({yi, y ′ i}) is connected, so there exist xi ∈ f −1 i (yi), x ′ i ∈ f−1i (y ′ i) such that xi ↔κi x ′ i or xi = x ′ i. For indices j 6= i, let xj ∈ f −1 j (yj). Then w = (x1, . . . , xv) and w ′ = (x1, . . . , xi−1, x ′ i, xi+1, . . . , xv) satisfy (9.3) w ∈ f−1(y), w′ ∈ f−1(y′), and w ↔×v i=1 κi w ′ or w = w′. From statements (9.1), (9.2), and (9.3), we conclude that f−1({y, y′}) is ×vi=1κi- connected. Therefore, f is shy. � c© AGT, UPV, 2018 Appl. Gen. Topol. 19, no. 1 50 Product adjacencies in digital topology For the lexicographic adjacency, we have the following. Theorem 9.6. Let fi : (Xi, κi) → (Yi, λi) be functions between digital images, 1 ≤ i ≤ v. Let X = Πvi=1Xi, Y = Π v i=1Yi, f = Π v i=1fi : (X, L(κ1, . . . , κv)) → (Y, L(λ1, . . . , λv)). If each fi is shy, then f is shy. Proof. Let y = (y1, . . . , yv) ∈ Y , where yi ∈ Yi. Then f −1(y) = Πvi=1f −1 i (yi). Since each fi is shy, f −1 i (yi) is κi-connected. By Theorem 4.6, f −1(y) is L(κ1, . . . , κv)-connected. Let p = (y′1, . . . , y ′ v) ↔L(λ1,...,λv) y in Y . Then for some smallest index k, y′k ↔λk yk and if k > 1 then yi = y ′ i for i < k. Since fk is shy, f −1 k ({yk, y ′ k}) is κk-connected. Further, if k > 1 then f −1 i ({yi, y ′ i}) = f −1 i (yi) is connected, since fi is shy. Now, (9.4) f−1(p) = Πikf −1 i (yi), (9.5) f−1(p′) = Πikf −1 i (y ′ i) By the shyness of the fi and Theorem 4.6, each of f −1(p) and f−1(p′) is L(κ1, . . . , κv)-connected. Further, since yi = y ′ i for i < k and, by shyness of fk, (9.6) f−1 k ({yk, y ′ k}) is κk-connected, from statements (9.4), (9.5), and (9.6) we can conclude that f−1(p) and f−1(p′) are L(κ1, . . . , κv)-adjacent sets. Therefore, f −1({p, p′}) = f−1(p) ∪ f−1(p′) is L(κ1, . . . , κv)-connected. Therefore, f is shy. � The following shows that the converse of Theorem 9.6 is not generally true. Example 9.7. Let f1 : ([0, 1]Z, c1) → {0} ⊂ (Z, c1) be the function f1(x) = 0. Let f2 : ({0, 2}, c1) → {0} ⊂ (Z, c1) be the function f2(x) = 0. Then f1 × f2 : ([0, 1]Z × {0, 2}, L(c1, c1)) → ({(0, 0)}, L(c1, c1)) is shy, but f2 is not shy. Proof. Since f−12 (0) is not connected, f2 is not shy. However, [0, 1]Z × {0, 2} is L(c1, c1)-connected, as discussed in Example 3.27, so, from Definition 2.30, f1 × f2 is shy. � 10. Further remarks We have studied the tensor product, Cartesian product, and lexicographic adjacencies for finite Cartesian products of digital images. We have obtained many results for “product” and “factor” properties that parallel results ob- tained for extensions of the normal product adjacency in [9]. However, there are many properties known [9] for the normal product ad- jacency whose analogs for the adjacencies studied here are either false or we c© AGT, UPV, 2018 Appl. Gen. Topol. 19, no. 1 51 L. Boxer were not able to derive. By comparing the results of [9] with those of the cur- rent paper, it appears that the normal product adjacency is the adjacency that yields the most satisfying results for Cartesian products of digital images. Acknowledgements. The anonymous reviewers were very helpful. Their corrections and suggestions are gratefully acknowledged. References [1] C. Berge, Graphs and hypergraphs, 2nd edition, North-Holland, Amsterdam, 1976. [2] K. Borsuk, Theory of retracts, Polish Scientific Publishers, Warsaw, 1967. [3] L. Boxer, Digitally continuous functions, Pattern Recognition Letters 15 (1994), 833– 839. [4] L. Boxer, A classical construction for the digital fundamental group, Pattern Recognition Letters 10 (1999), 51–62. [5] L. Boxer, Properties of digital homotopy, Journal of Mathematical Imaging and Vision 22 (2005), 19–26. [6] L. Boxer, Digital products, wedges, and covering spaces, Journal of Mathematical Imag- ing and Vision 25 (2006), 159–171. [7] L. Boxer, Remarks on digitally continuous multivalued functions, Journal of Advances in Mathematics 9, no. 1 (2014), 1755–1762. [8] L. Boxer, Digital shy maps, Applied General Topology 18, no. 1 (2017), 143–152. [9] L. Boxer, Generalized normal product adjacency in digital topology, Applied General Topology 18, no. 2 (2017), 401–427. [10] L. Boxer, O. Ege, I. Karaca, J. Lopez, and J. Louwsma, Digital fixed points, approximate fixed points and universal functions, Applied General Topology 17, no. 2 (2016), 159– 172. [11] L. Boxer and I. Karaca, Fundamental groups for digital products, Advances and Appli- cations in Mathematical Sciences 11, no. 4 (2012), 161–180. [12] L. Boxer and P. C. Staecker, Connectivity preserving multivalued functions in digital topology, Journal of Mathematical Imaging and Vision 55, no. 3 (2016), 370–377. [13] L. Boxer and P. C. Staecker, Remarks on pointed digital homotopy, Topology Proceed- ings 51 (2018), 19–37. [14] L. Chen, Gradually varied surfaces and its optimal uniform approximation, SPIE Pro- ceedings 2182 (1994), 300–307. [15] L. Chen, Discrete surfaces and manifolds, Scientific Practical Computing, Rockville, MD, 2004. [16] C. Escribano, A. Giraldo and M. Sastre, Digitally Continuous Multivalued Functions, in: Discrete Geometry for Computer Imagery, Lecture Notes in Computer Science, v. 4992, Springer, 2008, 81–92. [17] C. Escribano, A. Giraldo and M. Sastre, Digitally continuous multivalued functions, morphological operations and thinning algorithms, Journal of Mathematical Imaging and Vision 42 (2012), 76–91. [18] A. Giraldo and M. Sastre, On the composition of digitally continuous multivalued func- tions, Journal of Mathematical Imaging and Vision 53, no. 2 (2015), 196–209. [19] F. Harary, On the composition of two graphs, Duke Mathematical Journal 26, no. 1 (1959), 29–34. [20] F. Harary and C. A. Trauth, Jr., Connectedness of products of two directed graphs, SIAM Journal on Applied Mathematics 14, no. 2 (1966), 250–254. [21] S.-E. Han, Computer topology and its applications, Honam Math. Journal 25 (2003), 153–162. c© AGT, UPV, 2018 Appl. Gen. Topol. 19, no. 1 52 Product adjacencies in digital topology [22] S.-E. Han, Non-product property of the digital fundamental group, Information Sciences 171 (2005), 73–91. [23] E. Khalimsky, Motion, deformation, and homotopy in finite spaces, in: Proceedings IEEE International Conference on Systems, Man, and Cybernetics, 1987, 227–234. [24] V. A. Kovalevsky, A new concept for digital geometry, Shape in Picture, Springer-Verlag, New York, 1994, pp. 37–51. [25] A. Rosenfeld, ‘Continuous’ functions on digital images, Pattern Recognition Letters 4 (1987), 177–184. [26] G. Sabidussi, Graph multiplication, Mathematische Zeitschrift 72 (1960), 446–457. [27] R. Tsaur and M. Smyth, “Continuous” multifunctions in discrete spaces with applica- tions to fixed point theory, in: Bertrand, G., Imiya, A., Klette, R. (eds.), Digital and Image Geometry, Lecture Notes in Computer Science, vol. 2243, pp. 151–162. Springer Berlin / Heidelberg (2001). [28] J. H. van Lint and R. M. Wilson, A Course in combinatorics, Cambridge University Press, Cambridge, 1992. c© AGT, UPV, 2018 Appl. Gen. Topol. 19, no. 1 53