This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Countably compact group topologies on arbitrarily large free Abelian groups

Matheus K. Bellini Klaas Pieter Hart Faculty EEMCS
TU Delft
Postbus 5031
2600 GA Delft
the Netherlands
[email protected] http://fa.its.tudelft.nl/~hart
Vinicius O. Rodrigues  and  Artur H. Tomita Depto de Matemática, Instituto de Matemática e Estatística, Universidade de São Paulo, Rua do Matão, 1010 – CEP 05508-090, São Paulo, SP - Brazil [email protected] [email protected] [email protected]
Abstract.

We prove that if there are 𝔠\mathfrak{c} incomparable selective ultrafilters then, for every infinite cardinal κ\kappa such that κω=κ\kappa^{\omega}=\kappa, there exists a group topology on the free Abelian group of cardinality κ\kappa without nontrivial convergent sequences and such that every finite power is countably compact. In particular, there are arbitrarily large countably compact groups. This answers a 1992 question of D. Dikranjan and D. Shakhmatov.

Key words and phrases:
Topological group, countable compactness, selective ultrafilter, free Abelian group, Wallace’s problem
2010 Mathematics Subject Classification:
Primary 54H11, 22A05; Secondary 54A35, 54G20.
†The first author received financial support from FAPESP 2017/15709-6.
‡The third author received financial support from FAPESP 2017/15502-2. Corresponding Author.
¶The fourth author received financial support from FAPESP 2016/26216-8. The fourth author thanks the second author for the hospitality during his visit to TU Delft in April 2019.

1. Introduction

1.1. Some history

It is well known that a non-trivial free Abelian group does not admit a compact Hausdorff group topology. Tomita [22] showed that it does not admit even a group topology whose countable power is countably compact .

Tkachenko [20] showed in 1990 that the free Abelian group generated by 𝔠\mathfrak{c} elements can be endowed with a countably compact Hausdorff group topology under CH. Tomita [22], Koszmider, Tomita and Watson [15], and Madariaga-Garcia and Tomita [17] obtained such examples using weaker assumptions. Boero, Castro Pereira and Tomita obtained such an example using a single selective ultrafilter [2]. Using 2𝔠2^{\mathfrak{c}} selective ultrafilters, the example in [17] showed the consistency of a countably compact group topology on the free Abelian group of cardinality 2𝔠2^{\mathfrak{c}}. All forcing examples obtained so far had their cardinalities bounded by 2𝔠2^{\mathfrak{c}}.

Boero and Tomita [4] showed from the existence of 𝔠\mathfrak{c} selective ultrafilters that there exists a free Abelian group of cardinality 𝔠\mathfrak{c} whose square is countably compact. Tomita [26] showed that there exists a group topology on the free Abelian group of cardinality 𝔠\mathfrak{c} that makes all its finite powers countably compact.

E. van Douwen showed in [8] that the cardinality of a countably compact group cannot be a strong limit of countable cofinality.

Using the result in the abstract, we obtain the following:

Theorem 1.1.

Assume GCH\mathrm{GCH}. Then a free Abelian group of infinite cardinality κ\kappa can be endowed with a countably compact group topology (without non-trivial convergent sequences) if and only if κ=κω\kappa=\kappa^{\omega}.

The result above answers a question of Dikranjan and Shakhmatov that was posed in the survey by Comfort, Hoffman and Remus [6].

Because of the way our examples are constructed we can raise their weights in the same way as in the papers [23] or [5] and obtain the following result – the examples in these references are Boolean but the trick is similar.

Theorem 1.2.

It is consistent that there is a proper class of cardinals of countable cofinality that can occur as the weight of a countably compact free Abelian group.

1.2. Basic results, notation and terminology

We recall that a topological space is countably compact if, and only if, every countable open cover of it has a finite subcover.

Definition 1.3.

Let 𝒰\mathcal{U} be a filter on ω\omega and let (xn:nω)(x_{n}:n\in\omega) be a sequence in a topological space XX. We say that xXx\in X is a 𝒰\mathcal{U}-limit point of (xn:nω)(x_{n}:n\in\omega) if, for every neighborhood UU of xx, the set {nω:xnU}\{n\in\omega:x_{n}\in U\} belongs to 𝒰\mathcal{U}.

If XX is Hausdorff, every sequence has at most one 𝒰\mathcal{U}-limit and we write x=𝒰lim(xn:nω)x=\mathcal{U}\mathchar 45\relax\!\lim(x_{n}:n\in\omega) in that case.

The set of all free ultrafilters on ω\omega is denoted by ω\omega^{*}. The following proposition is a well known result on ultrafilter limits.

Proposition 1.4.

A topological space is countably compact if and only if each sequence in it has a 𝒰\mathcal{U}-limit point for some 𝒰ω\mathcal{U}\in\omega^{*}.

The concept of almost disjoint families will be useful in our construction.

Definition 1.5.

An almost disjoint family is an infinite family 𝒜\mathcal{A} of infinite subsets of ω\omega such that distinct elements of 𝒜\mathcal{A} have a finite intersection.

It is well known that there exists an almost disjoint family of size continuum (see [16]).

Definition 1.6.

The unit circle group 𝕋\T will be the metric group (/,δ)(\R/\Z,\delta) where the metric δ\delta is given by

δ(x+,y+)=min{|xy+a|:a}\delta(x+\Z,y+\Z)=\min\{|x-y+a|:a\in\Z\}

for every x,yx,y\in\R.

Given an open interval (a,b)(a,b) of \R with a<ba<b, we let δ((a,b))=ba\delta((a,b))=b-a.

An arc of 𝕋\T is a set of the form I+={a+:aI}I+\Z=\{a+\Z:a\in I\}, where II is an open interval of \R. An arc is said to be proper if it is distinct from 𝕋\T.

If UU is a proper arc and U={a+:aI}={b+:aJ}U=\{a+\Z:a\in I\}=\{b+\Z:a\in J\}, then the Euclidean length of II equals the Euclidean length of JJ, and we define the length of UU as δ(U)=δ(I)\delta(U)=\delta(I). We also let δ(𝕋)=1\delta(\T)=1.

Given an arc UU such that δ(U)12\delta(U)\leq\frac{1}{2}, it follows that diamδU=δ(U)\operatorname{diam}_{\delta}U=\delta(U).

Our free Abelian groups will all be represented as directs sums of copies of the group of integers \Z; we fix some notation. The additive group of rationals will also be used, so in the following definition one should read \Z or \Q for GG.

Definition 1.7.

If ff is a map from a set XX to a group GG then the support of ff, denotes suppf\operatorname{supp}f is defined to be the set {xX:f(x)0}\{x\in X:f(x)\neq 0\}.

We define G(X)={fGX:|suppf|<ω}G^{(X)}=\{f\in G^{X}:|\operatorname{supp}f|<\omega\}.

If YY is a subset of XX then, as an abuse of notation, we often write G(Y)={xG(X):suppxY}G^{(Y)}=\{x\in G^{(X)}:\operatorname{supp}x\subseteq Y\}.

Given xXx\in X, we denote by χx\chi_{x} the characteristic function of {x}\{x\}, whose support is {x}\{x\} and which value χx(x)=1\chi_{x}(x)=1.

For a sequence ζ:ωX\zeta:\omega\to X in XX we define χζ:ωGX\chi_{\zeta}:\omega\to G^{X} by χζ(n)=χζ(n)\chi_{\zeta}(n)=\chi_{\zeta(n)}.

Finally, for xXx\in X, we let x:ωX\vec{x}:\omega\to X be the constant sequence with value xx.

Note that then χx\chi_{\vec{x}} is also constant, with value χx\chi_{x}.

Definition 1.8.

Let 𝒰\mathcal{U} be a filter on ω\omega and XX a set. We say that the sequences f,gXωf,g\in X^{\omega} are 𝒰\mathcal{U}-equivalent and write f𝒰gf\equiv_{\mathcal{U}}g iff {nω:f(n)=g(n)}𝒰\{n\in\omega:f(n)=g(n)\}\in\mathcal{U}.

It is easy to verify that 𝒰\equiv_{\mathcal{U}} is an equivalence relation. We denote the equivalence class of fXωf\in X^{\omega} by [f]𝒰[f]_{\mathcal{U}}. We also denote the set of all equivalence classes by Xω/𝒰X^{\omega}/\mathcal{U}.

If RR is a ring and XX is an RR-module, then Xω/𝒰X^{\omega}/\mathcal{U} has a natural RR-module structure given by [f]𝒰+[g]𝒰=[f+g]𝒰[f]_{\mathcal{U}}+[g]_{\mathcal{U}}=[f+g]_{\mathcal{U}}, [f]𝒰=[f]𝒰[-f]_{\mathcal{U}}=-[f]_{\mathcal{U}}, r[f]𝒰=[rf]𝒰r\cdot[f]_{\mathcal{U}}=[r\cdot f]_{\mathcal{U}} and the class of the zero function as its zero element.

If pp is a free ultrafilter, then the ultrapower of the RR-module XX by pp is the RR-module Xω/pX^{\omega}/p.

For the rest of this paper we will fix a cardinal number κ\kappa that satisfies κω=κ\kappa^{\omega}=\kappa.

Throughout this article, we will work inside ultrapowers of (κ)\Q^{(\kappa)}. These ultrapowers contain copies of ultrapowers of (κ)\Z^{(\kappa)}, which will be useful for the construction. So it is useful to define some notation.

Definition 1.9.

Let pp be a free ultrafilter on ω\omega. We define Ult(,p)\operatorname{Ult}(\Q,p) as the \Q-vector space ((κ))ω/p(\Q^{(\kappa)})^{\omega}/p and Ult(,p)={[g]p:gω}\operatorname{Ult}(\Z,p)=\{[g]_{p}:g\in\Z^{\omega}\} with the subgroup structure.

Notice that each [g]p[g]_{p} in Ult(,p)\operatorname{Ult}(\Z,p) is formally an element of ((κ))ω/p(\Q^{(\kappa)})^{\omega}/p, not of ((κ))ω/p(\Z^{(\kappa)})^{\omega}/p. Nevertheless it is clear that ((κ))ω/p(\Z^{(\kappa)})^{\omega}/p is isomorphic to Ult(,κ)\operatorname{Ult}(\Z,\kappa) via the obvious isomorphism that carries the equivalence class of a sequence g(κ)ωg\in(\Z^{\kappa})^{\omega} in ((κ))ω/p(\Z^{(\kappa)})^{\omega}/p to its class in ((κ))ω/p(\Q^{(\kappa)})^{\omega}/p.

2. Selective Ultrafilters

In this section we review some basic facts about selective ultrafilters, the Rudin-Keisler order and some lemmas we will use in the next sections.

Definition 2.1.

A selective ultrafilter (on ω\omega), also called Ramsey ultrafilter, is a free ultrafilter pp on ω\omega with the property that for every partition (An:nω)(A_{n}:n\in\omega) of ω\omega, either there exists nn such that AnpA_{n}\in p or there exists BpB\in p such that |BAn|=1|B\cap A_{n}|=1 for every nωn\in\omega.

The following proposition is well known. We provide [14] as a reference.

Proposition 2.2.

Let pp be a free ultrafilter on ω\omega. Then the following are equivalent:

  1. a)

    pp is a selective ultrafilter,

  2. b)

    for every fωωf\in\omega^{\omega}, there exists ApA\in p such that ff is either constant or one-to-one on AA,

  3. c)

    for every function f:[ω]22f:[\omega]^{2}\to 2 there exists ApA\in p such that ff is constant on [A]2[A]^{2}.

The Rudin-Keisler order is defined as follows:

Definition 2.3.

Let 𝒰\mathcal{U} be a filter on ω\omega and f:ωωf:\omega\to\omega. We define f(𝒰)={Aω:f1[A]𝒰}f_{*}(\mathcal{U})=\{A\subseteq\omega:f^{-1}[A]\in\mathcal{U}\}.

It is easy to verify that f(𝒰)f_{*}(\mathcal{U}) is a filter; if 𝒰\mathcal{U} is an ultrafilter then so is f(𝒰)f_{*}(\mathcal{U}); if f,g:ωωf,g:\omega\to\omega, then (fg)=fg(f\circ g)_{*}=f_{*}\circ g_{*}; and (idω)(\text{id}_{\omega})_{*} is the identity over the set of all filters. This implies that if ff is bijective, then (f1)=(f)1(f^{-1})_{*}=(f_{*})^{-1}.

Definition 2.4.

Let 𝒰\mathcal{U} and 𝒱\mathcal{V} be filters. We say that 𝒰𝒱\mathcal{U}\leq\mathcal{V} (or 𝒰RK𝒱\mathcal{U}\leq_{\text{RK}}\mathcal{V}, if we need to be clear) iff there exists fωf\in\omega such that f(𝒰)=𝒱f_{*}(\mathcal{U})=\mathcal{V}.

The Rudin-Keisler order is the set of all free ultrafilters over ω\omega ordered by RK\leq_{\text{RK}}. We say that two ultrafilters pp and qq are equivalent iff pqp\leq q and qpq\leq p.

It is easy to verify that \leq is a preorder and that the equivalence defined above is indeed an equivalence relation. Moreover, the equivalence class of a fixed ultrafilter is the set of all fixed ultrafilters, so the relation restricts to ω\omega^{*} without modifying the equivalence classes. We refer to [14] for the following proposition:

Proposition 2.5.

The following are true:

  1. (1)

    If pp and qq are ultrafilters, then pqp\leq q and qpq\leq p is equivalent to the existence of a bijection f:ωωf:\omega\to\omega such that f(p)=qf_{*}(p)=q.

  2. (2)

    The selective ultrafilters are exactly the minimal elements of the Rudin-Keisler order.

This implies that if f:ωωf:\omega\to\omega and pp is a selective ultrafilter, then f(p)f_{*}(p) is either a fixed ultrafilter or a selective ultrafilter. If f(p)f_{*}(p) is the ultrafilter generated by nn, then f1[{n}]pf^{-1}[\{n\}]\in p, so, in particular, if ff is finite to one and pp is selective, then f(p)f_{*}(p) is a selective ultrafilter equivalent to pp.

The existence of selective ultrafilters is independent from ZFC. Martin’s Axiom for countable orders implies the existence of 2𝔠2^{\mathfrak{c}} pairwise incomparable selective ultrafilters in the Rudin-Keisler order.

The lemma below appears in [24].

Lemma 2.6.

Let (pk:kω)(p_{k}:k\in\omega) be a family of pairwise incomparable selective ultrafilters. For each kk let (ak,i:iω)(a_{k,i}:i\in\omega) be a strictly increasing sequence in ω\omega such that {ak,i:iω}pk\{a_{k,i}:i\in\omega\}\in p_{k} and i<ak,ii<a_{k,i} for all iωi\in\omega. Then there exists {Ik:kω}\{I_{k}:k\in\omega\} such that:

  1. a)

    {ak,i:iIk}pk\{a_{k,i}:i\in I_{k}\}\in p_{k}, for each kωk\in\omega.

  2. b)

    IjIj=I_{j}\cap I_{j}=\emptyset whenever i,jωi,j\in\omega and iji\neq j, and

  3. c)

    {[i,ak,i]:iIk\{[i,a_{k,i}]:i\in I_{k} and kω}k\in\omega\} is a pairwise disjoint family.

In the course of the construction we will often use families of ultrafilters indexed by ω\omega and finite sequences of infinite subsets of ω\omega. The following definition fixes some convenient notation.

Definition 2.7.

A finite tower in ω\omega is a finite sequence (A0,,Ak1)(A_{0},\dots,A_{k-1}) of infinite subsets of ω\omega such that At+1AtA_{t+1}\subseteq A_{t} for every t<k1t<k-1. The set of all finite towers in ω\omega is called 𝒯\mathcal{T}. If T=(A0,,Ak1)T=(A_{0},\dots,A_{k-1}) then l(T)=Ak1l(T)=A_{k-1}, the last term of the sequence TT. For the empty sequence we write l()=ωl(\emptyset)=\omega.

Lemma 2.8.

Assume there are 𝔠\mathfrak{c} incomparable selective ultrafilters. Then there is a family of incomparable selective ultrafilters (pT,n:T𝒯,nω)(p_{T,n}:T\in\mathcal{T},n\in\omega) such that l(T)pT,nl(T)\in p_{T,n} whenever T𝒯T\in\mathcal{T} and nωn\in\omega.

Proof.

Index the 𝔠\mathfrak{c} incomparable selective ultrafilters faithfully as {qT,n:T𝒯,nω}\{q_{T,n}:T\in\mathcal{T},n\in\omega\}. For each TT, let fT:ωl(T)f_{T}:\omega\to l(T) be a bijection and define pT,n=fT(qT,n)p_{T,n}={f_{T}}_{*}(q_{T,n}). Since ff is one-to-one, it follows that pT,np_{T,n} is a selective ultrafilter equivalent to qT,nq_{T,n}. The family (pT,n:T𝒯,nω)(p_{T,n}:T\in\mathcal{T},n\in\omega) is as required. ∎

3. Main Ideas

From now on we fix a family (pT,n:nω,T𝒯)(p_{T,n}:n\in\omega,T\in\mathcal{T}) of selective ultrafilters as provided by Lemma 2.8.

The idea will be to use these ultrafilters to assign pp-limits to enough injective sequences in (κ)\Z^{(\kappa)} to ensure countable compactness of the resulting topology. We take some inspiration from [2] where a large independent family was used such that, up to a permutation every injective sequence in (𝔠){\Z}^{(\mathfrak{c})} was part of this family. Since this group has cardinality 𝔠\mathfrak{c}, there were indeed enough permutations to accomplish this. For an arbitrarily large group, we shall consider large linearly independent pieces to make sure every sequence has an accumulation point.

The following definition will be used to construct a witness for linearly independence in an ultraproduct that does not depend on the free ultrafilter.

Definition 3.1.

Let \mathcal{F} be a subset of ((κ))ω({\Z}^{(\kappa)})^{\omega} and A[ω]ωA\in[\omega]^{\omega}. We shall call \mathcal{F} linearly independent mod AA^{*} iff for every free ultrafilter pp with ApA\in p the set

([f]p:f)˙([χξ]p:ξ<κ)([f]_{p}:f\in\mathcal{F})\mathbin{\dot{\cup}}([\chi_{\vec{\xi}}]_{p}:\xi<\kappa)

is linearly independent in the \Q-vector space Ult(,p)\operatorname{Ult}(\Q,p), and if [f]p[g]p[f]_{p}\neq[g]_{p} whenever ff and gg are distinct elements of \mathcal{F}.

Notice that it is implicit in our definition that {[f]p:f}\{[f]_{p}:f\in\mathcal{F}\} and {[χξ]p:ξ<κ}\{[\chi_{\vec{\xi}}]_{p}:\xi<\kappa\} are disjoint. We will abbreviate “linearly independent mod AA^{*}” to l.i. mod AA^{*}.

An application of Zorn’s Lemma will establish the following lemma.

Lemma 3.2.

Every set of sequences that is l.i. mod AA^{*} can be extended to a maximal linearly independent set mod AA^{*}.∎

It should be clear that ABωA\subseteq B\subseteq\omega and AA and BB are infinite, then a set that is l.i. mod BB^{*} is also l.i. mod AA^{*}. Then by using recursion, this easily implies the following corollary:

Corollary 3.3.

There exists a family (T:T𝒯)(\mathcal{E}_{T}:T\in\mathcal{T}) such that:

  1. (1)

    For every T𝒯T\in\mathcal{T} the set T\mathcal{E}_{T} is maximal l.i. mod l(T)l(T)^{*}, and

  2. (2)

    For every T𝒯T\in\mathcal{T}, if n|T|n\leq|T| then T|nT\mathcal{E}_{T|n}\subseteq\mathcal{E}_{T}.

We note explicitly that even though T\mathcal{E}_{T} is only demanded to be maximal l.i. mod l(T)l(T)^{*} it will, because of item (2), depend on all of TT, not just on l(T)l(T).

Lemma 3.4.

Let gg be an element of ((κ))ω({\Z}^{(\kappa)})^{\omega} and let ((κ))ω\mathcal{E}\subseteq(\Z^{(\kappa)})^{\omega} be maximal l.i. mod BB^{*}. Then there exist an infinite subset AA of BB, a finite subset EE of \mathcal{E}, a finite subset DD of κ\kappa, and sets {rf:fE}\{r_{f}:f\in E\} and {sν:νD}\{s_{\nu}:\nu\in D\} of rational numbers such that

g|A=fErff|A+νDsνχν|A.g|_{A}=\sum_{f\in E}r_{f}\cdot f|_{A}+\sum_{\nu\in D}s_{\nu}\cdot\chi_{\vec{\nu}}|_{A}.
Proof.

If gg\in\mathcal{E} or g=χνg=\chi_{\vec{\nu}} for some ν<κ\nu<\kappa, then we are done. Otherwise, by the maximality of \mathcal{E}, there exists a free ultrafilter pp with BpB\in p such that the set

{[g]p}{[h]p:h}{[χξ]p:ξ<κ}\{[g]_{p}\}\cup\{[h]_{p}:\,h\in\mathcal{E}\}\cup\{[\chi_{\vec{\xi}}]_{p}:\xi<\kappa\}

is not linearly independent.

This means that we can find finite subsets EE and DD of \mathcal{E} and κ\kappa respectively and finite sets {rf:fE}\{r_{f}:f\in E\} and {sν:νD}\{s_{\nu}:\nu\in D\} of rational numbers such that

[g]p=fErf[f]p+νDsν[χν]p.[g]_{p}=\sum_{f\in E}r_{f}\cdot[f]_{p}+\sum_{\nu\in D}s_{\nu}\cdot[\chi_{\vec{\nu}}]_{p}.

Now choose ApA\in p with ABA\subseteq B that witnesses this equality. ∎

Corollary 3.5.

If ((κ))ω\mathcal{E}\subseteq(\Z^{(\kappa)})^{\omega} is maximal l.i. mod BB^{*}, then ||=κ|\mathcal{E}|=\kappa.

Proof.

First notice that |||((κ))ω|=κω=κ|\mathcal{E}|\leq|(\Z^{(\kappa)})^{\omega}|=\kappa^{\omega}=\kappa. Assume ||<κ|\mathcal{E}|<\kappa. Then the set C={suppf(n):nω,f}C=\bigcup\{\operatorname{supp}f(n):n\in\omega,f\in\mathcal{E}\} has cardinality less than κ\kappa.

Take some injective sequence ξn:nω\langle\xi_{n}:n\in\omega\rangle in κC\kappa\setminus C and define g:ω(κ)g:\omega\to\Z^{(\kappa)} by g(n)=χξng(n)=\chi_{\xi_{n}} for all nn. Clearly then {suppg(n):nω}\bigcup\{\operatorname{supp}g(n):n\in\omega\} is disjoint from CC, all values of gg are non-zero and the values have disjoint supports.

Apply Lemma 3.4 to obtain sets AA, EE, DD, {rf:fE}\{r_{f}:f\in E\}, and {sν:νD}\{s_{\nu}:\nu\in D\} such that

() g|A=fErff|A+νDsνχν|A.g|_{A}=\sum_{f\in E}r_{f}\cdot f|_{A}+\sum_{\nu\in D}s_{\nu}\cdot\chi_{\vec{\nu}}|_{A}.

Since AA is infinite and DD is finite, there is a kAk\in A such that ξkD\xi_{k}\notin D. Now f(k)(ξk)=0f(k)(\xi_{k})=0 when fEf\in E because ξkC\xi_{k}\notin C, and χν(k)(ξk)=0\chi_{\vec{\nu}}(k)(\xi_{k})=0 when νD\nu\in D because ξkD\xi_{k}\notin D, and also g(k)(ξk)=1g(k)(\xi_{k})=1, which contradicts ()(*). ∎

Henceforth we fix a family (T:T𝒯)(\mathcal{E}_{T}:T\in\mathcal{T}) as in Corollary 3.3 and enumerate each T\mathcal{E}_{T} faithfully as T={fξT:κξ<κ+κ}\mathcal{E}_{T}=\{f^{T}_{\xi}:\kappa\leq\xi<\kappa+\kappa\}.

Definition 3.6.

For each T𝒯T\in\mathcal{T} and nωn\in\omega, we denote by GT,nG_{T,n} the intersection of Ult(,pT,n)\operatorname{Ult}(\Z,p_{T,n}) and the free Abelian group generated by {1n![fξT]pT,n:κξ<κ+κ}{1n![χξ]pT,n:ξ<κ}\{\frac{1}{n!}[f^{T}_{\xi}]_{{p}_{T,n}}:\kappa\leq\xi<\kappa+\kappa\}\cup\{\frac{1}{n!}[\chi_{\vec{\xi}}]_{{p}_{T,n}}:\xi<\kappa\}.

For the next lemma, we are going to use the following proposition:

Proposition 3.7.

If GG is an abelian group and HH is a subgroup of GG such that G/HG/H is an infinite cyclic group, then there exists aGa\in G such that G=HaG=H\oplus\langle a\rangle.

A proof may be found in [9, 14.4]. This is not the statement of the theorem but it is exactly what is proved by the author.

The main idea of the proof of the following lemma is to mimic the well known proof of the fact that every subgroup of a free abelian group is free.

Lemma 3.8.

The group GT,nG_{T,n} has a basis of the form {[χξ]pT,n:ξ<κ}˙{[f]pT,n:fT,n}\{[\chi_{\vec{\xi}}]_{p_{T,n}}:\xi<\kappa\}\mathbin{\dot{\cup}}\{[f]_{p_{T,n}}:f\in\mathcal{F}_{T,n}\} for some subset T,n\mathcal{F}_{T,n} of ((κ))ω({\Z}^{(\kappa)})^{\omega}.

Proof.

Let HμH_{\mu} the the group generated by {1n![χξ]pT,n:ξ<μ}\{\frac{1}{n!}[\chi_{\vec{\xi}}]_{p_{T,n}}:\xi<\mu\} if μκ\mu\leq\kappa and by the union of {1n![χξ]pT,n:ξ<κ}\{\frac{1}{n!}[\chi_{\vec{\xi}}]_{p_{T,n}}:\xi<\kappa\} and {1n![fξT]pT,n:κξ<μ}\{\frac{1}{n!}[f^{T}_{\xi}]_{p_{T,n}}:\kappa\leq\xi<\mu\} when κ<μκ+κ\kappa<\mu\leq\kappa+\kappa.

Let Gμ=HμUlt(,pT,n)G_{\mu}=H_{\mu}\cap\operatorname{Ult}(\Z,p_{T,n}) for all μ\mu.

For every μ<κ+κ\mu<\kappa+\kappa we shall find hμh_{\mu} so that Gμ+1=Gμ<{[hμ]pT,n}>G_{\mu+1}=G_{\mu}\oplus\bigl{<}\{[h_{\mu}]_{p_{T,n}}\}\bigr{>}, as follows.

For μ<κ\mu<\kappa the group GμG_{\mu} is generated by {[χξ]pT,n:ξ<μ}\{[\chi_{\vec{\xi}}]_{p_{T,n}}:\xi<\mu\}, so Gμ+1=Gμ{[χμ]}G_{\mu+1}=G_{\mu}\oplus\langle\{[\chi_{\vec{\mu}}]\}\rangle and we have hμ=χμh_{\mu}=\chi_{\vec{\mu}}.

For μκ\mu\geq\kappa observe that Gμ+1Hμ=GμG_{\mu+1}\cap H_{\mu}=G_{\mu}, so:

Gμ+1Gμ=Gμ+1Gμ+1HμGμ+1+HμHμHμ+1Hμ.\frac{G_{\mu+1}}{G_{\mu}}=\frac{G_{\mu+1}}{{G_{\mu+1}}\cap H_{\mu}}\approx\frac{G_{\mu+1}+H_{\mu}}{H_{\mu}}\leq\frac{H_{\mu+1}}{H_{\mu}}.

The group Hμ+1Hμ\frac{H_{\mu+1}}{H_{\mu}} is cyclic infinite, so either Gμ+1Gμ\frac{G_{\mu+1}}{G_{\mu}} is infinite and cyclic or Gμ+1=GμG_{\mu+1}=G_{\mu}. By Proposition 3.7 there exists aμGμ+1a_{\mu}\in G_{\mu+1} such that Gμ+1=Gμ{aμ}G_{\mu+1}=G_{\mu}\oplus\langle\{a_{\mu}\}\rangle (and aμ=0a_{\mu}=0 in case Gμ+1=GμG_{\mu+1}=G_{\mu}). Take hμh_{\mu} such that [hμ]pT,n=aμ[h_{\mu}]_{p_{T,n}}=a_{\mu}.

For every μ<κ+κ\mu<\kappa+\kappa, it follows that Gμ+1=Gμ<{[hμ]pT,n}>G_{\mu+1}=G_{\mu}\oplus\bigl{<}\{[h_{\mu}]_{p_{T,n}}\}\bigr{>}. Since GT,n=μ<κ+κGμG_{T,n}=\bigcup_{\mu<\kappa+\kappa}G_{\mu}, it follows that GT,n=μ<κ+κ<{[hμ]pT,n}>G_{T,n}=\bigoplus_{\mu<\kappa+\kappa}\bigl{<}\{[h_{\mu}]_{p_{T,n}}\}\bigr{>}.

The set T,n={hμ:κμ<κ+κ,[hμ]pT,n0}\mathcal{F}_{T,n}=\{h_{\mu}:\kappa\leq\mu<\kappa+\kappa,[h_{\mu}]_{p_{T,n}}\neq 0\} is as required. ∎

For the rest of this article we fix such a set T,n\mathcal{F}_{T,n} as above for each pair (T,n)(T,n) in 𝒯×ω\mathcal{T}\times\omega.

The next lemma makes good on the promise from the beginning of this section as it shows how to make our topology countably compact.

Lemma 3.9.

Assume that for every pair (T,n)(T,n) in 𝒯×ω\mathcal{T}\times\omega every sequence ff in T,n\mathcal{F}_{T,n} has a pT,np_{T,n}-limit in (κ){\Z}^{(\kappa)}. Then every finite power of (κ){\Z}^{(\kappa)} is countably compact.

Proof.

A sequence in some finite power of (κ){\Z}^{(\kappa)} is represented by finitely members of ((κ))ω({\Z}^{(\kappa)})^{\omega}, say g0,,gmg_{0},\dots,g_{m}. We show that there is one ultrafilter pp such that plimgip\mathchar 45\relax\!\lim g_{i} exists for all ii, namely pT,np_{T,n} for a suitable TT and nn.

Recursively, we define a tower T=(A0,,Am)T=(A_{0},\dots,A_{m}) and for imi\leq m finite subsets EiE_{i} and DiD_{i} of T|i\mathcal{E}_{T|_{i}} and κ\kappa respectively together with finite sets (rfi:fEi)(r_{f}^{i}:f\in E_{i}) and (sνi:νDi)(s_{\nu}^{i}:\nu\in D_{i}) of rational numbers such that

() gi|Ai=fEirfif|Ai+νDisνiχν|Aig_{i}|_{A_{i}}=\sum_{f\in E_{i}}r_{f}^{i}\cdot f|_{A_{i}}+\sum_{\nu\in D_{i}}s_{\nu}^{i}\cdot\chi_{\vec{\nu}}|_{A_{i}}

For i=0i=0, use Lemma 3.4 applied to \mathcal{E}_{\emptyset} to obtain A0A_{0}, E0E_{0}, D0D_{0}, (rf0:fE0)(r^{0}_{f}:f\in E_{0}) and (sν0:νD0)(s^{0}_{\nu}:\nu\in D_{0}) such that ()(*) holds with i=0i=0.

To go from ii to i+1i+1 apply Lemma 3.4 to (A0,,Ai)\mathcal{E}_{(A_{0},\dots,A_{i})} to obtain Ai+1A_{i+1}, Ei+1E_{i+1}, Di+1D_{i+1}, (rfi+1:fEi+1)(r^{i+1}_{f}:f\in E_{i+1}), and (sνi+1:νDi+1)(s^{i+1}_{\nu}:\nu\in D_{i+1}) so that ()(*) holds for i+1i+1.

Let A=AmA=A_{m} and let nn be sufficiently large so that n!rfin!r_{f}^{i} and n!sνin!s_{\nu}^{i} are integers, for all imi\leq m, fEif\in E_{i}, and νDi\nu\in D_{i}. Then gi|A=fEin!rfi(1n!f)|A+νDin!sνi(1n!χν)|Ag_{i}|_{A}=\sum_{f\in E_{i}}n!\cdot r_{f}^{i}\cdot(\frac{1}{n!}\cdot f)|_{A}+\sum_{\nu\in D_{i}}n!\cdot s_{\nu}^{i}\cdot(\frac{1}{n!}\cdot\chi_{\vec{\nu}})|_{A} for all ii.

As l(T)=ApT,nl(T)=A\in p_{T,n} and for each EiE_{i} is a subset of T\mathcal{E}_{T}, it follows that [gi]pT,nGT,n[g_{i}]_{p_{T,n}}\in G_{T,n}. Therefore, each [gi]pT,n[g_{i}]_{p_{T,n}} is an integer combination of {[f]pT,n:fT,n}{[χξ]pT,n:ξ<κ}\{[f]_{p_{T,n}}:f\in\mathcal{F}_{T,n}\}\cup\{[\chi_{\xi}]_{p_{T,n}}:\xi<\kappa\}. Then, by hypothesis, it follows that each gig_{i} has a pT,np_{T,n}-limit. This completes the proof. ∎

4. Constructing homomorphisms

Through this section, we let G=(κ)G={\Z}^{(\kappa)} and we let {hξ:ωξ<κ}\{h_{\xi}:\omega\leq\xi<\kappa\} be an enumeration of GωG^{\omega} such that supphξ(n)ξ\operatorname{supp}h_{\xi}(n)\subseteq\xi whenever nωn\in\omega and ωξ<κ\omega\leq\xi<\kappa, and so that each element of GωG^{\omega} appears at least 𝔠\mathfrak{c} many times.

Lemma 4.1.

There exists a family (JT,n:T𝒯,nω)(J_{T,n}:T\in\mathcal{T},n\in\omega) of pairwise disjoint subsets of κ\kappa such that {hξ:ξJT,n}=T,n\{h_{\xi}:\xi\in J_{T,n}\}=\mathcal{F}_{T,n}.

Proof.

For each fGωf\in G^{\omega} there is an injective map ϕf:𝒯×ω{ξκ:f=hξ}\phi_{f}:\mathcal{T}\times\omega\to\{\xi\in\kappa:f=h_{\xi}\}. Let JT,n={ϕf(T,n):fT,n}J_{T,n}=\{\phi_{f}(T,n):f\in\mathcal{F}_{T,n}\} and we are done. ∎

For the rest of this section, we fix a family (JT,n:T𝒯,nω)(J_{T,n}:T\in\mathcal{T},n\in\omega) as above.

The following lemma is the key to the main result.

Lemma 4.2.

Assume we have a non-zero element dd of GG, an injective sequence rr in GG, and a countably infinite subset DD of κ\kappa such that

  1. (1)

    ωsuppdnωsuppr(n)D\omega\cup\operatorname{supp}d\cup\bigcup_{n\in\omega}\operatorname{supp}r(n)\subseteq D,

  2. (2)

    DJT,nD\cap J_{T,n}\neq\emptyset for infinitely many (T,n)(T,n)’s and,

  3. (3)

    supphξ(n)D\operatorname{supp}h_{\xi}(n)\subseteq D for all nωn\in\omega and ξDω\xi\in D\setminus\omega

Then there exists a homomorphism ϕ:(D)𝕋\phi:\Z^{(D)}\to\T such that:

  1. (1)

    ϕ(d)0\phi(d)\neq 0

  2. (2)

    pT,nlimkϕ(hξ(k))=ϕ(χξ)p_{T,n}\mathchar 45\relax\!\lim_{k}\phi(h_{\xi}(k))=\phi(\chi_{\xi}), whenever T𝒯T\in\mathcal{T}, nωn\in\omega, and ξDJT,n\xi\in D\cap J_{T,n}.

  3. (3)

    ϕr\phi\circ r does not converge.

Before proving this lemma, we show how to use it to prove the main result. First, we use it to prove another lemma:

Lemma 4.3.

Assume dd is a non-zero element of GG and rr is an injective sequence in GG. Then there exists a homomorphism ϕ:(κ)𝕋\phi:\Z^{(\kappa)}\to\T such that

  1. (1)

    ϕ(d)0\phi(d)\neq 0

  2. (2)

    pT,nlimkϕ(hξ(k))=ϕ(χξ)p_{T,n}\mathchar 45\relax\!\lim_{k}\phi(h_{\xi}(k))=\phi(\chi_{\xi}), whenever T𝒯T\in\mathcal{T}, nωn\in\omega and ξJT,n\xi\in J_{T,n}.

  3. (3)

    ϕr\phi\circ r does not converge.

Proof.

Using a closing-off argument construct a countable subset DD of κ\kappa that intersects infinitely many sets JT,nJ_{T,n}, and that contains ω\omega, suppd\operatorname{supp}d, suppr(n)\operatorname{supp}r(n) for all nn as well as supphξ(n)\operatorname{supp}h_{\xi}(n) whenever ξDω\xi\in D\setminus\omega and nωn\in\omega.

By the previous Lemma, there exists a homomorphism ϕ0:(D)𝕋\phi_{0}:\Z^{(D)}\to\T such that ϕ0(d)0\phi_{0}(d)\neq 0, ϕr\phi\circ r does not converge, and pT,nlimkϕ0(hξ(k))=ϕ0(χξ)p_{T,n}\mathchar 45\relax\!\lim_{k}\phi_{0}(h_{\xi}(k))=\phi_{0}(\chi_{\xi}) whenever T𝒯T\in\mathcal{T}, nωn\in\omega and ξDJT,n\xi\in D\cap J_{T,n}.

We let αδ:δ<κ\langle\alpha_{\delta}:\delta<\kappa\rangle be the monotone enumeration of κD\kappa\setminus D. For γκ\gamma\leq\kappa, let Dγ=D{αδ:δ<γ}D_{\gamma}=D\cup\{\alpha_{\delta}:\delta<\gamma\}. So D0=DD_{0}=D and Dκ=κD_{\kappa}=\kappa.

Recursively, we construct, for γκ\gamma\leq\kappa, an increasing sequence of homomorphisms ϕγ:(Dγ)𝕋\phi_{\gamma}:\Z^{(D_{\gamma})}\to\T such that pT,nlimkϕγ(hξ(k))=ϕγ(χξ)p_{T,n}\mathchar 45\relax\!\lim_{k}\phi_{\gamma}(h_{\xi}(k))=\phi_{\gamma}(\chi_{\xi}) whenever T𝒯T\in\mathcal{T}, nωn\in\omega and ξDγJT,n\xi\in D_{\gamma}\cap J_{T,n}. Our homomorphism ϕ\phi will be ϕκ\phi_{\kappa}. The basis step 0 is already done, and for limit steps, we just unite all previous homomorphisms.

To define ϕγ+1\phi_{\gamma+1} given ϕγ\phi_{\gamma} it suffices to specify the value ϕγ+1(χαγ)\phi_{\gamma+1}(\chi_{\alpha_{\gamma}}).

If αγJT,n\alpha_{\gamma}\in J_{T,n} for some T𝒯T\in\mathcal{T} and nωn\in\omega then we put ϕγ+1(χαγ)=pT,nlimnϕγ(hγ(n))\phi_{\gamma+1}(\chi_{\alpha_{\gamma}})=p_{T,n}\mathchar 45\relax\!\lim_{n}\phi_{\gamma}(h_{\gamma}(n)). This is well defined because supphγ(n)γDγ\operatorname{supp}h_{\gamma}(n)\subseteq\gamma\subseteq D_{\gamma} for all nn and because 𝕋\T is compact. In the other case let ϕγ+1(χαγ)=0\phi_{\gamma+1}(\chi_{\alpha_{\gamma}})=0. ∎

We can now prove our main result.

Theorem 4.4.

Assume the existence of pairwise incompatible 𝔠\mathfrak{c} selective ultrafilters and that κ\kappa is an infinite cardinal such that κω\kappa^{\omega}. Then the free abelian group of cardinality κ\kappa has a Hausdorff group topology without nontrivial converging sequences such that all of its finite powers are countably compact.

Proof.

Following the notation of the rest of the article, given dG{0}d\in G\setminus\{0\} and an injective sequence rr in GG, Lemma 4.3 provides a homomorphism ϕd,r:G𝕋\phi_{d,r}:G\to\T such that ϕd(d)0\phi_{d}(d)\neq 0, such that ϕd,rr\phi_{d,r}\circ r does not converge, and such that pT,nlimkϕd,r(hξ(k))=ϕd,r(χξ)p_{T,n}\mathchar 45\relax\!\lim_{k}\phi_{d,r}(h_{\xi}(k))=\phi_{d,r}(\chi_{\xi}) whenever T𝒯T\in\mathcal{T}, nωn\in\omega and ξJT,n\xi\in J_{T,n}. We give GG the initial topology generated by the collection of homomorphisms {ϕd,r:dG{0}\{\phi_{d,r}:d\in G\setminus\{0\}, rGωr\in G^{\omega} is injective}\} thus obtained and the natural topology of 𝕋\T.

Since the initial topology generated by any collection of group homomorphisms is a group topology we do indeed obtain a group topology. Since 𝕋\T is Hausdorff and for every d0d\neq 0 there are many ϕd,r\phi_{d,r} with ϕd,r(d)0\phi_{d,r}(d)\neq 0 it follows at once that our topology is Hausdorff.

To see that every finite power of GG is countably compact we use Lemma 3.9.

Given T𝒯T\in\mathcal{T}, nωn\in\omega and fT,nf\in\mathcal{F}_{T,n}, there exist ξJT,n\xi\in J_{T,n} such that hξ=fh_{\xi}=f. For every dG{0}d\in G\setminus\{0\} and injective rGωr\in G^{\omega}, we have pT,nlimnϕd,r(hξ(n))=ϕd,r(χξ)p_{T,n}\mathchar 45\relax\!\lim_{n}\phi_{d,r}(h_{\xi}(n))=\phi_{d,r}(\chi_{\xi}). So pT,nlimf(n)=χξp_{T,n}\mathchar 45\relax\!\lim f(n)=\chi_{\xi} and we are done.

Since for a given injective sequence rr and any dGωd\in G^{\omega} the sequence ϕd,rr\phi_{d,r}\circ r does not converge and ϕd,r\phi_{d,r} is continuous, it follows that rr does not converge. So GG has no nontrivial convergent sequences. ∎

Towards the proof of Lemma 4.2 we formulate a definition and a (very) technical lemma.

Definition 4.5.

Let ϵ>0\epsilon>0. An ϵ\epsilon-arc function is a function ψ\psi from κ\kappa into the set of open arcs of 𝕋\T (including 𝕋\T itself) such that for all α\alpha either ψ(α)=𝕋\psi(\alpha)=\T or the length of ψ(α)\psi(\alpha) is equal to ϵ\epsilon, and the set {ακ:ψ(α)𝕋}\{\alpha\in\kappa:\psi(\alpha)\neq\T\} is finite. We will call this finite set the support of ψ\psi and denote it by suppψ\operatorname{supp}\psi.

Given two arc functions ψ\psi and ϱ\varrho we write ψϱ\psi\leq\varrho if ψ(α)¯ϱ(α)\overline{\psi(\alpha)}\subseteq\varrho(\alpha) or ψ(α)=ϱ(α)\psi(\alpha)=\varrho(\alpha) for each ακ\alpha\in\kappa.

We shall obtain our homomorphisms using limits of such arc functions. The following lemmas are instrumental in its construction.

The following result follows from an argument implicit in the construction of [2], but it may be difficult to extract it from that paper. We postpone its rather technical proof to the next section.

Lemma 4.6.

Let pp be a selective ultrafilter and \mathcal{F} a finite subset of GωG^{\omega} such that the set {[f]p:f}{[χα]p:α<κ}\{[f]_{p}:f\in\mathcal{F}\}\cup\{[\chi_{\vec{\alpha}}]_{p}:\alpha<\kappa\} is linearly independent.

Then for a given ϵ>0\epsilon>0 and a finite subset EE of κ\kappa there exist ApA\in p and a sequence (δn:nA)(\delta_{n}:n\in A) of positive real numbers such that

  • ()(\star)

    whenever {Uf:f}\{U_{f}:f\in\mathcal{F}\} is a family of arcs of length ϵ\epsilon and ϱ\varrho is an arc function of length at least ϵ\epsilon with suppϱE\operatorname{supp}\varrho\subseteq E there exist for each nAn\in A a δn\delta_{n}-arc function ψnϱ\psi_{n}\leq\varrho such that suppψn=fsuppf(n)E\operatorname{supp}\psi_{n}=\bigcup_{f\in\mathcal{F}}\operatorname{supp}f(n)\cup E, and μsuppff(n)(μ)ψn(μ)Uf\sum_{\mu\in\operatorname{supp}f}f(n)(\mu)\cdot\psi_{n}(\mu)\subseteq U_{f} for each ff\in\mathcal{F}.

Now we proceed to prove Lemma 4.2. We will use the following lemma:

Lemma 4.7.

Let (k:kω)(\mathcal{F}^{k}:k\in\omega) be a sequence of countable subsets of GωG^{\omega} and let (pk:kω)(p_{k}:k\in\omega) is a sequence of pairwise incomparable selective ultrafilters such that for each kωk\in\omega the set {[f]pk:fk}˙{[χξ]pk:ξκ}\{[f]_{p_{k}}:f\in\mathcal{F}^{k}\}\mathbin{\dot{\cup}}\{[\chi_{\vec{\xi}}]_{p_{k}}:\xi\in\kappa\} is linearly independent and [f]pk[g]pk[f]_{p_{k}}\neq[g]_{p_{k}} whenever fgf\neq g in k\mathcal{F}^{k}. Furthermore let for every fkkf\in\bigcup_{k}\mathcal{F}^{k} an ordinal ξf\xi_{f} in κ\kappa be given. In addition let dd and dd^{\prime} be non-zero in GG and with disjoint supports. Finally, let DD be a countable subset of κ\kappa that contains ωsuppdsuppd\omega\cup\operatorname{supp}d\cup\operatorname{supp}d^{\prime} and nsuppf(n)\bigcup_{n}\operatorname{supp}f(n) for every fkkf\in\bigcup_{k}\mathcal{F}^{k}.

Then there exists a homomorphism ϕ:(D)𝕋\phi:\Z^{(D)}\to\T such that ϕ(d)0\phi(d)\neq 0, ϕ(d)0\phi(d^{\prime})\neq 0 and pklimnϕ(f(n))=ϕ(χξf)p_{k}\mathchar 45\relax\!\lim_{n}\phi(f(n))=\phi(\chi_{\xi_{f}}), whenever kωk\in\omega and fkf\in\mathcal{F}^{k}.

Proof.

Write DD as the union of an increasing sequence (Dn:nω)(D_{n}:n\in\omega) of finite nonempty subsets, and take a similar sequence (nk:nω)(\mathcal{F}^{k}_{n}:n\in\omega) for each k\mathcal{F}^{k}.

Take a sufficiently small positive number ϵ0\epsilon_{0} and an ϵ0\epsilon_{0}-arc function ϱ\varrho_{*} such that suppdsuppdsuppϱ\operatorname{supp}d\cup\operatorname{supp}d^{\prime}\subseteq\operatorname{supp}\varrho_{*} and 0μsuppdd(μ)ϱ(μ)¯μsuppdd(μ)ϱ(μ)¯0\notin\overline{\sum_{\mu\in\operatorname{supp}d}d(\mu)\varrho_{*}(\mu)}\cup\overline{\sum_{\mu\in\operatorname{supp}d^{\prime}}d^{\prime}(\mu)\varrho_{*}(\mu)}.

Let E0=suppϱD0E_{0}=\operatorname{supp}\varrho_{*}\cup D_{0} and B0k=ωB^{k}_{0}=\omega for each kωk\in\omega.

We will define, by recursion, for mωm\in\omega: finite sequences (Bmk:0km)(B^{k}_{m}:0\leq k\leq m), finite sets EmκE_{m}\subseteq\kappa, and real numbers ϵm>0\epsilon_{m}>0 satisfying:

  1. (1)

    For all kk and mm in ω\omega we have BmkpkB^{k}_{m}\in p_{k},

  2. (2)

    For each m1m\geq 1 and kmk\leq m, we have a sequence (δm,nk:nω)(\delta^{k}_{m,n}:n\in\omega) of positive real numbers such that: if (Uf:fmk)(U_{f}:f\in\mathcal{F}^{k}_{m}) is a family of arcs of length ϵm1\epsilon_{m-1} and ϱ\varrho is an arc function of length ϵm1\epsilon_{m-1} and suppϱEm1\operatorname{supp}\varrho\subseteq E_{m-1} then for each nωn\in\omega there exists a δm,nk\delta^{k}_{m,n}-arc function ψ\psi with ψϱ\psi\leq\varrho, and suppψ=fmksuppf(n)Em1\operatorname{supp}\psi=\bigcup_{f\in\mathcal{F}^{k}_{m}}\operatorname{supp}f(n)\cup E_{m-1}, and μsuppff(n)(μ)ψ(μ)Uf\sum_{\mu\in\operatorname{supp}f}f(n)(\mu)\psi(\mu)\subseteq U_{f} for each fmkf\in\mathcal{F}^{k}_{m}.

  3. (3)

    For all kk and mm we have Bm+1kBmkB^{k}_{m+1}\subseteq B^{k}_{m}.

  4. (4)

    ϵm+1=12min({δl,nk:klm+1\epsilon_{m+1}=\frac{1}{2}\min(\{\delta^{k}_{l,n}:k\leq l\leq m+1 and n(m+2)Blk}{ϵm})n\in(m+2)\cap B^{k}_{l}\}\cup\{\epsilon_{m}\}).

Suppose we have defined BlkB^{k}_{l} for all kk as well as ElE_{l} and ϵl\epsilon_{l} for all lml\leq m. As will be clear from the step below the set BmkB^{k}_{m} is only non-trivial whenever kmk\leq m. Therefore we let Bm+1k=Bmk=ωB^{k}_{m+1}=B^{k}_{m}=\omega for k>m+1k>m+1 and we concentrate on the case km+1k\leq m+1.

Let km+1k\leq m+1. By Lemma 4.6, there exist Bm+1kpkB^{k}_{m+1}\in p_{k} and (δm+1,nk:nω)(\delta^{k}_{m+1,n}:n\in\omega) that satisfy (2) for m+1m+1. Without loss of generality we can assume that Bm+1kBmkB^{k}_{m+1}\subseteq B^{k}_{m}.

Condition (4) now specifies ϵm+1\epsilon_{m+1}.

Setting Em+1=Em{suppf(k):km,fkm+1m+1k}Dm+1E_{m+1}=E_{m}\cup\bigcup\{\operatorname{supp}f(k):k\leq m,f\in\bigcup_{k\leq m+1}\mathcal{F}_{m+1}^{k}\}\cup D_{m+1} completes the recursion.

For each kωk\in\omega, apply the selectivity of pkp_{k}, to choose an increasing sequence (ak,i:iω)(a_{k,i}:i\in\omega) with {ak,i:iω}pk\{a_{k,i}:i\in\omega\}\in p_{k} and such that ak,iBika_{k,i}\in B^{k}_{i} and ak,i>ia_{k,i}>i for all ii.

Next apply Lemma 2.6 and let (Ik:kω)(I_{k}:k\in\omega) be a sequence of pairwise disjoint subsets of ω\omega such that {ak,i:iIk}pk\{a_{k,i}:i\in I_{k}\}\in p_{k} and the family of intervals {[i,ak,i]:kω,iIk}\{[i,a_{k,i}]:k\in\omega,i\in I_{k}\} is pairwise disjoint. Without loss of generality we can assume that k<minIkk<\min I_{k}.

Enumerate kωIk\bigcup_{k\in\omega}I_{k} in increasing order as (it:tω)(i_{t}:t\in\omega). For each tωt\in\omega, let ktk_{t} be such that itIkti_{t}\in I_{k_{t}}. Thus, for each tt we have itIkti_{t}\in I_{k_{t}}, and hence itminIkt>kti_{t}\geq\min I_{k_{t}}>k_{t} and akt,it>ita_{k_{t},i_{t}}>i_{t}.

By recursion we define a sequence of arc functions, (ϱit:tω)(\varrho_{i_{t}}:t\in\omega), such that ϱi0ϱ\varrho_{i_{0}}\leq\varrho_{*} and ϱit+1ϱit\varrho_{i_{t+1}}\leq\varrho_{i_{t}}.

We start with t=0t=0. Then we have k0<i0<ak0,i0k_{0}<i_{0}<a_{k_{0},i_{0}}, and ak0,i0Bi0k0a_{k_{0},i_{0}}\in B^{k_{0}}_{i_{0}}, and ϵi01ϵ0\epsilon_{i_{0}-1}\leq\epsilon_{0}.

Since ϱ\varrho_{*} has length at least ϵi01\epsilon_{i_{0}-1}, there exists an arc function ϱi0\varrho_{i_{0}} of length δi0,ak0,i0k0\delta^{k_{0}}_{i_{0},a_{k_{0},i_{0}}} such that μsuppff(ak0,i0)(μ)ϱi0(μ)ϱ(ξf)\sum_{\mu\in\operatorname{supp}f}f(a_{k_{0},i_{0}})(\mu)\varrho_{i_{0}}(\mu)\subseteq\varrho_{*}(\xi_{f}), for each fi0k0f\in\mathcal{F}_{i_{0}}^{k_{0}}. We have by the definition that δi0,ak0,i0k0>ϵi11\delta^{k_{0}}_{i_{0},a_{k_{0},i_{0}}}>\epsilon_{i_{1}-1}.

Suppose t>0t>0 and that ϱit1\varrho_{i_{t-1}} has been defined with length at least ϵit1\epsilon_{i_{t-1}}.

Apply item (2) to the arc function ϱit1\varrho_{i_{t-1}}, the finite set =itkt\mathcal{F}=\mathcal{F}^{k_{t}}_{i_{t}}, the number ϵit1\epsilon_{i_{t-1}}, the finite set Eit1E_{i_{t-1}}, the arcs Uf=ϱit1(ξf)U_{f}=\varrho_{i_{t-1}}(\xi_{f}) for fitktf\in\mathcal{F}_{i_{t}}^{k_{t}}, and n=akt,itBitktn=a_{k_{t},i_{t}}\in B^{k_{t}}_{i_{t}} to obtain an arc function ϱitϱit1\varrho_{i_{t}}\leq\varrho_{i_{t-1}} such that μsuppff(akt,it)(μ)ϱit(μ)ϱit1(ξf)\sum_{\mu\in\operatorname{supp}f}f(a_{k_{t},i_{t}})(\mu)\varrho_{i_{t}}(\mu)\subseteq\varrho_{i_{t-1}}(\xi_{f}) for all fitktf\in\mathcal{F}_{i_{t}}^{k_{t}}, and ϱit\varrho_{i_{t}} has length δit,akt,itkt\delta^{k_{t}}_{i_{t},a_{k_{t},i_{t}}}.

Because kt<it<akt,itit+11k_{t}<i_{t}<a_{k_{t},i_{t}}\leq i_{t+1}-1 and akt,itBitkta_{k_{t},i_{t}}\in B^{k_{t}}_{i_{t}} we get δit,akt,itkt>ϵit+11\delta^{k_{t}}_{i_{t},a_{k_{t},i_{t}}}>\epsilon_{i_{t+1}-1}.

If ξDit\xi\in D_{i_{t}} then ξsuppϱit\xi\in\operatorname{supp}\varrho_{i_{t}} and the length of ϱit(ξ)\varrho_{i_{t}}(\xi) is not greater than ϵit1\epsilon_{i_{t}-1} which in turn is not larger than 12it112t\frac{1}{2^{i_{t-1}}}\leq\frac{1}{2^{t}}.

It follows that for all ξD\xi\in D the intersection tωϱit(ξ)\bigcap_{t\in\omega}\varrho_{i_{t}}(\xi) consists of a unique element; we define ϕ(χξ)\phi(\chi_{\xi}) to be that element and extend ϕ\phi to a group homomorphism.

By construction ϕ(f(akt,it))\phi(f(a_{k_{t},{i_{t}}})) is in μsuppff(akt,it)(μ)ϱit(μ)\sum_{\mu\in\operatorname{supp}f}f(a_{k_{t},i_{t}})(\mu)\varrho_{i_{t}}(\mu) which is a subset of ϱit1(ξf)\varrho_{i_{t-1}}(\xi_{f}) whenever fitktf\in\mathcal{F}_{i_{t}}^{k_{t}}. Therefore, the sequence (ϕ(f(ak,i)))iIk(\phi(f(a_{k,i})))_{i\in I_{k}} converges to ϕ(χξf)\phi(\chi_{\xi_{f}}), for each kωk\in\omega and fkf\in\mathcal{F}^{k}.

Furthermore ϕ(d)μsuppdd(μ)ϱ(μ)\phi(d)\in\sum_{\mu\in\operatorname{supp}d}d(\mu)\varrho_{*}(\mu), therefore, ϕ(d)0\phi(d)\neq 0; and likewise ϕ(d)0\phi(d^{\prime})\neq 0.

It is clear that this implies the conclusion of Lemma 4.7. ∎

Now we are ready to prove Lemma 4.2.

Proof of Lemma 4.2.

There are only a countably many of pairs (T,n)𝒯×ω(T,n)\in\mathcal{T}\times\omega such that JT,nDJ_{T,n}\cap D\neq\emptyset. We enumerate them faithfully as ((Tm,nm):m2)((T_{m},n_{m}):m\geq 2).

For m2m\geq 2 let m={hξ:ξDJTm,nm}\mathcal{F}^{m}=\{h_{\xi}:\xi\in D\cap J_{T_{m},n_{m}}\} and pm=pTm,nmp_{m}=p_{T_{m},n_{m}}. Let p0p_{0} and p1p_{1} be two ultrafilters that were not listed and let 0=1={r}\mathcal{F}^{0}=\mathcal{F}^{1}=\{r\}. For each m2m\geq 2 and ξJTm,nmD\xi\in J_{T_{m},n_{m}}\cap D, let ξhξ,m=ξ\xi_{h_{\xi},m}=\xi. Let hr,0=χkh_{r,0}=\chi_{k} and hr,1=χkh_{r,1}=\chi_{k^{\prime}} where k,kωk,k^{\prime}\in\omega are not in suppd\operatorname{supp}d. Then, by applying Lemma 4.7 with d=χkχkd^{\prime}=\chi_{k}-\chi_{k^{\prime}}, there exist ϕ:(D)𝕋\phi:\Z^{(D)}\to\T satisfying (1) and (2). To see it also satisfies (3), notice that p0limϕrp1limϕrp_{0}\mathchar 45\relax\!\lim\phi\circ r\neq p_{1}\mathchar 45\relax\!\lim\phi\circ r. ∎

5. Proof of Lemma 4.6

In this section we present a proof of Lemma 4.6. We will need the notion of integer stack, which was defined in [26].

The integer stacks are collections of sequences in (𝔠)\Z^{(\mathfrak{c})} that are usually associated to a selective ultrafilter. Given an finite set of sequences \mathcal{F} it is possible to associate it to a integer stack which generates the same \Q vector space as \mathcal{F}. The sequences in the stack have some nice properties that help us to construct well behaved arcs when constructing homomorphisms, and the linear relations between \mathcal{F} and the sequences of the stack helps us to transform these arcs into arcs that work for the functions of \mathcal{F}. Below, we give the definition of integer stack.

Definition 5.1.

An integer stack 𝒮\mathcal{S} on AA consists of

  1. (i)

    an infinite subset AA of ω\omega;

  2. (ii)

    natural numbers ss, tt, and MM; positive integers rir_{i} for 0i<s0\leq i<s and positive integers ri,jr_{i,j} for 0i<s0\leq i<s and 0j<ri0\leq j<r_{i};

  3. (iii)

    functions fi,j,k((𝔠))Af_{i,j,k}\in(\Z^{(\mathfrak{c})})^{A} for 0i<s0\leq i<s, 0j<ri0\leq j<r_{i} and 0k<ri,j0\leq k<r_{i,j} and elements gl((𝔠))Ag_{l}\in(\Z^{(\mathfrak{c})})^{A} for 0l<t0\leq l<t;

  4. (iv)

    sequences ξi𝔠A\xi_{i}\in\mathfrak{c}^{A} for 0i<s0\leq i<s and μl𝔠A\mu_{l}\in\mathfrak{c}^{A} for 0l<t0\leq l<t and

  5. (v)

    real numbers θi,j,k\theta_{i,j,k} for 0i<s0\leq i<s, 0j<ri0\leq j<r_{i} and 0k<ri,j0\leq k<r_{i,j}

These are required to satisfy the following conditions:

  1. (1)

    μl(n)suppgl(n)\mu_{l}(n)\in\operatorname{supp}g_{l}(n) for each nAn\in A;

  2. (2)

    μl(n)suppgl(n)\mu_{l^{*}}(n)\notin\operatorname{supp}g_{l}(n) for each nAn\in A and 0l<l<t0\leq l^{*}<l<t;

  3. (3)

    the elements of {μl(n):0l<t\{\mu_{l}(n):0\leq l<t and nA}n\in A\} are pairwise distinct;

  4. (4)

    |gl(n)|M|g_{l}(n)|\leq M for each nAn\in A and 0l<t0\leq l<t;

  5. (5)

    {θi,j,k:0k<ri,j}\{\theta_{i,j,k}:0\leq k<r_{i,j}\} is a linearly independent subset of \R as a \Q-vector space for each 0i<s0\leq i<s and 0j<ri0\leq j<r_{i};

  6. (6)

    limnAfi,j,k(n)(ξi(n))fi,j,0(n)(ξi(n))=θi,j,k\lim_{n\in A}\frac{f_{i,j,k}(n)(\xi_{i}(n))}{f_{i,j,0}(n)(\xi_{i}(n))}=\theta_{i,j,k} for each 0i<s0\leq i<s, 0j<ri0\leq j<r_{i} and 0k<ri,j0\leq k<r_{i,j};

  7. (7)

    the sequence (|fi,j,k(n)(ξi(n))|:nA)\bigl{(}|f_{i,j,k}(n)(\xi_{i}(n))|:n\in A\bigr{)} diverges monotonically to \infty, for each 0i<s0\leq i<s, 0j<ri0\leq j<r_{i} and 0k<ri,j0\leq k<r_{i,j};

  8. (8)

    |fi,j,k(n)(ξi(n))|>|fi,j,k(n)(ξi(n))||f_{i,j,k}(n)(\xi_{i}(n))|>|f_{i,j,k^{*}}(n)(\xi_{i}(n))| for each nAn\in A, i<si<s, j<rij<r_{i} and 0k<k<ri,j0\leq k<k^{*}<r_{i,j};

  9. (9)

    (|fi,j,k(n)(ξi(n))||fi,j,k(n)(ξi(n))|:nA)\left(\frac{|f_{i,j,k}(n)(\xi_{i}(n))|}{|f_{i,j^{*},k^{*}}(n)(\xi_{i}(n))|}:n\in A\right) converges monotonically to 0 for each 0i<s0\leq i<s, 0j<j<ri0\leq j^{*}<j<r_{i}, 0k<ri,j0\leq k<r_{i,j}, and 0k<ri,j0\leq k^{*}<r_{i,j^{*}}; and

  10. (10)

    {fi,j,k(n)(ξi(n)):nA}[M,M]\{f_{i,j,k}(n)(\xi_{i^{*}}(n)):n\in A\}\subseteq[-M,M] for each 0i<i<s0\leq i^{*}<i<s, 0j<ri0\leq j<r_{i} and 0k<ri,j0\leq k<r_{i,j}.

It is not difficult to show that the sequences of the stack are linearly independent. Moreover, if pp is a free ultrafilter, 𝒮\mathcal{S} is a stack over AA, and ApA\in p, then it is not difficult to see that ([gl]p:l<t)([fi,j,k]p:i<s,j<ri,k<ri,j)({[g_{l}]}_{p}:l<t)\cup({[f_{i,j,k}]}_{p}:i<s,j<r_{i},k<r_{i,j}) is linearly independent in the \Q-vector space (𝔠)/p\Q^{(\mathfrak{c})}/p. We leave the details as an exercise to the reader.

Definition 5.2.

Given an integer stack 𝒮\mathcal{S} and a natural number NN, the NthNth root of 𝒮\mathcal{S}, written 1N𝒮\frac{1}{N}\mathcal{S}, is obtained by keeping all the structure in 𝒮\mathcal{S} with the exception of the functions; these are divided by NN. Thus a function fi,j,k𝒮f_{i,j,k}\in\mathcal{S} is replaced by 1Nfi,j,k\frac{1}{N}f_{i,j,k} in 1N𝒮\frac{1}{N}\mathcal{S} for each 0i<s0\leq i<s, 0j<ri0\leq j<r_{i} and 0k<ri,j0\leq k<r_{i,j} and a function gl𝒮g_{l}\in\mathcal{S} is replaced by 1Ngl\frac{1}{N}g_{l} in 1N𝒮\frac{1}{N}\mathcal{S} for each 0l<t0\leq l<t.

A stack (unqualified) is then defined to be the NthNth root of an integer stack for some positive integer NN.

The lemma below gives the relation between a finite sequence of sequences in (𝔠)\Z^{(\mathfrak{c})} and a stack 𝒮\mathcal{S} that is associated to it. The first part of this lemma is proved in [26]. The second part was stated in [3] with no proof presented there, since it follows directly from statements of several lemmas and constructions from [26]. Since the construction there is long and complicated, we sketch in this paper, for the sake of completeness, a proof for the second part by indicating which statements and proofs from [26] are used, without repeating the arguments.

Lemma 5.3.

Let h0h_{0}, …, hm1h_{m-1} be sequences in (𝔠)\Z^{(\mathfrak{c})} and 𝒰ω\mathcal{U}\in\omega^{*} a selective ultrafilter. Then there exists A𝒰A\in\mathcal{U} and a stack 1N𝒮\frac{1}{N}\mathcal{S} on AA such that: if the elements of the stack have a 𝒰\mathcal{U}-limit in (𝔠)\Z^{(\mathfrak{c})} then hih_{i} has a 𝒰\mathcal{U}-limit in (𝔠)\Z^{(\mathfrak{c})} for each 0i<m0\leq i<m.

We will say in this case that the finite sequence {h0,,hm1}\{h_{0},\ldots,h_{m-1}\} is associated to (1N𝒮,A,𝒰)(\frac{1}{N}\mathcal{S},A,\mathcal{U}).

  1. (#)(\#)

    If {[h0]𝒰,,[hm1]𝒰}\{[h_{0}]_{\mathcal{U}},\ldots,[h_{m-1}]_{\mathcal{U}}\} is a \Q-linearly independent set and the group generated by it does not contain nonzero constant classes, then each restriction hi|Ah_{i}|_{A} is an integer combination of the stack 1N𝒮\frac{1}{N}\mathcal{S} on AA. On the other hand, each element of the integer stack 𝒮\mathcal{S} is an integer combination of {h0,,hm1}\{h_{0},\ldots,h_{m-1}\} restricted to AA.

Proof.

We prove (#)(\#). All numbered references in this proof are to the paper [26].

First, notice that if {[h0]𝒰,,[hm1]𝒰}\{[h_{0}]_{\mathcal{U}},\ldots,[h_{m-1}]_{\mathcal{U}}\} is a \Q-linearly independent set and the group generated by it does not contain nonzero constant classes, then it satisfies the conclusion of Lemma 4.1. Then, following the proof of Lemma 7.1, using the ff’s as the hh’s themselves, we see that the functions h0h_{0}, …, hm1h_{m-1} are integer combinations of the stack 1N.𝒮\frac{1}{N}.\mathcal{S} that was constructed.

It remains to see that the functions of 𝒮\mathcal{S} are integer combinations of the functions hih_{i} restricted to AA. First, notice that in the statement of Lemma 5.4, by x), xi), xii) and xiv) the functions fqi,jf_{q}^{i,j} and gq0g_{q}^{0} are integer combinations of the hih_{i}. This Lemma is used in the proof of Lemma 5.5, where the functions fqi,jf_{q}^{i,j} become the functions fi,j,kf_{i,j,k}, so there are integer combinations of the hih_{i}’s.

Now notice that in Lemma 6.1, by g), c) and finite induction, the functions gjig_{j}^{i} are integer combinations of the hih_{i}, and some of these become the gig_{i}’s in the proof of Lemma 6.2. As in the proof of 7.1 the stack is constructed by applying Lemma 5.5 or Lemma 6.2 or Lemma 5.5 followed by Lemma 6.2 (depending on the case), it follows that the stack constructed consists of functions that are linear combinations of functions the hih_{i}’s (restricted to AA). ∎

Now we define some integers related to Kronecker’s Theorem that will be useful in our proof. The existence of these integers follows from Lemma 4.3. of [26]. These integers were also defined and used in that paper.

Definition 5.4.

If {θ0,,θr1}\{\theta_{0},\ldots,\theta_{r-1}\} is a linearly independent subset of the \Q-vector space \R and ϵ>0\epsilon>0 then L(θ0,,θr1,ϵ)L(\theta_{0},\ldots,\theta_{r-1},\epsilon) denotes a positive integer, LL, such that {(θ0x+,,θr1x+):xI}\{(\theta_{0}x+\Z,\ldots,\theta_{r-1}x+\Z):x\in I\} is ϵ\epsilon-dense in 𝕋r\T^{r} in the usual Euclidean metric product topology, for any interval II of length at least LL.

The last lemma we are going to need is Lemma 8.3 from [26], stated below.

Lemma 5.5.

Let ϵ\epsilon, γ\gamma and ρ\rho be positive reals, NN a positive integer and ψ\psi be an arc function. Let 𝒮\mathcal{S} be an integer stack on A[ω]ωA\in[\omega]^{\omega} and ss, tt, rir_{i}, ri,jr_{i,j}, MM, fi,j,kf_{i,j,k}, glg_{l}, ξi\xi_{i}, μj\mu_{j} and θi,j,k\theta_{i,j,k} be as in Definition 5.1.

Let LL be an integer greater or equal to max{L(θi,j,0,,θi,j,ri,j1,ϵ24):0i<s\max\{L(\theta_{i,j,0},\ldots,\theta_{i,j,r_{i,j}-1},\frac{\epsilon}{24}):0\leq i<s and 0j<ri}0\leq j<r_{i}\} and let r=max{ri,j:0i<sr=\max\{r_{i,j}:0\leq i<s and 0j<ri}0\leq j<r_{i}\}.

Suppose that nAn\in A is such that

  1. (a)

    {Vi,j,k:0i<s,0j<ri\{V_{i,j,k}:0\leq i<s,0\leq j<r_{i} and 0k<ri,j}{Wl:0l<t}0\leq k<r_{i,j}\}\cup\{W_{l}:0\leq l<t\} is a family of open arcs of length ϵ\epsilon;

  2. (b)

    δ(ψ(β))ϵ\delta(\psi(\beta))\geq\epsilon for each βsuppψ\beta\in\operatorname{supp}\psi;

  3. (c)

    ϵ>3Nρmax({gl(n):0l<t}{fi,j,k(n):0i<s,0j<ri,0k<ri,j})\epsilon>3N\cdot\rho\cdot\max\bigl{(}\{\|g_{l}(n)\|:0\leq l<t\}\cup\bigcup\{\|f_{i,j,k}(n)\|:0\leq i<s,0\leq j<r_{i},0\leq k<r_{i,j}\}\bigr{)};

  4. (d)

    3MNsγ<ϵ3MNs\gamma<\epsilon;

  5. (e)

    |fi,ri1,0(n)(ξi(n))|γ>3L|f_{i,r_{i}-1,0}(n)(\xi_{i}(n))|\cdot\gamma>3L for each 0i<s0\leq i<s;

  6. (f)

    |fi,j1,0(n)(ξi(n))|ϵ6ri,j|fi,j,0(n)|>3L|f_{i,j-1,0}(n)(\xi_{i}(n))|\cdot\frac{\epsilon}{6\sqrt{r_{i,j}}|f_{i,j,0}(n)|}>3L for each 0i<s0\leq i<s and 0<j<ri0<j<r_{i};

  7. (g)

    |θi,j,kfi,j,k(n)(ξi(n))fi,j,0(n)(ξi(n))|<ϵ24rL\left|\theta_{i,j,k}-\frac{f_{i,j,k}(n)(\xi_{i}(n))}{f_{i,j,0}(n)(\xi_{i}(n))}\right|<\frac{\epsilon}{24\sqrt{r}L} for each i<si<s, j<rij<r_{i} and k<ri,jk<r_{i,j} and

  8. (h)

    suppψ{μ0(n),,μt1(n)}=\operatorname{supp}\psi\cap\{\mu_{0}(n),\ldots,\mu_{t-1}(n)\}=\emptyset.

Then there exists an arc function ϕ\phi such that

  1. (A)

    Nϕ(β)Nϕ(β)¯ψ(β)N\cdot\phi(\beta)\subseteq N\cdot\overline{\phi(\beta)}\subseteq\psi(\beta) for each βsuppψ\beta\in\operatorname{supp}\psi;

  2. (B)

    βsuppgl(n)gl(n)(β)ϕ(β)Wl\sum_{\beta\in\operatorname{supp}g_{l}(n)}g_{l}(n)(\beta)\phi(\beta)\subseteq W_{l} for each l<tl<t;

  3. (C)

    βsuppfi,j,k(n)fi,j,k(n)(β)ϕ(β)Vi,j,k\sum_{\beta\in\operatorname{supp}f_{i,j,k}(n)}f_{i,j,k}(n)(\beta)\cdot\phi(\beta)\subseteq V_{i,j,k} for each i<si<s, j<rij<r_{i} and k<ri,jk<r_{i,j};

  4. (D)

    δ(ϕ(β))=ρ\delta(\phi(\beta))=\rho for each βsuppϕ\beta\in\operatorname{supp}\phi and

  5. (E)

    suppϕ\operatorname{supp}\phi can be chosen to be any finite set containing

    suppψ0i<s,0j<ri,0k<ri,jsuppfi,j,k(n)0l<tsuppgl(n).\operatorname{supp}\psi\cup\bigcup_{0\leq i<s,0\leq j<r_{i},0\leq k<r_{i,j}}\operatorname{supp}f_{i,j,k}(n)\cup\bigcup_{0\leq l<t}\operatorname{supp}g_{l}(n).\qed

Now we are ready to prove Lemma 4.6.

Proof of Lemma 4.6.

Write ={u0,uq1}\mathcal{F}=\{u_{0},\dots u_{q-1}\} with no repetition. Let 𝒮\mathcal{S} be an integer stack on ApA^{\prime}\in p and let NN be a positive integer such that (1N𝒮,A,p)\left(\frac{1}{N}\mathcal{S},A^{\prime},p\right) is associated to \mathcal{F}.

As in Definition 5.1 the components of 𝒮\mathcal{S} will be denoted ss, tt, MM, (ri:i<s)(r_{i}:i<s), (ri,j:i<s,j<ri)(r_{i,j}:i<s,j<r_{i}), (fi,j,k:i<s,j<ri,k<ri,j)(f_{i,j,k}:i<s,j<r_{i},k<r_{i,j}), (gl:l<t)(g_{l}:l<t), (ξi:i<s)(\xi_{i}:i<s), (μp:i<t)(\mu^{p}:i<t) and (θi,j,k:0i<s,0j<ri,k<ri,j)(\theta_{i,j,k}:0\leq i<s,0\leq j<r_{i},k<r_{i,j}).

We write {fi,j,k:i<sp,j<ri,k<ri,j}{gl:l<t}\{f_{i,j,k}:i<s_{p},j<r_{i},k<r_{i,j}\}\cup\{g_{l}:l<t\} as {v0,,vq1}\{v_{0},\ldots,v_{q-1}\}.

Let \mathcal{M} be the q×qq\times q matrix of integer numbers such that Nui(n)=j<qi,jvj(n)Nu_{i}(n)=\sum_{j<q}\mathcal{M}_{i,j}v_{j}(n) for all nAn\in A and i<qi<q.

By (#)(\#) in Lemma 5.3, each vjv_{j} is an integer combination of the uiu_{i}’s, therefore the inverse matrix of 1N\frac{1}{N}\mathcal{M}, which we denote by 𝒩\mathcal{N}, has integer entries.

Let ϵ=ϵ(i,j<l|i,j|)1\epsilon^{\prime}=\epsilon\cdot(\sum_{i,j<l}|\mathcal{M}_{i,j}|)^{-1} and γ<ϵ/(3MNs)\gamma<\epsilon^{\prime}/(3MNs). Let LL be larger than or equal to the maximum of the set {L(θi,j,0,,θi,j,ri,j1,ϵ/24):i<s,j<ri}\{L(\theta_{i,j,0},\ldots,\theta_{i,j,r_{i,j}-1},\epsilon^{\prime}/24):i<s,j<r_{i}\}.

For each nAn\in A^{\prime}, let δn<12\delta_{n}<\frac{1}{2} be such that:

ϵ>3Nmax({gl(n):0l<t}{fi,j,k(n):0i<s,0j<ri,0k<ri,j})δnN\epsilon^{\prime}>3N\cdot\max\bigl{(}\{\|g_{l}(n)\|:0\leq l<t\}\cup\bigcup\{\|f_{i,j,k}(n)\|:0\leq i<s,0\leq j<r_{i},0\leq k<r_{i,j}\}\bigr{)}\cdot\frac{\delta_{n}}{N}

We note that both NN’s above cancel but we write this way as we will use δn/N{\delta_{n}}/{N} in the place of ρ\rho in item c) of Lemma 5.5.

Let r=max{ri,j:0i<s,0j<ri}r=\max\{r_{i,j}:0\leq i<s,0\leq j<r_{i}\}. Let AA be the set of nn’s in AA^{\prime} such that:

  • |fi,ri1,0(n)(ξi(n))|γ>3L for each 0i<s,\displaystyle|f_{i,r_{i}-1,0}(n)(\xi_{i}(n))|\gamma>3L\text{ for each }0\leq i<s,,

  • |fi,j1,0(n)(ξi(n))|ϵ6ri,j|fi,j,0(n)|>3L\displaystyle|f_{i,j-1,0}(n)(\xi_{i}(n))|\cdot\frac{\epsilon^{\prime}}{6\sqrt{r_{i,j}}|f_{i,j,0}(n)|}>3L for each 0i<s0\leq i<s and 0<j<ri0<j<r_{i},

  • |θi,j,kfi,j,k(n)(ξi(n))fi,j,0(n)(ξi(n))|<ϵ24rL for each i<s,j<ri and k<ri,j,\displaystyle\left|\theta_{i,j,k}-\frac{f_{i,j,k}(n)(\xi_{i}(n))}{f_{i,j,0}(n)(\xi_{i}(n))}\right|<\frac{\epsilon^{\prime}}{24\sqrt{r}L}\text{ for each }i<s,j<r_{i}\text{ and }k<r_{i,j}, and

  • E{μ0(n),,μt1(n)}=.\displaystyle E\cap\{\mu_{0}(n),\ldots,\mu_{t-1}(n)\}=\emptyset.

Notice that AA is cofinite in AA^{\prime}, therefore ApA\in p.

We claim this AA and this sequence (δn:nA)(\delta_{n}:n\in A) work.

Fix nAn\in A.

Let (Uf:f)(U_{f}:f\in\mathcal{F}) be a family of arcs of length ϵ\epsilon and let ϱ\varrho be an arc function of length at least ϵ\epsilon with suppϱE\operatorname{supp}\varrho\subseteq E. We rewrite the family of arcs as (Ui:i<q)(U_{i}:i<q), where Ui=UfiU_{i}=U_{f_{i}} for each i<qi<q. For each i<qi<q let yiy_{i} be a real such that yi+y_{i}+\Z is the center of UiU_{i}. Let zj=i<q𝒩j,iyiNz_{j}=\sum_{i<q}\mathcal{N}_{j,i}\frac{y_{i}}{N} and, for each jj let RjR_{j} be the arc of center zjz_{j} and length ϵ\epsilon^{\prime}. Since 𝒩\mathcal{N} is a matrix of integers, zj+=i<q𝒩j,i(yiN+)z_{j}+\Z=\sum_{i<q}\mathcal{N}_{j,i}(\frac{y_{i}}{N}+\Z). Then the arc j<qi,jRj\sum_{j<q}\mathcal{M}_{i,j}R_{j} is a subset of UiU_{i} for each i<qi<q.

Now we aim to apply Lemma 5.5. Set ψ=ϱ\psi=\varrho, ρ=δn/N\rho=\delta_{n}/N and ϵ\epsilon^{\prime} in the place of ϵ\epsilon. For i<si<s, j<rij<r_{i}, k<ri,jk<r_{i,j} we put Vi,j,k=RxV_{i,j,k}=R_{x} if fi,j,k=vxf_{i,j,k}=v_{x} for some x<qx<q, and for j<tj<t we put Wj=RxW_{j}=R_{x} if gj=vxg_{j}=v_{x} for some x<qx<q.

Then there exists an arc function ψn~\tilde{\psi_{n}} such that

  1. (A)

    Nψn~Nψn~¯ϱ(β)N\tilde{\psi_{n}}\subseteq N\overline{\tilde{\psi_{n}}}\subseteq\varrho(\beta) for each βsuppψ\beta\in\operatorname{supp}\psi;

  2. (B)

    βsuppgl(n)gl(n)(β)ψn~(β)Wl\sum_{\beta\in\operatorname{supp}g_{l}(n)}g_{l}(n)(\beta)\tilde{\psi_{n}}(\beta)\subseteq W_{l} for each l<tl<t;

  3. (C)

    βsuppfi,j,k(n)fi,j,k(n)(β)ψn~(β)Vi,j,k\sum_{\beta\in\operatorname{supp}f_{i,j,k}(n)}f_{i,j,k}(n)(\beta)\cdot\tilde{\psi_{n}}(\beta)\subseteq V_{i,j,k} for each i<si<s, j<rij<r_{i} and k<ri,jk<r_{i,j};

  4. (D)

    δ(ψn~(β))=δn/N\delta(\tilde{\psi_{n}}(\beta))={\delta_{n}}/N for each βsuppψn~\beta\in\operatorname{supp}\tilde{\psi_{n}} and

  5. (E)

    suppψn~\operatorname{supp}\tilde{\psi_{n}} is equal to

    0i<s,0j<ri,0k<ri,jsuppfi,j,k(n)0l<tsuppgl(n)E=fsuppf(n)E.\bigcup_{0\leq i<s,0\leq j<r_{i},0\leq k<r_{i,j}}\operatorname{supp}f_{i,j,k}(n)\cup\bigcup_{0\leq l<t}\operatorname{supp}g_{l}(n)\cup E=\bigcup_{f\in\mathcal{F}}\operatorname{supp}f(n)\cup E.

Let ψn=Nψn~\psi_{n}=N\tilde{\psi_{n}}. By (A), ψnϱ\psi_{n}\leq\varrho. By (E) and (D), suppψn=fsuppf(n)E\operatorname{supp}\psi_{n}=\bigcup_{f\in\mathcal{F}}\operatorname{supp}f(n)\cup E and for each βsuppψn\beta\in\operatorname{supp}\psi_{n}, we have δ(ψn(β))=δn\delta(\psi_{n}(\beta))=\delta_{n}. Let S=suppψnS=\operatorname{supp}\psi_{n}. Now notice that given uiu_{i}\in\mathcal{F} we have:

μsuppuiui(n)(μ)ψn(μ)\displaystyle\sum_{\mu\in\operatorname{supp}u_{i}}u_{i}(n)(\mu)\psi_{n}(\mu) =μSui(n)(μ)Nψ~n(μ)\displaystyle=\sum_{\mu\in S}u_{i}(n)(\mu)N\tilde{\psi}_{n}(\mu)
=μS(j<qi,jvj(n)(μ))ψ~n(μ)\displaystyle=\sum_{\mu\in S}\left(\sum_{j<q}\mathcal{M}_{i,j}v_{j}(n)(\mu)\right)\tilde{\psi}_{n}(\mu)
=j<qi,j(μSvj(n)(μ)ψ~n(μ))\displaystyle=\sum_{j<q}\mathcal{M}_{i,j}\left(\sum_{\mu\in S}v_{j}(n)(\mu)\tilde{\psi}_{n}(\mu)\right)

Then by (B), (C) and the definitions of the WlW_{l}’s and Vi,j,kV_{i,j,k}’s:

μsuppuiui(n)(μ)ψn(μ)=μSui(n)(μ)Nψ~n(μ)j<qi,jRjUi.\sum_{\mu\in\operatorname{supp}u_{i}}u_{i}(n)(\mu)\psi_{n}(\mu)=\sum_{\mu\in S}u_{i}(n)(\mu)N\tilde{\psi}_{n}(\mu)\subseteq\sum_{j<q}\mathcal{M}_{i,j}R_{j}\subseteq U_{i}.

As intended. ∎

6. Final comments

The method to construct countably compact free Abelian groups came from the technique to construct countably compact groups without non-trivial convergent sequences. It is not known if there is an easier method to produce countably compact group topologies on free Abelian groups if we do not care if the resulting topology has convergent sequences.

In fact, even to produce a countably compact group topology with convergent sequences in non-torsion groups it is used a modification of the technique to construct countably compact groups without non-trivial convergent sequences, see [1] and [2].

The first examples of countably compact groups without non-trivial convergent sequences were obtained by Hajnal and Juhász [11] under CH. E. van Douwen [7] obtained an example from MA and asked for a ZFC example. Other examples were obtained using MAcountable\mathrm{MA_{countable}} [15], a selective ultrafilter [10] and in the Random real model [19]. Only recently, Hrušak, van Mill, Shelah and Ramos obtained an example in ZFC ([13]).

This motivates the following questions in ZFC:

Question 6.1.

Are there large countably compact groups without non-trivial convergent sequences in ZFC? Is there an example of cardinality 2𝔠2^{\mathfrak{c}}?

The example of Hrušak et al has size continuum and it is not clear if their construction could yield larger examples.

Question 6.2.

Is there a countably compact free Abelian group in ZFC? A countably compact free Abelian group without non-trivial convergent sequences in ZFC?

It is still open if there exists a torsion-free group in ZFC that admits a countably compact group topology without non-trivial convergent sequences. If such example exists then there is a countably compact group topology without non-trivial convergent sequences in the free Abelian group of cardinality 𝔠\mathfrak{c} (see [25] or [27]).

Question 6.3.

Is there a both-sided cancellative semigroup that is not a group that admits a countably compact semigroup topology (a Wallace semigroup) in ZFC?

The known examples were obtained in [18] under CH, in [21] under MAcountable\mathrm{MA_{countable}}, in [17] from 𝔠\mathfrak{c} incomparable selective ultrafilters and in [2] from one selective ultafilter. The last two use the known fact that a free Abelian group without non-trivial convergent sequences contains a Wallace semigroup, which was used in [18]. The example in [21] was a modification of [12].

References

  • [1] M. K. Bellini, A. C. Boero, I. Castro-Pereira, V. O. Rodrigues, and A. H. Tomita, Countably compact group topologies on non-torsion abelian groups of size 𝔠\mathfrak{c} with non-trivial convergent sequences, Topology Appl. (2019), 106894.
  • [2] A. C. Boero, I. Castro-Pereira, and A. H. Tomita, Countably compact group topologies on the free Abelian group of size continuum (and a Wallace semigroup) from a selective ultrafilter, Acta Math. Hungarica 159(2) (2019), 414–428.
  • [3] by same author, Countably compact group topologies on the free abelian group of size continuum (and a wallace semigroup) from a selective ultrafilter, Acta Math. Hungarica 159(2) (2019), 414–428.
  • [4] A. C. Boero and A. H. Tomita, A group topology on the free Abelian group of cardinality 𝔠\mathfrak{c} that makes its square countably compact, Fund. Math. 212 (2011), 235–260.
  • [5] I. Castro-Pereira and A. H. Tomita, Abelian torsion groups with a countably compact group topology, Topology Appl. 157 (2010), 44–52.
  • [6] W. W. Comfort, K. H. Hofmann, and D. Remus, Topological groups and semigroups, Recent progress in general topology (M. Husek and J. van Mill, eds.), North-Holland, 1992, pp. 57–144.
  • [7] E. K. van Douwen, The product of two countably compact topological groups, Trans. Amer. Math. Soc. 262 (1980), 417–427.
  • [8] by same author, The weight of a pseudocompact (homogeneous) space whose cardinality has countable cofinality, Proc. Amer. Math. Soc. 80 (1980), 678–682.
  • [9] L. Fuchs, Infinite abelian groups, Pure and Applied Mathematics, Elsevier Science, 1970.
  • [10] S. Garcia-Ferreira, A. H. Tomita, and S. Watson, Countably compact groups from a selective ultrafilter, Proc. Amer. Math. Soc. 133 (2005), 937–943.
  • [11] A. Hajnal and L. Juhász, A separable normal topological group need not be Lindelöf, Gen. Topology Appl. 6 (1976), 199–205.
  • [12] K. Hart and J. van Mill, A countably compact group H{H} such that H×H{H}\times{H} is not countably compact, Trans. Amer. Math. Soc. 323 (1991), 811–821.
  • [13] M. Hrušák, J. van Mill, U. A. Ramos-García, and S. Shelah, Countably compact groups without non-trivial convergent sequences, To appear in Trans. Amer. Math. Soc.DOI: https://doi.org/10.1090/tran/8222, 2020.
  • [14] T. Jech, Set theory, Springer, 2003.
  • [15] P. B. Koszmider, A. H. Tomita, and S. Watson, Forcing countably compact group topologies on a larger free Abelian group, Topology Proc. 25 (2000), 563–574.
  • [16] K. Kunen, Set theory: an introduction to independence proofs, North Holland, 1983.
  • [17] R. E. Madariaga-Garcia and A. H. Tomita, Countably compact topological group topologies on free Abelian groups from selective ultrafilters, Topology Appl. 154 (2007), 1470–1480.
  • [18] D. Robbie and S. Svetlichny, An answer to A D Wallace’s question about countably compact cancellative semigroups, Proc. Amer. Math. Soc. 124 (1996), 325–330.
  • [19] P. J. Szeptycki and A. H. Tomita, HFD groups in the Solovay model, Topology Appl. 156 (2009), 1807–1810.
  • [20] M. G. Tkachenko, Countably compact and pseudocompact topologies on free Abelian groups, Soviet Math. (Izv. VUZ) 34 (1990), 79–86.
  • [21] A. H. Tomita, The Wallace problem: a counterexample from MAcountable\textrm{MA}_{\textrm{countable}} and pp-compactness, Canad. Math. Bull. 39 (1996), 486–498.
  • [22] by same author, The existence of initially ω1\omega_{1}-compact group topologies on free Abelian groups is independent of ZFC, Comment. Math. Univ. Carolinae 39 (1998), 401–413.
  • [23] by same author, Two countably compact topological groups: one of size ω\aleph_{\omega} and the other of weight ω\aleph_{\omega} without non-trivial convergent sequences, Proc. Amer. Math. Soc. 131 (2003), 2617–2622.
  • [24] by same author, A solution to Comfort’s question on the countable compactness of powers of a topological group, Fund. Math. 186 (2005), 1–24.
  • [25] by same author, Square of countably compact groups without non-trivial convergent sequences, Topology Appl. 153(1) (2005), 107–122.
  • [26] by same author, A group topology on the free abelian group of cardinality 𝔠{\mathfrak{c}} that makes its finite powers countably compact, Topology Appl. 196 (2015), 976–998.
  • [27] by same author, A van Douwen-like ZFC theorem for small powers of countably compact groups without non-trivial convergent sequences, Topology Appl. 259 (2019), 347–364.