CasertaWatsonAGT.dvi @ Applied General Topology c© Universidad Politécnica de Valencia Volume 10, No. 2, 2009 pp. 245-267 Michael spaces and Dowker planks Agata Caserta∗ and Stephen Watson Abstract. We investigate the Lindelöf property of Dowker planks. In particular, we give necessary conditions such that the product of a Dowker plank with the irrationals is not Lindelöf. We also show that if there exists a Michael space, then, under some conditions involving singular cardinals, there is one that is a Dowker plank. 2000 AMS Classification: Primary 54C50, 54D20, 54G15. Keywords: Michael space, Michael function, NL property, Haydon plank. 1. Introduction In 1963 E. Michael constructed, under the continuum hypothesis, a Lindelöf space whose product with the irrationals is not normal (see [7]). Such a space is known as a Michael space. An open problem is to construct a Michael space in ZFC without additional axioms. The aim of this paper is to provide necessary conditions for the existence of a Michael space, and to give some examples of Michael spaces. Our work is associated to the results in [8]. In this note, P stands for the set of the irrational numbers, and the Cantor set C is viewed as a compactification of P obtained by adding a countable set QC . Ordinal numbers are denoted by Greek letters; when viewed as topological spaces, they are given the order topology. Products of topological spaces are endowed with the standard product topology. The symbol [A]λ denotes the family of subsets of A having size exactly λ. The symbols [A]≤λ and [A]<λ have similar meaning. Let ≤∗ be the quasi-order on a countable product of ordered sets that is associated to the coordinate-wise order on each set. Thus f ≤∗ g stands for f (n) ≤ g(n) for all but finitely many n ∈ ω. A subset of ωω is unbounded if it is unbounded in (ωω, ≤∗). A dominating family is an unbounded set that is ∗Corresponding author. 246 A. Caserta and S. Watson cofinal in (ω ω, ≤∗). A subset of ωω is a scale if it is a dominating family and is well-ordered by ≤∗. Recall that P can be identified with ωω with the product topology. For each ξ ∈ <ωω = {η | η : [0, n] → ω for some n}, a basic open neighborhood of ξ in the product topology is {f ∈ ω ω : ξ ⊆ f}. For every g ∈ ωω, the sets {f ∈ ωω : f ≤ g} and {f ∈ ωω : f ≤∗ g} are respectively compact and σ-compact (see [2]). Let X and Y be topological spaces. A set A ⊆ X is Y-analytic if it is a projection on X of a closed subset of X × Y . In particular, A ⊆ X is analytic if it is P-analytic. Given a function f : X → Y , the small image of A ⊆ X is defined by f ♯(A) = {y ∈ Y : f −1(y) ⊆ A}. Sometimes we abuse of terminology and say that f ♯ is open, with the meaning that for each open subset A of X, f ♯(A) is an open subset of Y . In most cases we will employ the notation used in [4] and [6]. 2. Michael sequences and Michael functions We start the section with the definition of a Michael sequence. The first goal of this section is to show that Michael sequences may be assumed to be continuous. Definition 2.1. Let {Xξ}ξ≤θ be a decreasing sequence of sets. It is a contin- uous sequence if for any γ ≤ θ, with γ limit ordinal, Xγ = ∩ξ<γ Xξ. Definition 2.2 (Moore [8]). A decreasing sequence {Xξ}ξ≤θ of subsets of a topological space Z is said to be a K-Michael sequence if the following conditions hold: (i) for each K compact subset of Z \ Xθ the ordinal δK = min{ξ ≤ θ : Xξ ∩ K = ∅} does not have uncountable cofinality. In particular an F-Michael sequence is a K-Michael sequence satisfying the following additional condition: (ii)F for each F closed subset of Z \ Xθ the ordinal δF = min{ξ ≤ θ : Xξ ∩ F = ∅} is either θ or does not have uncountable cofinality. Also given a topological space Y , an A(Y )-Michael sequence is a K-Michael sequence satisfying the following additional condition: (ii)A for each A which is Y -analytic in Z \ Xθ the ordinal δA = min{ξ ≤ θ : Xξ ∩ A = ∅} is either θ or does not have uncountable cofinality. Remark 2.3. In the definition of a K-Michael sequence, we observe that the property of being a continuous sequence is partially satisfied . In other words, for every limit ordinal γ < θ with cfγ > ω it follows that Xγ = ∩ξ<γ Xξ. Indeed, let x ∈ ∩ξ<γ Xξ \ Xγ . Then {x} is a compact subset of Z \ Xθ, and δ{x} = γ, so that cfδ{x} > ω in contradiction with the definition of K-Michael sequence. Michael spaces and Dowker planks 247 Lemma 2.4. Let θ be a cardinal and {Xξ}ξ≤θ (strictly) decreasing sequence such that Xγ = ∩ξ<γ Xξ for every limit ordinal γ < θ with cfγ > ω. Then there exists {Yξ}ξ≤θ continuous (strictly) decreasing sequence, such that Yα = Xα for every α < θ with cfα 6= ω. Proof. Let {Xξ}ξ≤θ be decreasing sequence. Define {Yξ}ξ≤θ such that Yα = Xα for every α < θ with cfα > ω, otherwise Yα = ∩ξ≤αXξ. Clearly Yη ⊇ Yξ for every η < ξ ≤ θ. Moreover for every α < θ with cfα = ω, Yα ⊇ Xα. By construction, we have that {Yξ}ξ≤θ is a continuous sequence. Assume that all the subsets Xξ ∈ {Xξ}ξ≤θ are distinct. Then Yα ⊇ Xα ⊃ Xα+1 = Yα+1 implies that Yα’s are distinct. � In case we have two or more sequences of subsets of Z of length θ + 1, having the same last element, and given H, we denote δH with respect the sequence {Xξ}ξ≤θ with δ X̃ H . Lemma 2.5. Let θ be a cardinal with cfθ > ω, {Xξ}ξ≤θ and {Yξ}ξ≤θ two decreasing sequences of subsets of a topological space Z, such that Yα = Xα for every α < θ with cfα 6= ω. Then δX̃H < δ Ỹ H ⇒ ( δ Ỹ H = δ X̃ H + 1 ) ∧ ( cfδ X̃ H = ω ) with H ⊆ Z. Proof. From Remark 2.3 it follows that for every α < θ with cfα > ω, Xα = ∩ξ<αXξ, and Xα = Yα ⊇ ∩ξ<αYξ. We have also that for every α < θ with cfα > ω there exists a cofinal sequence (αη)η ω and (ii) cfδ Ỹ H = ω. If (i) holds, then YδỸ H ∩ H = X δỸ H ∩ H = ∅, therefore δX̃H ≤ δ Ỹ H . If δ X̃ H < δ Ỹ H there exists αη, such that δ X̃ H < αη < δ Ỹ H . Then by minimality of δỸH we have Yαη ∩ K 6= ∅ and Yαη ∩ K ⊆ Xαη ∩ K. Moreover Xαη ⊆ XδX̃ H , so X δX̃ H ∩ K 6= ∅ which is in contradiction with the definition of δK . Thus δ X̃ H = δ Ỹ H . For (ii), assume by contradiction, that δ X̃ H 6= δ Ỹ H . Since cfδX̃H = cfδ Ỹ H = ω, there exists α successor ordinal such that δ Ỹ H < α < δ X̃ H . Then Xα = Yα, and so YδỸ H ⊇ Yα = Xα ⊇ XδX̃ H . Therefore Xα ∩ H = ∅ which is in contradiction with the minimality of δX̃H . Thus δ X̃ H = δ Ỹ H . � Corollary 2.6. Let θ be a cardinal with cfθ > ω, {Xξ}ξ≤θ and {Yξ}ξ≤θ two decreasing sequences of subsets of Z, such that Yα = Xα for every α < θ with cfα 6= ω. Let H ⊂ Z, then δX̃H = δ Ỹ H if either one has uncountable cofinality. Corollary 2.7. Let θ be a cardinal with cfθ > ω, {Xξ}ξ≤θ and {Yξ}ξ≤θ two decreasing sequences of subsets of a topological space Z, such that Yα = Xα for every α < θ with cfα 6= ω. Then {Xξ}ξ≤θ is a K-Michael (resp., F-Michael or A(Y )-Michael) sequence if and only if {Yξ}ξ≤θ is a K-Michael (resp., F-Michael or A(Y )-Michael) sequence. 248 A. Caserta and S. Watson Proof. Let {Xξ}ξ≤θ be a K-Michael (resp., F-Michael or A(Y )-Michael) se- quence. By hypothesis Yθ = Xθ. Let H ⊆ (Z \ Xθ) compact (resp., closed or analytic). Then cfδX̃H ≤ ω (resp., either δ X̃ H ≤ ω or δ X̃ H = θ). We want to check that cfδỸK ≤ ω (resp., either δ Ỹ H ≤ ω or δ Ỹ H = θ). Assume not, i.e., cfδỸK > ω, (resp., ω < cfδ Ỹ K < θ) Corollary 2.6 implies that δ X̃ K = δ Ỹ K , which is a contradiction. � Corollary 2.8. Let θ be a cardinal with cfθ > ω. The following are equivalent: (i) there exists {Xξ}ξ≤θ which is K-Michael (resp., F-Michael or A(Y )- Michael strictly decreasing) sequence; (ii) there exists {Xξ}ξ≤θ continuous K-Michael (resp., F-Michael or A(Y )- Michael strictly decreasing) sequence. Next we introduce the definition of Michael function and we analyze the relationship between Michael functions and Michael sequences. Definition 2.9. Let Z be a topological space and f : Z → θ + 1 an arbitrary function. Then f is said to be a K-Michael function if the following condition holds: (i) for each K compact subset of Z \ f −1({θ}), supx∈Kf(x) + 1 does not have uncountable cofinality. In particular an F-Michael function is a K-Michael function satisfying the following additional condition: (ii) for every F closed subset of Z \ f −1({θ}), supx∈Ff(x) + 1 is either θ or does not have uncountable cofinality. Also given a topological space Y , an A(Y )-Michael function is a K-Michael function satisfying the following additional condition: (ii) for every A which is Y -analytic in Z \f −1({θ}), supx∈Af(x)+ 1 is either θ or does not have uncountable cofinality. In the next proposition we will show the equivalence of continuous K-Michael sequences with K-Michael functions f : Z → θ + 1. Lemma 2.10. Let Z be a topological space, f : Z → θ + 1 be an arbitrary function with θ cardinal. If Xξ = {x ∈ Z : f (x) ≥ ξ} for every ξ ∈ θ, then δH = supx∈Hf(x) + 1 for every H ⊆ Z \ Xθ. Proof. By definition we have that δH = min{ξ ≤ θ : K ∩ Xξ = ∅} = min{ξ ≤ θ : ∀x ∈ K (x /∈ Xξ)} = min{ξ ≤ θ : ∀x ∈ K (f (x) < ξ)} = sup{f(x) + 1 : x ∈ K}. � Lemma 2.11. Let θ be a cardinal, {Xξ}ξ≤θ a continuous sequence of subsets of topological space Z, and f : Z → θ + 1 a function defined by f (x) = sup{γ ∈ θ + 1 : x ∈ Xγ}. Then we have: (i) Xξ = {x ∈ Z : f (x) ≥ ξ} for every ξ ∈ θ; (ii) f is surjective if and only if {Xξ}ξ≤θ is strictly decreasing. Michael spaces and Dowker planks 249 Proof. To show (i), we have that for every ξ ∈ θ, {x ∈ Z : f (x) ≥ ξ} = {x ∈ Z : sup{γ ∈ θ : x ∈ Xγ} ≥ ξ}. From the continuity follow {x ∈ Z : sup{γ ∈ θ : x ∈ Xγ} ≥ ξ} = {x ∈ Z : x ∈ Xξ} = Xξ. For (ii), first assume that f is surjective. By (i) we have that Xξ = {x ∈ Z : f (x) ≥ ξ} for every ξ ∈ θ, and so {Xξ}ξ≤θ is a decreasing sequence. Assume that there exist α, β ∈ θ with α < β such that Xα = Xβ. Thus there exist ξ ∈ θ with α < ξ ≤ β and z ∈ Z such that f (x) = ξ. Hence x ∈ Xβ but x /∈ Xα, a contradiction. On the other hand, assume that {Xξ}ξ≤θ is strictly decreasing, and f is not surjective. Then there exists α < θ such that f (x) 6= α for any x ∈ Z \ Xθ, with Xθ = f −1({θ}). Let f (x) > α. From (i) it follows that there exists α < θ such that (Z \ Xθ) ⊆ Xα. Thus Xβ = Xα for any β ≤ α, which contradicts the fact that the sequence is strictly decreasing. If f (x) < α, follow that (Z \ Xθ) ∩ Xα = ∅, which is a contradiction. � Proposition 2.12. Let Z and Y be two topological spaces, θ a cardinal with cfθ > ω. For every Q ⊆ Z, the following statements are equivalent: (i) there exists a continuous K-Michael (resp., F-Michael or A- Michael) sequence {Xξ}ξ≤θ with Q = Xθ; (ii) there exists K-Michael (resp., F-Michael or A(Y )-Michael) function f : Z → θ + 1, with Q = f −1({θ}); (iii) there exists K-Michael (resp., F-Michael or A(Y )-Michael) sequence {Xξ}ξ≤θ with Q = Xθ. Proof. (i) ⇒ (ii). Let {Xξ}ξ≤θ be a continuous K-Michael sequence. Define the following map f : Z → θ + 1 such that f (x) = sup{γ ∈ θ + 1 : x ∈ Xγ}. Clearly f −1({θ}) = Xθ. Now let H ⊂ (Z \ Q) be a compact (resp., closed or analytic) subset. By Lemma 2.11, for any α ≤ θ, Xα = {x ∈ Z : f (x) ≥ α}. By Lemma 2.10, supx∈Hf(x) + 1) = δH. Since cfδK ≤ ω (resp., either cfδH ≤ ω or cfδH = θ), then cf (supx∈Kf(x) + 1)) ≤ ω (resp., either cf (supx∈Hf(x) + 1) ≤ ω or cf (supx∈Hf(x) + 1) = θ ). (ii) ⇒ (iii). Let f : Z → θ + 1 be a K-Michael function with Q = f −1({θ}). For any α ≤ θ define Xα = {x ∈ Z : f (x) ≥ α}. Clearly Xθ = f −1({θ}) and Xξ ⊇ Xη for any ξ < η ≤ θ. Let now H ⊂ (Z \Q) be a compact (resp., closed or analytic) subset, we want to show that cfδH ≤ ω. By Lemma 2.10, supx∈Kf(x)+ 1 = δH. Since cf (supx∈Hf(x) + 1) ≤ ω (resp., either cf (supx∈Hf(x) + 1) ≤ ω or cf (supx∈Hf(x) + 1) = θ ), then cfδH ≤ ω (resp., either cfδH ≤ ω or cfδH = θ). (iii) ⇒ (i). Follow from Corollary 2.8. � Corollary 2.13. Let θ be a cardinal of uncountable cofinality, Z and Y two topological spaces. There exists a continuous K-Michael (resp., F-Michael or A(Y )-Michael) strictly decreasing sequence {Xξ}ξ≤θ of subsets of Z, if and only if there exists K-Michael (resp., F-Michael or A(Y )-Michael) function f : Z → θ + 1 that is surjective. 250 A. Caserta and S. Watson 3. Local properties of Michael functions In this section we want to analyze and characterize the properties of being a Michael function. First we need the following definition. Definition 3.1. Let Z be a topological space, h : Z → θ + 1 an arbitrary function with θ cardinal. For every α ≤ θ we say that h is Michael at α if cfα > ω ⇒ (∀F ⊆ Z closed (∀z ∈ F h(z) < α) ⇒ (supz∈Fh(z) < α)). Moreover h is σ-Michael at α if cfα > ω ⇒ (∀C ⊆ Z Fσ-set (∀z ∈ C h(z) < α) ⇒ (supz∈Ch(z) < α)). Directly from the definition follow: Lemma 3.2. Let Z be a topological space, h : Z → θ + 1 an arbitrary function with θ cardinal. The following statements are equivalent: (i) h is Michael at α, (ii) cfα > ω ⇒ (∀U ⊆ Z open (h−1[α, θ] ⊆ U) ⇒ (supz∈Z\Uh(z) < α)). Lemma 3.3. Let Z be a topological space, h : Z → θ + 1 an arbitrary function with θ cardinal. The following statements are equivalent: (i) h is σ-Michael at α, (ii) cfα > ω ⇒ (∀G ⊆ Z Gδ-set (h −1([α, θ]) ⊆ G) ⇒ (supz∈Z\Gh(z) < α)). Lemma 3.4. Let Z be a topological space, h : Z → θ + 1 an arbitrary function with θ cardinal. Then h is σ-Michael at α if and only if h is Michael at α. Proof. Let cfα > ω, and C = ⋃ n∈ω Fn with Fn closed subset of Z such that for every z ∈ C h(z) < α. Then for every n ∈ ω and for every z ∈ Fn, h(z) < α. Let αn = supz∈Fnh(z). Since h is Michael at α, follow αn < α for every n ∈ ω. Then supz∈Ch(z) = supn∈ωαn. From cfα > ω if follows that supn∈ωαn < α. � An arbitrary function h : Z → θ+1, induces a new function ĥ : Z×Y → θ+1, defined by ĥ(x) = h(π1(x)) for every x ∈ Z × Y , where Y is an arbitrary topological space, and π1 the projection of Z × Y onto its first coordinate space. Clearly this raises the question whether ĥ is Michael at some ordinal α ≤ θ. Lemma 3.5. Let Z, Y be two topological spaces, h : Z → θ + 1 is an arbitrary function, ĥ : Z × Y → θ + 1 with θ cardinal. Then the following statements are equivalent: (i) ĥ is Michael at α, (ii) cfα > ω ⇒ (∀A ⊆ Z Y-analytic (∀z∈A h(z) < α) ⇒ supz∈Ah(z) < α). Moreover, if ĥ is Michael at α for some α ≤ θ, then h is Michael at the same ordinal. But the converse does not hold. Now, given a function h : Z → θ + 1, we want to characterize the property of being Michael at some ordinal for h, in term of a Michael function. Michael spaces and Dowker planks 251 Proposition 3.6. Let Z be a topological space, h : Z → θ + 1 a function with θ cardinal. Then the following statements are equivalent: (i) h is a F-Michael function; (ii) h is Michael at α for every α ≤ θ. Proof. Assume that h is not a F-Michael function. Then there is a closed set F ⊆ (Z \ h−1({θ})) such that cf(supz∈Fh(z) + 1) > ω. Let α = supz∈Kh(z) + 1. Note that h(z) < α for every z ∈ F . Since h is Michael at α, from Lemma 3.2 follow that supz∈Fh(z) < α, which is a contradiction. Vice versa, Assume that h is a F-Michael function. Let α ∈ θ such that cfα > ω. We want to show that h is Michael at α. Let U be an open set of Z such that h−1([α, θ]) ⊂ U . Then Z \ U is such that h(z) < α for every z ∈ Z. Therefore cf(supZ\Uh(z) + 1) ≤ ω. Since cfα > ω there exists β < α such that supz∈Z\Uh(z) + 1 ≤ β < α. � From the previous proof we can argue that if h is Michael at α for every α ∈ θ, then h is F-Michael function, and so K-Michael function, but the vice versa does not hold. Clearly it is true in case Z is a compact space. Moreover we have shown that if h is a F-Michael function, then there exists an ordinal α such that h is Michael at α. The vice versa does not hold, we needed the property of being Michael to be satisfied at each ordinal into the codomain of h. Proposition 3.7. Let Z, Y be two topological spaces, h : Z → θ + 1 and ĥ : Z × Y → θ + 1 functions with θ cardinal. Then the following statements are equivalent: (i) h is a A(Y )-Michael function, (ii) ĥ is Michael at α for each α ≤ θ. Proof. Assume that h is not a A(Y )-Michael function. Then there is a set A ⊆ (Z \h−1({θ})) which is the projection onto Z of a closed subset F of Z ×Y , such that ω < cf(supz∈Ah(z) + 1) < θ. Let α = supz∈Ah(z) + 1. Since A is a Y - analytic subset of Z and ĥ is Michael at α it follows that supz∈Ah(z) < α which is in contradiction with supz∈Ah(z) = {β ≤ α : h(z) ≥ β for same z ∈ A} = α. Assume that h is a A(Y )-Michael function. Let α < θ with cfα > ω. Let A be a Y -analytic subset of Z, i.e., A = π(F ) where F is a closed subset of Z × Y such that h(z) < α for every z ∈ A. Then cf(supz∈Ah(z) + 1) ≤ ω. Since cfα > ω, it follows that supz∈Ah(z) < α. � Remark 3.8. Note that by Proposition 2.12 and Proposition 3.6, it follows that if h : C → θ + 1 is such that QC = h −1({θ}), the property of being Michael at α for every α ≤ θ is equivalent to the notion of K-Michael sequence {Xξ}ξ≤θ [M [8]], where for every ξ ∈ θ, Xξ ⊆ C and Xθ = QC . The next Proposition give us conditions on the function h : Z → θ + 1, so that the function ĥ is not Michael at θ. 252 A. Caserta and S. Watson Proposition 3.9. Let θ be a cardinal with cfθ > ω, Z a topological space. Let h : Z → θ + 1 be a function such that h(Z) ∩ (α, θ) 6= ∅ for every α < θ. Then ĥ is not Michael at θ, where ĥ : Z × (Z \ h−1({θ})) → θ + 1. Proof. Set ∆ = {(z, z) : z ∈ Z \ h−1({θ})}. Then ∆ is a closed subset of Z × (Z \ h−1({θ})) such that for every z ∈ Z \ h−1({θ}) we have h(z) < θ. But supz∈Z\h−1({θ})h(z) = θ. If not, there exists α < θ such that supz∈Z\h−1({θ})h(z) = α. Since h(Z) ∩ (α, θ) 6= ∅, there exists β with α < β < θ, z ∈ Z \ h−1({θ}) such that h(z) = β which is a contradiction. � 4. NL Property In this section we introduce the new definition of NL Property at some ordinal, and we give examples of functions which have this property. Definition 4.1. Let X be a topological space, θ a cardinal and  : X → θ an arbitrary function. For each α ≤ θ with cfα > ω, we say that  has the property NL at α if for every A ⊆ X such that (A) is cofinal in α, A is not Lindelöf Remark 4.2. A banal case for the function  : X → θ + 1 to have the property NL at each α ≤ θ with cfα > ω, is for  = idθ+1. Indeed every subset of α which is cofinal in α cannot be Lindelöf Another simple case in which  has the property NL at each α ≤ θ with cfα > ω is when −1(β) is open in X for every β < α. Indeed, assume that A ⊆ X such that (A) is cofinal in α, and by contradiction A is Lindelöf. Then {−1(β)}β∈α is an open cover for A, therefore there exist β0 ∈ α countable such that A ⊆ ⋃ β∈β0 −1(β). Thus A ⊆ −1(β0) which is a contradiction. Other examples of function with the property NL are given. Before we need the following definitions. Definition 4.3. Let θ be a cardinal and X a topological space. The family {Aα}α∈θ is a special Gδ family of X, if for every α ∈ θ, Aα = ⋂ n∈ω A n α where each Anα is open in X and for every n ∈ ω, {A n α}α∈θ is an increasing family. Definition 4.4. Let θ be a cardinal and X a topological space. The function  : X → θ + 1 is a special at α with α ≤ θ, if there exists a sequence of continuous functions (n)n∈ω with n : X → θ + 1 such that for every n ∈ ω, (i) −1(α) ⊆ −1n (α), (ii) (x) ≤ n(x) for every x ∈ X, (iii) {−1n (α)}α≤θ is an increasing family. Lemma 4.5. Let θ be a cardinal, X a topological space and  : X → θ + 1 a function. The following statements are equivalent: (i)  is special at each α ≤ θ (ii) {−1(α)}α≤θ is a special Gδ family of X. Proof. Let α ≤ θ and  be special at α. Let (n,α)n∈ω be a sequence of con- tinuous functions n,α : X → θ + 1 satisfying properties in Definition 4.4. By Michael spaces and Dowker planks 253 continuity of each n,α, the set  −1 n,α(α) is open in X for each n ∈ ω. Since −1(α) ⊆ −1n,α(α), it follows that  −1(α) ⊆ ⋂ n∈ω  −1 n,α(α) for each α ∈ ω. We show that ⋂ n∈ω  −1 n,α(α) ⊆  −1(α). Let x ∈ ⋂ n∈ω  −1 n,α(α), hence x ∈  −1 n,α(α) for each n ∈ ω, i.e., for each n, n,α(x) ∈ α. Since (x) ≤ n,α(x) for all n ∈ ω and x ∈ X, we have that (x) ≤ α. Thus x ∈ −1(α) and −1(α) = ⋂ n∈ω  −1 n,α) (α) for each α ≤ θ. Moreover for every n ∈ ω, we have that {−1n,α(α)}α≤θ is an increasing family. Vice versa, assume that {−1(α)}α≤θ is a special Gδ family of X. Let α ≤ θ. By hypothesis, −1(α) = ⋂ n∈ω A n α with the property that A n α is an open set and for every n the family {Anα}α≤θ is increasing. Define for each n ∈ ω, the function n : X → θ + 1 by n(x) = min{ξ ∈ θ + 1 : x ∈ A n ξ }. We have that for each α ≤ θ, −1n (α) = A n α. Indeed, A n α ⊆  −1 n (α) and for each γ > α there is not y ∈ Anγ \ A n α such that y ∈  −1 n (α). Otherwise from y ∈  −1 n (α), it follows that y ∈ Anα which is a contradiction. Thus n is continuous for each n and the family {−1n (α)}α≤θ is an increasing. Since  −1(α) = ⋂ n∈ω  −1 n (α), we have that for each n ∈ ω, −1(α) ⊆ −1n (α). Let x ∈ X. It remains to prove that (x) ≤ n(x) for every n ∈ ω. Let (x) = α. Hence x ∈  −1(α) and x ∈ −1n (α) for every n ∈ ω, i.e., the point x is such that min{ξ ∈ θ + 1 : x ∈ Anξ } = α for each n. Therefore n(x) ≥ α for each n ∈ ω. � Proposition 4.6. Let X be a topological space, θ a cardinal and  : X → θ + 1 a function. If {−1(α)}α∈θ is a special Gδ family, then  has the property NL for every α ≤ θ. Proof. Let A ⊆ X, α ≤ θ with cfα > ω and (A) is cofinal in α. For every β ∈ θ, we have −1(β) = ⋂ n∈ω G n β such that for every n ∈ ω, {G n β}β∈θ is an increasing family of open sets. Since (A) is cofinal in α, for all β ∈ α A \ ⋂ n∈ω G n β 6= ∅, i.e., for all β ∈ α there exists n ∈ ω such that A \ Gnβ 6= ∅. There exist n ∈ ω and (βξ)ξ∈cfα increasing sequence with βξ < α, such that A \ G n βξ 6= ∅. Now, fixed n ∈ ω, we have that A ⊆ ⋃ ξ∈cfα G n βξ . Therefore the family {Gnβξ }ξ∈cfα is an open cover of A. If A was Lindelöf, there should be β0 countable such that Gnβ0 would cover A, which is a contradiction. � Proposition 4.7. Let X = ∏ n∈ω θ + 1,  : X → θ + 1 defined by (f ) = min{ξ ∈ θ + 1 : f ≤ fξ}, where {fα}α∈θ ⊆ ∏ n∈ω θ + 1 such that for every α < α ′ fα ≤ fα′ . Then  has the property NL at every α ≤ θ. Proof. By definition −1(α) = {f ∈ X : ∀n ∈ ω f (n) < fα(n)} = ⋂ n∈ω{f ∈ X : f (n) < fα(n)}. Set G n α = {f ∈ X : f (n) < fα(n)}, then for every α ∈ θ+1, Gnα is an open set in X, and moreover for every n ∈ ω, {G n α}α∈θ is an increasing family. Hence {−1(α)}α∈θis a special Gδ family of X. Proposition 4.6 ends the proof. � Corollary 4.8. Let X = ∏ n∈ω θ + 1,  : X → θ + 1 defined by (f ) = min{ξ ∈ θ + 1 : f ≤ fξ}, where fξ is a constant function with value ξ for every ξ ≤ θ. Then  has the property NL at every α ≤ θ. 254 A. Caserta and S. Watson Remark 4.9. Given X = ∏ n∈ω θ + 1, a family {fα}α∈θ ⊆ ∏ n∈ω θ + 1, a sequence of function n(f ) = min{ξ ∈ θ + 1 : f (n) ≤ fξ(n)} and a function (f ) = min{ξ ∈ θ + 1 : f ≤ fξ}, all of them defined in X with value in θ + 1. Then  > supn, and the equality does not hold. Indeed let f : ω → θ + 1 defined by f (n) = 0 for every n 6= 0 and f (0) = 2, and {fξ}ξ∈θ defined by fξ = ~ξ for every ξ ∈ θ with ξ 6= 2 and f2(n) = 0 for every n ∈ ω \ {0, 2} and f (0) = 2, f (2) = 0. Then (f ) = 3 and supnn(f) = 2. There are examples of chain for countable product of ordered spaces, not considering the constant value function, which is a banal example. For example X = ∏ n∈ω\{0} ℵω·n. In (X, ≤) there exists a chain C such that ot(C) = ℵω·ω but not ot(C) = ℵω·ω+1. Given {αn}n∈ω ordinals, what is the set of β such that there exists a function f : β →֒ παn? Remark 4.10. Let κ be a cardinal with cfκ > ω, X = ∏ n∈ω κ + 1 and {fξ}ξ∈κ ⊆ X such that fα ≤ fβ for every α < β < κ. Let  : X → κ + 1, defined by (f ) = min{ξ ∈ κ : f ≤ fξ}. We have that the function  has the property NL at κ. Let A ⊂ X such that (A) is cofinal in κ. Then for every α ∈ κ A * −1(α), i.e., for every α ∈ κ and for every n ∈ ω A * {g ∈ X : g(n) ≤ fα(n)}. Let V n, α = {g ∈ X : g(n) ≤ fα(n)}. Then {Vn,α}n,α is an uncountable open cover of A. If A was Lindelöf, there should exist α0 ∈ κ countable such that {Vn,α} n∈ω α∈α0 is a cover for A which is a contradiction with cfκ > ω. We give an example of function which has the property NL only at some ordinal. Proposition 4.11. Let X = ∏ n∈ω θn +1 with every θn cardinal with cfθn > ω, and  : X → κ + 1 defined by (f ) = min{ξ ∈ κ : f ≤∗ fξ} where κ is a cardinal with cfκ > ω and κ > θn for every n ∈ ω, {fξ}ξ∈κ a dominating family in ( ∏ n∈ω θn, ≤∗). Then  has the property NL at κ. Proof. Let A ⊂ X such that (A) is cofinal in κ. Then for every α ∈ κ A * −1(α), i.e., for every α ∈ κ A * {g ∈ X : g ≤∗ fα}. Then there exists n ∈ ω such that {g(n) : g ∈ A} is unbounded in θn. If not, for every n ∈ ω {g(n) : g ∈ A} is bounded in θn, and since the family {fξ}ξ∈κ is an ≤∗- dominating in ( ∏ n∈ω θn, ≤∗), there should exists ξ ∈ κ such that for every g ∈ A g ≤∗ fξ, which is a contradiction. Thus there exist n ∈ ω such that for every α ∈ θn A * {g ∈ A : g(n) < α}. Let Vn,α = {g ∈ A : g(n) < α}. Then {Vn,α}n,α is an uncountable open cover of A. If A was Lindelöf, there should exist α0 ∈ θn countable such that {Vn,α} n∈ω α∈α0 is a cover for A which is a contradiction with cfθn > ω. � 5. Closed mapping properties In this section we investigate different properties of the projection map, introducing two new definitions. Michael spaces and Dowker planks 255 Let us recall that if f : X → Y is a function and A ⊆ X, then the restriction of f to A, f↾A, is closed if the image of a closed subset of A is a closed subset of Y . Definition 5.1. Given two arbitrary topological spaces X and Y , we say that the function f : X → Y is σ-closed if the image of a closed subset of X is an Fσ subset of Y . Definition 5.2. Let X, Y be two topological spaces. f : X → Y is strongly σ-closed if there exists (Kn)n∈ω with Kn’s closed subsets of X such that X =⋃ n∈ω Kn and f↾Kn is closed for every n ∈ ω. Remark 5.3. We are dealing with three different properties of the function f : X → Y . The following implications hold f closed ⇒ f strongly σ-closed ⇒ f σ-closed Example 5.4. First note that for every countable topological space X which is T1, the map f : X → Y is strongly σ-closed for every topological space Y which is T1. Therefore the map f : Q → R with f = idQ is strongly σ-closed, but it is not a closed map. Example 5.5. [AC] Under the Axiom of choice, the set ω1 can be partitioned in ω stationary sets Sn such that ω1 = ⋃ n∈ω Sn. In other words,there exists a function f : ω1 → ω + 1 defined by f −1(n) = Sn for every n ∈ ω. By definition of stationary set, it follows that for every n ∈ ω and for every club C in ω1 we have C ∩ f −1(n) 6= ∅. Clearly f is σ-closed. We claim that f is not strongly σ-closed, which is equivalent to show that for every (Kn)n∈ω with Kn’s closed subsets of X such that X = ⋃ n∈ω Kn there exists n0 ∈ ω such that the map f ↾ Kn0 is not closed. Indeed let (Kn)n∈ω be any countable family of closed subsets of ω1 such that ω1 = ⋃ n∈ω Kn. Then there exist n0 ∈ ω such that |Kn0| > ℵ0. Then Kn0 is a club in ω1, therefore Kn0 ∩ f −1(n) 6= ∅ for every n ∈ ω. Thus f (Kn0 ) = ω, and so f↾Kn0 is not closed, because the set ω is not closed in its compactification ω + 1. Lemma 5.6. Let X, Z be topological spaces, such that X = ⋃ n∈ω Kn. Let F be a subset of X × Z and Fn = F ∩ (Kn × Z). Let π : X × Z → Z be the projection map. Then π(F ) = ⋃ n∈ω π↾(Kn × Z)(Fn) Proof. Note that X × Z = ⋃ n∈ω(Kn × Z), and for every n ∈ ω, Fn is a subset of Kn × Z such that F = ⋃ n∈ω Fn. Let pn = π↾(Kn × Z). For every n ∈ ω, pn(Fn) ⊆ π(F ). Indeed if z ∈ pn(Fn), there exists (x, z) ∈ Fn such that pn(x, z) = z, therefore there exists (x, z) ∈ F such that π(x, z) = z. Thus z ∈ π(F ). On the other side, if z ∈ π(F ), there exists (x, z) ∈ ⋃ n∈ω Fn such that π(x, z) = z. Therefore there exists n ∈ ω such that (x, z) ∈ Fn such that pn(x, z) = z. � The Kuratowski Theorem is useful: Theorem 5.7. Given a compact Hausdorff space X, the projection map π : X × Z → Z is a closed map, for every topological space Z. 256 A. Caserta and S. Watson An application is given by: Proposition 5.8. Given an Hausdorff space X and the projection map π : X × Z → Z, the following implications hold X σ-compact ⇒ π strongly σ-closed ⇒ π σ-closed Proof. First we show that π is a strongly σ-closed map. From X σ−compact,let X = ⋃ n∈ω Kn where Kn’s are compact in X. Therefore Kn × Z is closed in X × Z. The Kuratowski Theorem assures that the projection map π↾Kn × Z is a closed map, for every topological space Z. For the second implication, let X × Z = ⋃ n∈ω Kn where Kn’s are closed. Let F be a closed subset of X × Z, and Fn = F ∩ Kn. Then for every n ∈ ω Fn is a closed subset of Kn such that F = ⋃ n∈ω Fn. From Lemma 5.6, π(F ) = ⋃ n∈ω π↾Kn(Fn); moreover for every n ∈ ω π↾Kn(Fn) is closed. It follows that π(F ) is Fσ in Z. � The use of the small image of the projection map will recur often. So let us state an useful basic property: Lemma 5.9. Let X and Y be two topological spaces and π : X × Y → Y a projection map. Then for every A ⊆ Y and B, K ⊆ X × Y , (i) A ⊆ (π↾K)♯(B ∩ K) ⇔ (X × A) ∩ K ⊆ B; (ii) A ⊆ π♯(B) ⇔ X × A ⊆ B. Proof. A ⊆ (π ↾K)♯(B ∩ K) = {y ∈ Y : π−1(y) ∩ K ⊆ B ∩ K} ⇔ ∀y ∈ A π−1(y) ∩ K ⊆ B ∩ K ⇔ ∀y ∈ A {(x, y) ∈ K : x ∈ X ∧ π(x, y) = y} ⊆ B ⇔ {(x, y) ∈ K ∩ (X × A) : π(x, y) = y} ⊆ B ⇔ (X × A) ∩ K ⊆ B. � Lemma 5.10. Let X, Y be two topological spaces, f : X → Y an arbitrary function. If f is a closed map, then for every U open in X, f ♯(U ) is an open subset of Y . Moreover if f is σ-closed map, then f ♯(U ) is a Gδ subset of Y . Proposition 5.11. Let the projection π : K × Z → Z be σ-closed, and X ⊆ Z. Let U be an open subset in K × Z which cover K × X, then there exists H ⊇ X which is a Gδ in Z such that U cover K × H. Proof. Set H = π♯(U ). Then, since π is σ-closed, H is a Gδ in Z. By Lemma 5.9 follow that K × H ⊆ U , and X ⊆ H. � Proposition 5.12. Let X be a subset of a topological space Z, and K × Z ⊆ ⋃ n∈ω Kn with every Kn Lindelöf, and for every n ∈ ω π↾Kn is closed, where π : K ×Z → Z is the projection map. If X is Lindelöf, then K ×X is Lindelöf. Proof. Let U be a cover of K × X made by open sets of K × Z. Without loss of generality we can assume that U is closed under countable unions. Fix n ∈ ω, for each z ∈ X, Kn ∩ (K × {z}) is Lindelöf. For every z ∈ X, since U is closed under countable unions there exists Uz ∈ U such that Kn ∩ (K × {z}) ⊂ Uz. Set Az,n = (π↾Kn) ♯(Uz ∩ Kn). Then Az,n is an open subset of Z containing z. From Lemma 5.9 follow that (K × Az,n) ∩ Kn ⊆ Uz. For a fixed n ∈ ω, {Az,n}z∈X is a family of open sets in Z which covers X. Since X is Lindelöf Michael spaces and Dowker planks 257 there exists countably many zni ’s such that {Azni ,n}i∈ω cover X. Moreover we have that for every n ∈ ω (K × Azn i ,n) ∩ Kn ⊆ Uzn i . We claim that {Uzn i }i,n∈ω covers K × X. Indeed, let (k, z) ∈ K × X, then there exists n ∈ ω such that (k, z) ∈ (K × X) ∩ Kn. Fixed such n, there exists i ∈ ω such that z ∈ Azn i ,n. Thus (k, z) ∈ (K × Azn i ,n) ∩ Kn ⊆ Uzn i . Therefore we have that {Uzn i }i,n∈ω is a countable family of open sets of U which covers K × X. � Corollary 5.13. Let X be a subset of a topological space Z, and K = ⋃ n∈ω Kn with every Kn Lindelöf, and for every n ∈ ω π↾Kn × Z is closed, where π : K × Z → Z is the projection map. If X is Lindelöf, then K × X is Lindelöf. Remark 5.14. From the proof of Lemma 5.12 we can also get that if K, X and π satisfy the assumptions, there exists U ∈ U such that it covers K × X, where U is an open cover of K × X made by open set in K × Z closed under countable union. Corollary 5.15. Let K, Z two topological spaces, π : K×Z → Z the projection map, and X ⊂ Z. If (i) X is Lindelöf, (ii) K is Lindelöf, (iii) π is strongly σ-closed, then K × X is Lindelöf. Corollary 5.16. Let K, Z two topological spaces, π : K×Z → Z the projection map and X ⊂ Z. Let U be a family of open sets of K × Z which covers K × X. If (i) X is Lindelöf, (ii) K is Lindelöf, (iii) π is strongly σ-closed, then there exists H ⊃ X which is a Gδ in Z and a countable subfamily of U which covers K × H. Proof. Without loss of generality we can assume that U is closed under count- able unions. By Corollary 5.15, K ×X is Lindelöf, therefore there exists U0 ⊆ U countable such that cover K × X. Set U0 = ⋃ U0. Then U0 ∈ U. By Proposi- tion 5.11, there exists H ⊇ X which is a Gδ in Z, such that U ⊇ K × H. � Lemma 5.17. Let U be a family of open sets in K × Z which covers K × X, with X ⊂ Z. Let π : K × Z → Z be the projection map. If (i) X is Lindelöf, (ii) K is Lindelöf, (iii) K = ⋃ n∈ω Kn with Kn closed and π↾Kn × Z is closed, then there exists a countable subfamily of U that covers K × X. Corollary 5.18. Let K be a σ- compact space, X a Lindelöf subset of a topo- logical space Z. Let U be a family of open subsets in K × Z which cover K × X, then there exists H ⊇ H which is a Gδ in Z, and U0 ⊆ U countable which cover K × H. 258 A. Caserta and S. Watson 6. Lindelof Haydon planks In this section we construct a Dowker-Style plank, i.e., a variation of Dowker’s idea of 1955 in which we take the subspace of all points in the product lying below the graph of a function (see [3]). Planks have been extensively studied by Watson in [9]. Definition 6.1. Let X, Z be topological spaces, θ a cardinal, h : Z → θ + 1 an arbitrary function, and  : X → θ + 1 surjective. Define the plank Y,h = {(x, z) ∈ X × Z : h(z) ≥ (x)} For every ξ ≤ γ ≤ θ denote Y,h↾(ξ, ·) = {(x, z) ∈ Y,h : (x) < ξ} and Y,h↾(ξ, γ) = {(x, z) ∈ Y,h : (x) < ξ ∧ h(z) < γ}. We investigate more in detail the relation between the plank and the func- tions. In the following, unless we state otherwise, we assume that the X, Z and the function h and  are defined as in the Definition 6.1 Proposition 6.2. Let α ≤ θ, if  has the property NL at α, then ( ∃B Lindelöf : Y,h↾(α, α) ⊆ B ⊆ Y,h ⇒ h is M ichael at α). Proof. Let α ≤ θ with cfα > ω, and B Lindelöf subset of Y,h such that Y,h ↾(α, α) ⊆ B. Let F ⊂ Z be closed such that for every z ∈ F, h(z) < α. Then B ∩ (X × F ) is Lindelöf. Let A = πX (B ∩ (X × F )). Thus A is a Lindelöf subset of X, such that for every x ∈ A (x) < α. From  NL at α we have (A) is not cofinal in α, i.e, there exist β < α such that for every x ∈ A (x) ≤ β. Since  is surjective, for every z ∈ F we can choose x ∈ X with (x) = h(z). Then (x, z) ∈ Y,h↾(α, α) ∩ (X × F ), therefore (x, z) ∈ B ∩ (X × F ). It follows that x ∈ A. Hence for every z ∈ F there exists x ∈ A with (x) = h(z) ≤ β. Thus supz∈Fh(z) ≤ β < α, i.e., h is Michael at α. � Corollary 6.3. Let θ be a cardinal. Assume that  has the property NL at each α ≤ θ. If Y,h is Lindelöf then for each α ∈ θ, h is Michael at α. Proof. Let α ∈ θ with cfα > ω. Assume by contradiction that h is not Michael at α. Then Y,h is Lindelöf and Y,h↾(α, α) ⊂ Y,h. From Proposition 6.2 follows that h is Michael at α. � From Proposition 3.9 and Proposition 6.2 follow: Corollary 6.4. Let θ be a cardinal with cfθ > ω. If (i)  has the property NL at θ, (ii) for each α < θ, h(Z) ∩ (α, θ) 6= ∅, then Y,h × (Z \ h −1({θ}) is not Lindelöf. Now we want to investigate when the plank Y,h is Lindelöf, and we give an inductive proof. First we need the following lemma. Michael spaces and Dowker planks 259 Lemma 6.5. Let U be a family of open sets in −1([0, α]) × Z which covers Y,h↾(α + 1, ·). Let π :  −1([0, α]) × Z → Z be the projection map. If (i) there exists U0 ∈ U which covers  −1([0, α]) × h−1([α, θ]), (ii) h is Michael at α, (iii) foe each ξ < α, Y,h↾(ξ + 1, ·) is Lindelöf;, (iv) π is σ-closed, then there exists a countable subfamily of U that covers Y,h↾(α + 1, ·). Proof. Note that Y,h↾(α+1, ·) = ( −1([0, α])×h−1([α, θ]))∪( ⋃ ξ<α Y,h↾(ξ+1, ·). Let U0 ∈ U that covers  −1([0, α]) × h−1([α, θ]). If cfα = ω, there exists an increasing sequence of ordinal (αn)n∈ω such that⋃ ξ<α Y,h↾(ξ + 1, ·) = ⋃ n∈ω Y,h↾(α + 1, ·), therefore there exists U1 ⊂ U which cover ⋃ ξ<α Y,h↾(ξ + 1, ·). Then U1 ∪ {U0} is a countable subcover of U that covers Y,h↾(α + 1, ·). Assume that cfα > ω. Let U c0 = ( −1([0, α]) × Z) \ U0. Then U c 0 is closed in −1([0, α]) × Z and for every (x, z) ∈ U c0 ∩ Y,h↾(α + 1, ·) we have that (x) < α and h(z) < α. From (iv) follow that C = π(U c0 ) is an Fσ subset of Z, and for every z ∈ C h(z) < α. From (ii) follow that δ = supz∈Ch(z) < α. We claim that Y,h↾(α + 1, ·) \ U0 ⊆ Y,h↾(δ + 1, ·). Let (x, z) ∈ Y,h↾(α + 1, ·) \ U0. From (x, z) ∈ U c0 , follow that z ∈ C; from h(z) < δ and (x) ≤ h(z), follow that (x, z) ∈ Y,h↾(δ + 1, ·). By hypothesi, Y,h↾(δ + 1, ·) is Lindelöf, therefore there exists a countable subfamily U1 ⊂ U which is a cover for a Y,h↾(δ + 1, ·). Thus {U0} ∪ U1 is a countable subcover for U that covers Y,h↾(α + 1, ·). � Proposition 6.6. Let Z be a Lindelöf space, θ a cardinal and α ≤ θ. If (i) h is Michael at α, (ii) h−1([α, θ]) is Lindelöf, (iii) −1([0, α]) is Lindelöf, (iv) for each ξ < α, Y,h↾(ξ + 1, ·) is Lindelöf, (v) π : −1([0, α]) × Z → Z is strongly σ-closed, then Y,h↾(α + 1, ·) is Lindelöf. Proof. Let U be a cover of Y,h↾(α + 1, ·) made by open sets of  −1([0, α]) × Z, and without loss of generality we can assume that it is closed under countable union. Note that Y,h ↾ (α + 1, ·) = ( −1([0, α]) × h−1([α, θ])) ∪ ( ⋃ ξ<α Y,h ↾ (ξ + 1, ·). From Corollary 5.15, there exists U0 ⊂ U countable such that it covers −1([0, α]) × h−1([α, θ]). Let U0 = ⋃ U0, then U0 ∈ U. Lemma 6.5 ends the proof. � Proposition 6.7. Let Z be a Lindelöf space, θ a cardinal, and α ≤ θ. If for each β ≤ α (i) h is Michael at β, (ii) h−1([β, θ]) is Lindelöf, (iii) −1([0, β]) is Lindelöf, (iv) π : −1([0, β]) × Z → Z is strongly σ-closed, then Y,h↾(α + 1, ·) is Lindelöf. 260 A. Caserta and S. Watson Proof. Assume that h is Michael at β for every β ≤ α. From Proposition 6.6, it remains to show that Y,h↾(β + 1, ·) is Lindelöf for every β < α. Suppose not, there exists β < α such that Y,h↾(β + 1, ·) is not Lindelöf, and assume that β is the minimum ordinal with this property. Then, for every γ < β, Y,h↾(γ + 1, ·) is Lindelöf, and for every γ < β, h is Michael at γ. From Proposition 6.6 follow that Y,h↾(β + 1, ·) is Lindelöf, a contradiction. � Theorem 6.8. Let Z be a Lindelöf space, θ a cardinal. If for each α ≤ θ (i) h is Michael at α;, (ii) h−1([α, θ]) is Lindelöf;, (iii) −1([0, α]) is Lindelöf, (iv) π : −1([0, α]) × Z → Z is strongly σ-closed, then Y,h is Lindelöf. Note that the problem to determinate when given an arbitrary topological space Y , the product Y,h × Y is Lindelöf becomes a problem to find condition on ĥ and  so that Y ,ĥ is a Lindelöf space where ĥ : Z × Y → θ + 1. Simply applying Proposition 6.2 and Corollary 6.8 to ĥ the following corol- lary give us conditions to determinate when Y,h × Y is Lindelöf. Corollary 6.9. Let Z be a Lindelöf space, X, Y a topological spaces, θ a cardinal, ĥ : Z × Y → θ + 1. If for each α ≤ θ (i) ĥ is Michael at α, (ii) ĥ−1([α, θ]) × Y is Lindelöf, (iii) −1([0, α]) is Lindelöf, (iv) π : ̂−1([0, α]) × Z × Y → Z × Y is strongly σ-closed, then Y,h × Y is Lindelöf. The next Theorem give us a necessary condition to find a Michael space: Theorem 6.10. Let X be a topological space, θ a cardinal with uncountable cofinality and h : C → θ + 1. If (i) QC = h −1({θ}), (ii) for each α ≤ θ, h is Michael at α, (iii) for each α < θ, h(C) ∩ (α, θ) 6= ∅, (iv)  has the property NL at θ, (v) for each α ≤ θ, −1([0, α]) is Lindelöf, (vi) for each α ≤ θ, π : −1([0, α]) × Z → Z is strongly σ-closed, then Y,h is a Michael space. 7. Special cases One special case is obtained choosing X = θ+1 and the map  as the identity map on θ + 1. The plank Y,h becomes Yh = {(α, z) ∈ (θ + 1) × Z : h(z) ≥ α} Michael spaces and Dowker planks 261 subset of (θ + 1) × Z, and it is an Haydon Plank [(see [5]). For every α ∈ θ denote Yh↾α = {(δ, z) ∈ Yh : δ < α}. In this case, the plank is characterized as Lindelöf, and it is also an example of Michael space. Theorem 7.1. Let Z be a Lindelöf space, h : Z → θ + 1 a function, θ a cardinal with cfθ > ω. Then Yh is Lindelöf if and only if for every α ≤ θ (i) h is Michael at α; (ii) h−1([α, θ]) is Lindelöf. Proof. Let Yh be Lindelöf. Then, for every α ∈ θ, Yh ↾α + 1 is Lindelöf. By Proposition 6.2, h is Michael at α for every α ∈ θ. Moreover Yh ∩ ({α} × Z) ∼= h−1([α, θ]). Theorem 6.8 ends the proof. � Corollary 7.2. Let θ be a cardinal with cfθ > ω, h : C → θ + 1 an arbitrary function. Then Yh is Lindelöf if and only if for each α ≤ θ, h is Michael at α. Lemma 7.3. Let Y be a topological space, θ cardinal with cfθ > ω. If Yh × Y is Lindelöf, then for every α≤θ, ĥ is Michael at α, where ĥ : Z × Y → θ + 1. Moreover, if h−1([α, θ]) × Y is Lindelöf for every α ≤ θ, then the converse holds. Proof. Follows from Corollary 6.8 (applied to ĥ), Proposition 6.2 and Re- mark 4.2. � Corollary 7.4. Let Y be a topological space, θ cardinal, h : Z → θ + 1 a func- tion such that for every α ≤ θ h−1([α, θ]) × Y is Lindelöf. Then the following statements are equivalent: (i) h is A(Y )-Michael function, (ii) Yh × Y is Lindelöf. Proof. Follows from Theorem 7.3 and Proposition 3.7. � Corollary 7.5. Let θ be a cardinal with uncountable cofinality, Z a Lindelöf space and h : Z → θ + 1 a function such that for every α < θ h(Z)∩(α, θ) 6= ∅. Then Yh × (Z \ h −1({θ})) is not Lindelöf Proof. Follows from Corollary 6.4. � Corollary 7.6. Let θ be a cardinal with uncountable cofinality, Z a Lindelöf space and h : Z → θ + 1 a function. If (i) for each α ≤ θ, h is Michael at α, (ii) for each α < θ, h(Z) ∩ (α, θ) 6= ∅, (iii) for each α ≤ θ, h−1([α, θ]) is Lindelöf, then Yh ⊆ (θ + 1) × Z is a Lindelöf space such that Yh × (Z \ h −1({θ})) is a non-Lindelöf space. Proof. Follows from Corollary 6.8 and Corollary 7.5. � 262 A. Caserta and S. Watson Theorem 7.7. Let θ be a cardinal with uncountable cofinality and h a K- Michael function defined on C such that (i) QC = h −1({θ}), (ii) h(C) ∩ (α, θ) 6= ∅ for every α < θ. Then Yh, subspace of (θ + 1) × C, is a Michael space. We give some other examples of planks which are Michael spaces. Definition 7.8. A special plank is given by choosing X = ∏ n∈ω θ + 1 with θ of uncountable cofinality, and the map  : X → θ + 1 defined by (f ) = min{ξ ∈ θ + 1 : f ≤ fξ}, where fα is a constant function with value α for every α ≤ θ. We denote this plank Y P,h. Theorem 7.9. Let θ be a cardinal with cfθ > ω and h a K- Michael function defined on C such that (i) QC = h −1({θ}), (ii) for every α < θ, h(C) ∩ (α, θ) 6= ∅. Then Y P,h is a Michael space. Proof. By Corollary 4.8, follow that the map  has the property NL at α for every α ≤ θ. Moreover, for every α ≤ θ, −1([0, α]) = {f ∈ ∏ n∈ω θ + 1 : ∀n ∈ ω f (n) ≤ α} is a compact subset of ∏ n∈ω θ + 1. By Lemma 5.8, the projection π : −1([0, α]) × Z → Z is strongly σ-closed. Theorem 6.10 ends the proof. � Another special plank is obtained for a particular choice of the map . Definition 7.10. Let X = ∏ n∈ω θn + 1 with every θn cardinal with uncount- able cofinality, and  : X → κ + 1 defined by (f ) = min{ξ ∈ κ : f ≤∗ fξ} where κ is a cardinal with cfκ > ω and {fξ}ξ∈κ is a dominating family in ( ∏ n∈ω θn, ≤∗). We denote this plank Y ∏ ,h . Remark 7.11. The definition of a dominating family and the definition of the map  in the plank Y ∏ ,h , imply that κ > θn for every n ∈ ω. Indeed considering the special case in which the family {fξ}ξ∈κ is a family of constant functions, we need to have the function which assumes constant value θn. Therefore κ > θn + 1 for every n ∈ ω. Remark 7.12. Let (X, ≤) be a partial order, F ⊆ X with F = {fξ}ξ∈κ a dominating family in X, (i.e. for all x ∈ X, there exists fξ ∈ F such that x ≤ fξ). Define  : X → F by (x) = min{fξ ∈ F : x ≤ fξ}. We have that  is surjective if and only if fα � fβ for every α < β. Remark 7.13. Let X = ∏ n∈ω θn + 1 with every θn cardinal with uncountable cofinality, κ a cardinal with cfκ > ω and {fξ}ξ∈κ a dominating family in ( ∏ n∈ω θn, ≤∗). The map  : X → κ + 1 defined by (f ) = min{ξ ∈ κ : f ≤∗ fξ} might not be surjective. Since F ′ = {fξ ∈ F : fξ ∈ (X)} is still a dominating family of X, when (X) has order type κ, we can assume without loss of generality that  is surjective. Further,if the dominating family is a scale of X, Michael spaces and Dowker planks 263 we can consider F ′ , the dominating family of minimum cardinality which is a scale, i.e., |F ′ | = d. Such a family is a dominating family with order type d and the map  ′ : X → F ′ defined by  ′ (x) = min{fξ ∈ F ′ : x ≤ fξ} is surjective. An example of Y ∏ ,h -plank is given by cardinal of countable cofinality. Indeed, from the Theorem of Shelah [B.M. [1]], given θ with cfθ = ω, there exists an increasing sequence of regular cardinals {θn}n∈ω cofinal in θ, and a scale {fξ}ξ∈θ+ on ( ∏ n∈ω θn, ≤∗). In this case choose X = ∏ n∈ω θn + 1 and the map  : X → θ+ + 1 defined by (f ) = min{ξ ∈ θ+ : f ≤∗ fξ}. Then we have: Theorem 7.14. Let θ be a cardinal with cfθ > ω and h a K- Michael function defined on C such that (i) QC = h −1({θ}), (ii) for every α < θ, h(C) ∩ (α, θ) 6= ∅. Then Y ∏ ,h is a Michael space. Proof. For every α ∈ κ, we have −1([0, α]) = {f ∈ X : f ≤∗ fα} = ⋃ F ∈[ω]<ω {f ∈ X : ∀n /∈ F f (n) ≤ fα(n)}. Therefore for every α ∈ κ, the set  −1([0, α]) ⊂ X is σ-compact, hence by Lemma 5.8, the projection map π : −1([0, α]) × Z → Z is strongly σ-closed. Moreover from Proposition 4.11, the map  has the property NL at κ. Theorem 6.10 ends the proof. � 8. The cardinal L If X is a non-Lindelöf space, L(X) denote the minimum cardinality of an uncountable open cover of X with no countable subcover, and if X is Lindelöf, define L(X) = ∞. Note that for a non-Lindelöf space, L(X) ≤ w(X), where w(X) denote the weight of the topological space X, and L(X) is either a regular cardinal or has countable cofinality. The following lemma give us some relations between the L cardinals of re- lated spaces. Lemma 8.1. Let X, Y be topological spaces. The following properties hold: (i) If X is Lindelöf and X × Y is not Lindelöf, then L(X × Y ) ≤ |Y |. (ii) If F ⊆ X is closed and not Lindelöf space, then L(X) ≤ L(F ). (iii) If f is a continuous open map, such that f (X) is not Lindelöf space, then L(f (X)) = L(X). Proof. (i) Let U be an open cover of X × Y witnessing L(X × Y ). For every y ∈ Y , let U(y) = {Un(y) : n ∈ ω} ⊂ U be a countable open subcover of X × {y}. Thus V = {Un(y) : n ∈ ω ∧ y ∈ Y } ⊂ U is an open cover of X × Y , such that |V| ≤ |Y | with no countable subcover. Therefore L(X × Y ) ≤ |Y |. (ii) Let U be an open cover of F with |U| = L(F ) with no countable subcover. Then U ∪ {F c} is an open cover for X of the same kind. 264 A. Caserta and S. Watson (iii) Let U be an open cover of f (X) with |U| = L(f (X)) with no countable subcover. Then f −1(U) is an open cover for X. Thus L(X) ≤ L(f (X)). If V is an open cover of X with |V| = L(X) with no countable subcover. Then f (V) is an open cover for f (X) of the same kind. � Lemma 8.2. Let X, Y be topological with X Lindelöf. For every F ⊆ Y closed such that L(X × Y ) > |F |, X × F is Lindelöf. Proof. If X × Y is Lindelöf, then L(X × Y ) = ∞, and X × F is Lindelöf. Now, assume that X ×Y and X ×F are not Lindelöf. Since X ×F is closed in X ×Y , from Lemma 8.1 we have that |F | < L(X × Y ) ≤ L(X × F ). Lemma 8.1 ends the proof. � Corollary 8.3. Let X, Y be topological spaces with X Lindelöf and L(X×Y ) = |Y |. Then for every closed F ⊆ Y with |F | < |Y | follow that X × F is Lindelöf. Lemma 8.4. Let X, Y be topological spaces with X Lindelöf and L(X × Y ) = |Y |, then Y is not union of less than |Y | many closed subsets of Y with cardi- nality less than |Y |. Proof. Let U = {Uξ}ξ<|Y | be an open cover of X ×Y witnessing L(X ×Y ). Let κ cardinal with κ < |Y |. Assume by contradiction that Y = ⋃ ξ∈κ Yξ where for every ξ ∈ κ Yξ are closed in Y and |Yξ| < |Y |. Therefore from Corollary 8.3 follow that X × Yξ is Lindelöf for every ξ ∈ κ, and so there exists Uξ ⊂ U countable subcover of X × Yξ . Set V = {Uξ : ξ ∈ κ} ⊂ U. Then V ⊆ U is an open cover of X × Y of size κ. From L(X × Y ) = |Y | follow that there exist a countable subcover V ′ ⊂ V of X × Y . Then V ′ is also a countable subcover from U which is a contradiction. � Let X, Y be topological spaces, θ a cardinal, and P (X, Y, θ) states that X is a Lindelöf space such that X × Y is not Lindelöf space and L(X × Y ) = θ. Theorem 8.5. Let X be a topological space and θ a cardinal. If Y satisfies P (X, Y, θ) and |Y | < κ, with κ infinite cardinal, then there exists Y ′ which satisfies P (X, Y ′ , θ) and |Y ′ | = κ. Proof. Let Y ′ = Y ⊕ αD(κ), where αD(κ) is the one-point compactification of a discrete set of cardinality κ. Clearly |Y ′ | = κ. Since the space X × Y is a closed subset of X × Y ′ , it follows that X × Y ′ is not Lindelöf. It remains to show that L(X×Y ′ ) = θ, assuming that L(X×Y ) = θ. Since the space X×Y is a closed subset of X × Y ′ , from Lemma 8.1 follow that L(X × Y ) ≤ L(X × Y ′ ), and so L(X × Y ′ ) ≥ θ. Now, let U be an open cover for X × Y of size θ with no countable subcover. We have that X × Y ′ is homeomorphic to (X × Y ) ⊕ (X × αD(κ)) hence it follows that U is an open family in X × Y ′ such that U ∩ (X × αD(κ)) = ∅ for every U ∈ U . Let V = U ∪ {X × αD(κ)}. Then V is an open cover of Y ′ of size θ with no countable subcover. � Michael spaces and Dowker planks 265 Remark 8.6. In other words we have that for a fixed topological space X and a cardinal θ, if there exists Y such that P (X, Y, θ), then the set AX,θ = {κ : κ is cardinal ∧ ∃Y P (X, Y, θ) ∧ |Y | = κ} is non empty and AX,θ = [min AX,θ, +∞). We conclude this work showing that if there is a Michael space, then under some conditions involving singular cardinals, there must be one which is a Haydon plank. Theorem 8.7. Let X be a Lindelöf space, Y a topological space such that X × Y is not Lindelöf, θ a cardinal with cfθ = ω and L(X × Y ) = |Y | = θ. Let cY be any compactification of Y . Then there exists a function f : cY → θ + 1 such that (i) f −1({θ}) = cY \ Y , (ii) for every α ≤ θ, f is Michael at α, (iii) for every α < θ, f (cY ) ∩ (α, θ) 6= ∅. Proof. Let θ = L(X × Y ), and U be an open cover of X × Y witnessing L(X × Y ). Fix an enumeration {yξ}ξ<θ of Y of order type θ. Given y ∈ Y , let U(y) = {Un(y) : n ∈ ω} ⊂ U a countable open subcover of X × {y}. Thus V = {Un(y) : n ∈ ω ∧ y ∈ Y } ⊂ U is an open cover of X ×Y , such that |V| = θ. Let cY be a compactification of Y . Define the function f : cY → θ + 1 as follows: for every y ∈Y , f (y) = sup{γ ∈ θ : X ×{y} * ∪ξ<γ (∪n∈ωUn(yξ))} and for every y ∈ cY \Y f (y) = θ. Then, by definition of V, there is not y ∈ Y such that X × {y} * ∪ξ<α(∪n∈ωUn(yξ)) for every α ≤ θ. Thus f −1({θ}) = cY \ Y . Let α ∈ θ with cfα > ω, and F ⊂ cY closed such that f (y) < α for every y ∈ F . Assume by contradiction that supy∈Ff(y) = α. By definition of α we have that X × F ⊆ ∪ξ<α(∪n∈ω Un(yξ)). Then {Un(yξ) : n ∈ ω ∧ ξ < α} is an uncountable cover of X × F with F compact. We want to show that it has no countable subcover which contradict X × F to be Lindelöf. Indeed if {Um(yξn )}n,m∈ω was a countable subcover of X × F . Let ν = supn∈ωξn. Since cfα > ω, ν < α. By definition of ν there exists y ∈ F such that X × {y} * ∪n∈ωUm(yξn ) which is a contradiction. Now, by contradiction, there exists α ∈ θ such that f (cY ) ∩ (α, θ) = ∅, i.e., there exists α ∈ θ such that for every y ∈ Y, f (y) < α. Therefore for every y ∈ Y, X × {y} ⊆ ∪ξ<α(∪n∈ωUn(yξ)). Thus {Un(yξ) : n ∈ ω ∧ ξ < α} is an open cover of X × Y with α < θ. By definition of L(X × Y ) = θ, there exists a countable subcover for X × Y from {Un(yξ) : n ∈ ω ∧ ξ < α}, and therefore from U, which is a contradiction. � Theorem 8.8. Let X be a Lindelöf space such that X × Y is not Lindelöf, θ a regular cardinal such that L(X × Y ) = θ. Let cY be any compactification of Y . Then there exists a function f : cY → θ + 1 such that (i) f −1({θ}) = cY \ Y , (ii) for every α ≤ θ, f is Michael at α, (iii) for every α < θ, f (cY ) ∩ (α, θ) 6= ∅. 266 A. Caserta and S. Watson Proof. Let θ = L(X × Y ). Fix an enumeration {Uξ}ξ<θ of an open cover of X × Y witnessing L(X × Y ). Let cY be a compactification of Y . Define the function f : cY → θ + 1 as follows: for every y ∈ Y , f (y) = sup{γ ∈ θ : X × {y} * ∪ξ<γ Uξ} and for every y ∈ cY \ Y f (y) = θ. Since θ is regular and {Uξ}ξ<θ is an open cover of X × Y , there is not y ∈ Y such that for every α ≤ θ X × {y} * ∪ξ<αUξ. Thus f −1({θ}) = cY \ Y . Let now α ∈ θ with cfα > ω, and F ⊂ cY closed such that f (y) < α for every y ∈ F . Assume by contradiction that supy∈Ff(y) = α. By definition of α we have that for every β ≥ α, X × {y} ⊆ ∪ξ<β Uξ for every y ∈ F , therefore X × F ⊆ ∪ξ<αUξ. Then {Uξ}ξ<θ is an uncountable cover of X × F with F compact. We want to show that it has no countable subcover which contradict X × F to be Lindelöf. Indeed if {Uξn}n∈ω was a countable subcover of X × F . Let ν = supn∈ωξn. Since cfα > ω, ν < α. By definition of ν there exists y ∈ F such that X × {y} * ∪n∈ωUξn , contradiction. Now, by contradiction, there exists α ∈ θ such that f (cY ) ∩ (α, θ) = ∅, i.e., there exists α ∈ θ such that for every y ∈ Y, f (y) < α. Therefore for every y ∈ Y, X × {y} ⊆ ∪ξ<αUξ. Thus {Uξ}ξ<α is an open cover of X × Y with α < θ. By definition of L(X × Y ) = θ, there exists a countable subcover for X ×Y from {Uξ}ξ<α, and therefore from {Uξ}ξ<θ, which is a contradiction. � In the Theorem 8.8, θ is a regular cardinal, we do not need any assumption about the cardinality of Y (as in Theorem 8.7) because the open cover{Uξ}ξ<θ witnessing L(X × Y ) = θ is never cofinal. This guarantee that f −1({θ}) = cY \ Y . From Theorem 7.7 it follows: Theorem 8.9. Let M be a Michael space, θ a regula cardinal such that L(M × P) = θ. Then there exists a function f and Yf ⊆ (θ + 1) × C which is a Michael space. The aim of Proposition 8.7 and Proposition 8.8 is to produce the following statement: given a Lindelöf space X such that L(X × Y ) = θ, there exists f and Yf ⊆ (θ + 1) × cY such that Yf is Lindelöf and Yf × Y is not Lindelöf, where cY is any compactification of Y . We require the property that for all α ≤ θ, f −1([α, θ]) is Lindelöf. Clearly this is always true when Y admits an hereditarily Lindelöf compactification. When does the property hold? Is there a function f : cY → θ + 1 such that satisfy the property of Proposi- tion 8.7 when X is a Lindelöf space, Y a topological space such that X × Y is not Lindelöf, θ a cardinal of countable cofinality such that L(X × Y ) = θ, and |Y | > θ? Michael spaces and Dowker planks 267 References [1] M. Burke and M. Magidor, Shelah’s pcf theory and its applications, Ann. Pure Appl. Logic 50, (1990) 207–254. [2] E. K. van Douwen, The integers and topology, in K. Kunen and J. Vaughan, editors, Handbook of Set-Theoretic Topology, 111–169, North-Holland, Amsterdam, (1984). [3] C. H. Dowker, Local dimension of normal spaces, Quart. J. Math. Oxford 2 (1990), no. 6, 101–120. [4] R. Engelking, General Topology, Heldermann Verlag, Berlin 1989. [5] R. Haydon, On compactness in spaces of measure and measure compact spaces, Proc. London Math. Soc. 29 (1974), no. 6, 1–16. [6] K. Kunen, Set Theory. An Introduction to Independence Proofs, North-Holland, Ams- terdam 1980. [7] E. Michael, The product of a normal space and a metric space need not be normal, Bull. Amer. Math. Soc. 69 (1963), 375–376. [8] J. Tatch Moore, Some of the combinatorics related to Michael’s problem, Proc. Amer. Math. Soc. 127, (1999), no. 8, 2459–2467. [9] S. Watson, The Construction of Topological Spaces: Planks and Resolutions, in M. Hus̄ek and J. van Mill (eds.), Recent Progress in General Topology, 673–757, North- Holland 1992. Received January 2009 Accepted June 2009 A. Caserta (agata.caserta@unina2.it) Dipartmento di Matematica, Seconda Universitá degli Studi di Napoli, Caserta 81100, Italia S. Watson (watson@hilbert.math.yorku.ca) Department of Mathematics and Statistics, York University, Toronto M3J1P3, Canada