Countably compact group topologies on arbitrarily large free Abelian groups
Abstract.
We prove that if there are incomparable selective ultrafilters then, for every infinite cardinal such that , there exists a group topology on the free Abelian group of cardinality 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 problem2010 Mathematics Subject Classification:
Primary 54H11, 22A05; Secondary 54A35, 54G20.1. 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 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 selective ultrafilters, the example in [17] showed the consistency of a countably compact group topology on the free Abelian group of cardinality . All forcing examples obtained so far had their cardinalities bounded by .
Boero and Tomita [4] showed from the existence of selective ultrafilters that there exists a free Abelian group of cardinality whose square is countably compact. Tomita [26] showed that there exists a group topology on the free Abelian group of cardinality 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 . Then a free Abelian group of infinite cardinality can be endowed with a countably compact group topology (without non-trivial convergent sequences) if and only if .
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 be a filter on and let be a sequence in a topological space . We say that is a -limit point of if, for every neighborhood of , the set belongs to .
If is Hausdorff, every sequence has at most one -limit and we write in that case.
The set of all free ultrafilters on is denoted by . 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 -limit point for some .
The concept of almost disjoint families will be useful in our construction.
Definition 1.5.
An almost disjoint family is an infinite family of infinite subsets of such that distinct elements of 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 will be the metric group where the metric is given by
for every .
Given an open interval of with , we let .
An arc of is a set of the form , where is an open interval of . An arc is said to be proper if it is distinct from .
If is a proper arc and , then the Euclidean length of equals the Euclidean length of , and we define the length of as . We also let .
Given an arc such that , it follows that .
Our free Abelian groups will all be represented as directs sums of copies of the group of integers ; we fix some notation. The additive group of rationals will also be used, so in the following definition one should read or for .
Definition 1.7.
If is a map from a set to a group then the support of , denotes is defined to be the set .
We define .
If is a subset of then, as an abuse of notation, we often write .
Given , we denote by the characteristic function of , whose support is and which value .
For a sequence in we define by .
Finally, for , we let be the constant sequence with value .
Note that then is also constant, with value .
Definition 1.8.
Let be a filter on and a set. We say that the sequences are -equivalent and write iff .
It is easy to verify that is an equivalence relation. We denote the equivalence class of by . We also denote the set of all equivalence classes by .
If is a ring and is an -module, then has a natural -module structure given by , , and the class of the zero function as its zero element.
If is a free ultrafilter, then the ultrapower of the -module by is the -module .
For the rest of this paper we will fix a cardinal number that satisfies .
Throughout this article, we will work inside ultrapowers of . These ultrapowers contain copies of ultrapowers of , which will be useful for the construction. So it is useful to define some notation.
Definition 1.9.
Let be a free ultrafilter on . We define as the -vector space and with the subgroup structure.
Notice that each in is formally an element of , not of . Nevertheless it is clear that is isomorphic to via the obvious isomorphism that carries the equivalence class of a sequence in to its class in .
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 ), also called Ramsey ultrafilter, is a free ultrafilter on with the property that for every partition of , either there exists such that or there exists such that for every .
The following proposition is well known. We provide [14] as a reference.
Proposition 2.2.
Let be a free ultrafilter on . Then the following are equivalent:
-
a)
is a selective ultrafilter,
-
b)
for every , there exists such that is either constant or one-to-one on ,
-
c)
for every function there exists such that is constant on .
The Rudin-Keisler order is defined as follows:
Definition 2.3.
Let be a filter on and . We define .
It is easy to verify that is a filter; if is an ultrafilter then so is ; if , then ; and is the identity over the set of all filters. This implies that if is bijective, then .
Definition 2.4.
Let and be filters. We say that (or , if we need to be clear) iff there exists such that .
The Rudin-Keisler order is the set of all free ultrafilters over ordered by . We say that two ultrafilters and are equivalent iff and .
It is easy to verify that 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 without modifying the equivalence classes. We refer to [14] for the following proposition:
Proposition 2.5.
The following are true:
-
(1)
If and are ultrafilters, then and is equivalent to the existence of a bijection such that .
-
(2)
The selective ultrafilters are exactly the minimal elements of the Rudin-Keisler order.
This implies that if and is a selective ultrafilter, then is either a fixed ultrafilter or a selective ultrafilter. If is the ultrafilter generated by , then , so, in particular, if is finite to one and is selective, then is a selective ultrafilter equivalent to .
The existence of selective ultrafilters is independent from ZFC. Martin’s Axiom for countable orders implies the existence of pairwise incomparable selective ultrafilters in the Rudin-Keisler order.
The lemma below appears in [24].
Lemma 2.6.
Let be a family of pairwise incomparable selective ultrafilters. For each let be a strictly increasing sequence in such that and for all . Then there exists such that:
-
a)
, for each .
-
b)
whenever and , and
-
c)
and is a pairwise disjoint family.
In the course of the construction we will often use families of ultrafilters indexed by and finite sequences of infinite subsets of . The following definition fixes some convenient notation.
Definition 2.7.
A finite tower in is a finite sequence of infinite subsets of such that for every . The set of all finite towers in is called . If then , the last term of the sequence . For the empty sequence we write .
Lemma 2.8.
Assume there are incomparable selective ultrafilters. Then there is a family of incomparable selective ultrafilters such that whenever and .
Proof.
Index the incomparable selective ultrafilters faithfully as . For each , let be a bijection and define . Since is one-to-one, it follows that is a selective ultrafilter equivalent to . The family is as required. ∎
3. Main Ideas
From now on we fix a family of selective ultrafilters as provided by Lemma 2.8.
The idea will be to use these ultrafilters to assign -limits to enough injective sequences in 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 was part of this family. Since this group has cardinality , 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 be a subset of and . We shall call linearly independent mod iff for every free ultrafilter with the set
is linearly independent in the -vector space , and if whenever and are distinct elements of .
Notice that it is implicit in our definition that and are disjoint. We will abbreviate “linearly independent mod ” to l.i. mod .
An application of Zorn’s Lemma will establish the following lemma.
Lemma 3.2.
Every set of sequences that is l.i. mod can be extended to a maximal linearly independent set mod .∎
It should be clear that and and are infinite, then a set that is l.i. mod is also l.i. mod . Then by using recursion, this easily implies the following corollary:
Corollary 3.3.
There exists a family such that:
-
(1)
For every the set is maximal l.i. mod , and
-
(2)
For every , if then .
We note explicitly that even though is only demanded to be maximal l.i. mod it will, because of item (2), depend on all of , not just on .
Lemma 3.4.
Let be an element of and let be maximal l.i. mod . Then there exist an infinite subset of , a finite subset of , a finite subset of , and sets and of rational numbers such that
Proof.
If or for some , then we are done. Otherwise, by the maximality of , there exists a free ultrafilter with such that the set
is not linearly independent.
This means that we can find finite subsets and of and respectively and finite sets and of rational numbers such that
Now choose with that witnesses this equality. ∎
Corollary 3.5.
If is maximal l.i. mod , then .
Proof.
First notice that . Assume . Then the set has cardinality less than .
Take some injective sequence in and define by for all . Clearly then is disjoint from , all values of are non-zero and the values have disjoint supports.
Apply Lemma 3.4 to obtain sets , , , , and such that
Since is infinite and is finite, there is a such that . Now when because , and when because , and also , which contradicts . ∎
Henceforth we fix a family as in Corollary 3.3 and enumerate each faithfully as .
Definition 3.6.
For each and , we denote by the intersection of and the free Abelian group generated by .
For the next lemma, we are going to use the following proposition:
Proposition 3.7.
If is an abelian group and is a subgroup of such that is an infinite cyclic group, then there exists such that .
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 has a basis of the form for some subset of .
Proof.
Let the the group generated by if and by the union of and when .
Let for all .
For every we shall find so that , as follows.
For the group is generated by , so and we have .
For observe that , so:
The group is cyclic infinite, so either is infinite and cyclic or . By Proposition 3.7 there exists such that (and in case ). Take such that .
For every , it follows that . Since , it follows that .
The set is as required. ∎
For the rest of this article we fix such a set as above for each pair in .
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 in every sequence in has a -limit in . Then every finite power of is countably compact.
Proof.
A sequence in some finite power of is represented by finitely members of , say . We show that there is one ultrafilter such that exists for all , namely for a suitable and .
Recursively, we define a tower and for finite subsets and of and respectively together with finite sets and of rational numbers such that
For , use Lemma 3.4 applied to to obtain , , , and such that holds with .
To go from to apply Lemma 3.4 to to obtain , , , , and so that holds for .
Let and let be sufficiently large so that and are integers, for all , , and . Then for all .
As and for each is a subset of , it follows that . Therefore, each is an integer combination of . Then, by hypothesis, it follows that each has a -limit. This completes the proof. ∎
4. Constructing homomorphisms
Through this section, we let and we let be an enumeration of such that whenever and , and so that each element of appears at least many times.
Lemma 4.1.
There exists a family of pairwise disjoint subsets of such that .
Proof.
For each there is an injective map . Let and we are done. ∎
For the rest of this section, we fix a family as above.
The following lemma is the key to the main result.
Lemma 4.2.
Assume we have a non-zero element of , an injective sequence in , and a countably infinite subset of such that
-
(1)
,
-
(2)
for infinitely many ’s and,
-
(3)
for all and
Then there exists a homomorphism such that:
-
(1)
-
(2)
, whenever , , and .
-
(3)
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 is a non-zero element of and is an injective sequence in . Then there exists a homomorphism such that
-
(1)
-
(2)
, whenever , and .
-
(3)
does not converge.
Proof.
Using a closing-off argument construct a countable subset of that intersects infinitely many sets , and that contains , , for all as well as whenever and .
By the previous Lemma, there exists a homomorphism such that , does not converge, and whenever , and .
We let be the monotone enumeration of . For , let . So and .
Recursively, we construct, for , an increasing sequence of homomorphisms such that whenever , and . Our homomorphism will be . The basis step is already done, and for limit steps, we just unite all previous homomorphisms.
To define given it suffices to specify the value .
If for some and then we put . This is well defined because for all and because is compact. In the other case let . ∎
We can now prove our main result.
Theorem 4.4.
Assume the existence of pairwise incompatible selective ultrafilters and that is an infinite cardinal such that . Then the free abelian group of cardinality 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 and an injective sequence in , Lemma 4.3 provides a homomorphism such that , such that does not converge, and such that whenever , and . We give the initial topology generated by the collection of homomorphisms , is injective thus obtained and the natural topology of .
Since the initial topology generated by any collection of group homomorphisms is a group topology we do indeed obtain a group topology. Since is Hausdorff and for every there are many with it follows at once that our topology is Hausdorff.
To see that every finite power of is countably compact we use Lemma 3.9.
Given , and , there exist such that . For every and injective , we have . So and we are done.
Since for a given injective sequence and any the sequence does not converge and is continuous, it follows that does not converge. So 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 . An -arc function is a function from into the set of open arcs of (including itself) such that for all either or the length of is equal to , and the set is finite. We will call this finite set the support of and denote it by .
Given two arc functions and we write if or for each .
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 be a selective ultrafilter and a finite subset of such that the set is linearly independent.
Then for a given and a finite subset of there exist and a sequence of positive real numbers such that
-
whenever is a family of arcs of length and is an arc function of length at least with there exist for each a -arc function such that , and for each .
Now we proceed to prove Lemma 4.2. We will use the following lemma:
Lemma 4.7.
Let be a sequence of countable subsets of and let is a sequence of pairwise incomparable selective ultrafilters such that for each the set is linearly independent and whenever in . Furthermore let for every an ordinal in be given. In addition let and be non-zero in and with disjoint supports. Finally, let be a countable subset of that contains and for every .
Then there exists a homomorphism such that , and , whenever and .
Proof.
Write as the union of an increasing sequence of finite nonempty subsets, and take a similar sequence for each .
Take a sufficiently small positive number and an -arc function such that and .
Let and for each .
We will define, by recursion, for : finite sequences , finite sets , and real numbers satisfying:
-
(1)
For all and in we have ,
-
(2)
For each and , we have a sequence of positive real numbers such that: if is a family of arcs of length and is an arc function of length and then for each there exists a -arc function with , and , and for each .
-
(3)
For all and we have .
-
(4)
and .
Suppose we have defined for all as well as and for all . As will be clear from the step below the set is only non-trivial whenever . Therefore we let for and we concentrate on the case .
Let . By Lemma 4.6, there exist and that satisfy (2) for . Without loss of generality we can assume that .
Condition (4) now specifies .
Setting completes the recursion.
For each , apply the selectivity of , to choose an increasing sequence with and such that and for all .
Next apply Lemma 2.6 and let be a sequence of pairwise disjoint subsets of such that and the family of intervals is pairwise disjoint. Without loss of generality we can assume that .
Enumerate in increasing order as . For each , let be such that . Thus, for each we have , and hence and .
By recursion we define a sequence of arc functions, , such that and .
We start with . Then we have , and , and .
Since has length at least , there exists an arc function of length such that , for each . We have by the definition that .
Suppose and that has been defined with length at least .
Apply item (2) to the arc function , the finite set , the number , the finite set , the arcs for , and to obtain an arc function such that for all , and has length .
Because and we get .
If then and the length of is not greater than which in turn is not larger than .
It follows that for all the intersection consists of a unique element; we define to be that element and extend to a group homomorphism.
By construction is in which is a subset of whenever . Therefore, the sequence converges to , for each and .
Furthermore , therefore, ; and likewise .
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 such that . We enumerate them faithfully as .
For let and . Let and be two ultrafilters that were not listed and let . For each and , let . Let and where are not in . Then, by applying Lemma 4.7 with , there exist satisfying (1) and (2). To see it also satisfies (3), notice that . ∎
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 that are usually associated to a selective ultrafilter. Given an finite set of sequences it is possible to associate it to a integer stack which generates the same vector space as . 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 and the sequences of the stack helps us to transform these arcs into arcs that work for the functions of . Below, we give the definition of integer stack.
Definition 5.1.
An integer stack on consists of
-
(i)
an infinite subset of ;
-
(ii)
natural numbers , , and ; positive integers for and positive integers for and ;
-
(iii)
functions for , and and elements for ;
-
(iv)
sequences for and for and
-
(v)
real numbers for , and
These are required to satisfy the following conditions:
-
(1)
for each ;
-
(2)
for each and ;
-
(3)
the elements of and are pairwise distinct;
-
(4)
for each and ;
-
(5)
is a linearly independent subset of as a -vector space for each and ;
-
(6)
for each , and ;
-
(7)
the sequence diverges monotonically to , for each , and ;
-
(8)
for each , , and ;
-
(9)
converges monotonically to for each , , , and ; and
-
(10)
for each , and .
It is not difficult to show that the sequences of the stack are linearly independent. Moreover, if is a free ultrafilter, is a stack over , and , then it is not difficult to see that is linearly independent in the -vector space . We leave the details as an exercise to the reader.
Definition 5.2.
Given an integer stack and a natural number , the root of , written , is obtained by keeping all the structure in with the exception of the functions; these are divided by . Thus a function is replaced by in for each , and and a function is replaced by in for each .
A stack (unqualified) is then defined to be the root of an integer stack for some positive integer .
The lemma below gives the relation between a finite sequence of sequences in and a stack 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 , …, be sequences in and a selective ultrafilter. Then there exists and a stack on such that: if the elements of the stack have a -limit in then has a -limit in for each .
We will say in this case that the finite sequence is associated to .
-
If is a -linearly independent set and the group generated by it does not contain nonzero constant classes, then each restriction is an integer combination of the stack on . On the other hand, each element of the integer stack is an integer combination of restricted to .
Proof.
We prove . All numbered references in this proof are to the paper [26].
First, notice that if is a -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 ’s as the ’s themselves, we see that the functions , …, are integer combinations of the stack that was constructed.
It remains to see that the functions of are integer combinations of the functions restricted to . First, notice that in the statement of Lemma 5.4, by x), xi), xii) and xiv) the functions and are integer combinations of the . This Lemma is used in the proof of Lemma 5.5, where the functions become the functions , so there are integer combinations of the ’s.
Now notice that in Lemma 6.1, by g), c) and finite induction, the functions are integer combinations of the , and some of these become the ’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 ’s (restricted to ). ∎
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 is a linearly independent subset of the -vector space and then denotes a positive integer, , such that is -dense in in the usual Euclidean metric product topology, for any interval of length at least .
The last lemma we are going to need is Lemma 8.3 from [26], stated below.
Lemma 5.5.
Let , and be positive reals, a positive integer and be an arc function. Let be an integer stack on and , , , , , , , , and be as in Definition 5.1.
Let be an integer greater or equal to and and let and .
Suppose that is such that
-
(a)
and is a family of open arcs of length ;
-
(b)
for each ;
-
(c)
;
-
(d)
;
-
(e)
for each ;
-
(f)
for each and ;
-
(g)
for each , and and
-
(h)
.
Then there exists an arc function such that
-
(A)
for each ;
-
(B)
for each ;
-
(C)
for each , and ;
-
(D)
for each and
-
(E)
can be chosen to be any finite set containing
Now we are ready to prove Lemma 4.6.
Proof of Lemma 4.6.
Write with no repetition. Let be an integer stack on and let be a positive integer such that is associated to .
As in Definition 5.1 the components of will be denoted , , , , , , , , and .
We write as .
Let be the matrix of integer numbers such that for all and .
By in Lemma 5.3, each is an integer combination of the ’s, therefore the inverse matrix of , which we denote by , has integer entries.
Let and . Let be larger than or equal to the maximum of the set .
For each , let be such that:
We note that both ’s above cancel but we write this way as we will use in the place of in item c) of Lemma 5.5.
Let . Let be the set of ’s in such that:
-
•
,
-
•
for each and ,
-
•
and
-
•
Notice that is cofinite in , therefore .
We claim this and this sequence work.
Fix .
Let be a family of arcs of length and let be an arc function of length at least with . We rewrite the family of arcs as , where for each . For each let be a real such that is the center of . Let and, for each let be the arc of center and length . Since is a matrix of integers, . Then the arc is a subset of for each .
Now we aim to apply Lemma 5.5. Set , and in the place of . For , , we put if for some , and for we put if for some .
Then there exists an arc function such that
-
(A)
for each ;
-
(B)
for each ;
-
(C)
for each , and ;
-
(D)
for each and
-
(E)
is equal to
Let . By (A), . By (E) and (D), and for each , we have . Let . Now notice that given we have:
Then by (B), (C) and the definitions of the ’s and ’s:
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 [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 ?
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 (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 , in [17] from 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 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 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 such that 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 and -compactness, Canad. Math. Bull. 39 (1996), 486–498.
- [22] by same author, The existence of initially -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 and the other of weight 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 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.