On some topological realizations of groups and homomorphisms
Abstract
Let be a homomorphism of groups. We construct a topological space such that its group of homeomorphisms is isomorphic to , its group of homotopy classes of self-homotopy equivalences is isomorphic to and the natural map between the group of homeomorphisms of and the group of homotopy classes of self-homotopy equivalences of is . In addition, we consider realization problems involving homology, homotopy groups and groups of automorphisms.
1 Introduction
††2020 Mathematics Subject Classification: 20B25, 06A06, 06A11 05E18, 55P10, 55P99.††Keywords: Automorphism group, homotopy equivalence, Alexandroff spaces, posets, homology groups, homotopy groups.††This research is partially supported by Grants PGC2018-098321-B-100 and BES-2016-076669 from Ministerio de Ciencia, Innovación y Universidades (Spain).Alexandroff spaces are topological spaces with the property that the arbitrary intersection of open sets is open. That sort of topological spaces was first studied by P.S. Alexandroff in [1], where it is shown that they can also be seen as partially ordered sets. This viewpoint can be used to express topological notions in combinatorial terms. A particular case of Alexandroff spaces are finite topological spaces. There are two foundational papers on this subject that were published independently in 1966 [15, 17]. In [17], R.E. Stong made an analysis of the homeomorphism classification of finite topological spaces using matrices and also introduced combinatorial techniques to study their homotopy type. In [15], M.C. McCord studied the singular homology groups and homotopy groups of Alexandroff spaces proving that for every Alexandroff space there exists a simplicial complex sharing the same homotopy groups and singular homology groups. In fact, it is shown that there is a continuous map inducing isomorphism on all homotopy groups. The converse result is also obtained, that is, given a simplicial complex , there exists an Alexandroff space having the same singular homology groups and homotopy groups of .
Finite topological spaces or Alexandroff spaces are a good tool to solve realization problems: given a category and a group , is there an object in such that the group of automorphisms of is isomorphic to ? In [6, 18, 5], it is proved that for every finite group there exists a finite topological space such that its group of homeomorphisms is isomorphic to . Recently, in [4], J.A. Barmak constructed a finite topological space with points and lower cardinality than the finite topological spaces obtained in [6, 18, 5]. On the other hand, L. Babai in [2] obtained a finite topological space with points realizing as a group of homeomorphisms. This topological space has the disadvantage that it requires to find a “good” set of generators of satisfying a list of non-trivial properties. A generalization of these results for non-finite groups was made in [7], where the realization problem for the homotopical category and the pointed homotopical category was also solved. To be precise, it was proved that for every group there exists an Alexandroff space such that its group of automorphisms in the topological category , and is isomorphic to .
The restriction of to topological spaces with the homotopy type of a CW-complex is denoted by . In and the pointed version , the realizability problem has a long history, see for instance [12]. Recently, C. Costoya and A. Viruel solved the realization problem in for finite groups in [9].
We introduce a bit of notation. Let be a topological space. Let us denote by the group of homeomorphisms of . Let denote the group of homotopy classes of self-homotopy equivalences of . Let denote the group of pointed homotopy classes of pointed self-homotopy equivalences of .
The following result answers a natural question posed by professor Jesús Antonio Álvarez López during a talk in the VIII Encuentro de Jóvenes Topólogos (A Coruña, 2019).
Lemma 1.1.
Let and be two groups. There exists a topological space such that is isomorphic to and is isomorphic to .
An immediate consequence of Lemma 1.1 is the following: for a general topological space there is no relation between its group of homeomorphisms and its group of homotopy classes of self-homotopy equivalences. Moreover, if and are two groups, then we can produce infinitely many non-homeomorphic topological spaces having as their group of homeomorphisms and having as their group of homotopy classes of self-homotopy equivalences.
For every topological space there is a natural homomorphism of groups sending each homeomorphism to its homotopy class . Given two groups and , we consider the topological space obtained in Lemma 1.1. The kernel of corresponds precisely to . The image of is the homotopy class of the identity map. However, modifying the construction of the topological space , we can obtain a stronger version of Lemma 1.1.
Theorem 1.2.
Let be a homomorphism of groups. There exists a topological space such that , and .
Theorem 1.2 clearly generalizes Lemma 1.1. We prefer to prove Lemma 1.1 first for the sake of exposition. Omitting Lemma 1.1, the proof of Theorem 1.2 becomes less intuitive. In addition, subsequent results are obtained using the topological space given in Lemma 1.1.
Given a topological space and a non-negative integer number , we can consider the -th homology group of or the -th homotopy group of . It is natural to consider more realization problems involving these groups, the group of homeomorphisms and the group of homotopy classes of self-homotopy equivalences.
Theorem 1.3.
Let and be finite groups and let be a topological space with the homotopy type of a compact CW-complex. There exists an Alexandroff space such that is isomorphic to , is isomorphic to and is weak homotopy equivalent to , which implies that is isomorphic to and is isomorphic to for every .
As an immediate consequence of Theorem 1.3, we can deduce the following corollaries.
Corollary 1.4.
Let and be finite groups and let be a set of finitely generated Abelian groups, where is a finite set. There exists a topological space such that is isomorphic to , is isomorphic to and is isomorphic to for every .
Corollary 1.5.
Let and be finite groups, and let be a finitely presented (Abelian) group (if ). There exists a topological space such that is isomorphic , is isomorphic to and is isomorphic to .
Roughly speaking, these corollaries say that for a general topological space its group of automorphisms in or does not have any relation to its -th homology or homotopy group and vice versa. In contrast, for the category , the situation is completely different since contains normal subgroups that are nilpotent. For instance, given a topological space , we denote by the set of self-homotopy equivalences that induce the identity map in homotopy (homology). It is trivial to check that is a normal subgroup of and if is a finite -complex, then is a nilpotent group. See [10] for more details. Using the construction obtained in Theorem 1.3, we can find topological spaces that do not satisfy the previous properties. From this we get that some of the techniques used to study the group of self-homotopy equivalences for -complexes cannot be adapted in a natural way to general spaces.
The organization of the paper is as follows. In Section 2 we introduce basic concepts and results from the literature. In Section 3 we provide an example of one of the main results in order to motivate the main ideas of the proof of Lemma 1.1. In Section 4 we prove Lemma 1.1 and give some remarks. In Section 5 we prove Theorem 1.2. In Section 6 we define a sequence of topological spaces whose homotopy and homology groups are all trivial, their group of automorphisms in and are also trivial, but they are not homeomorphic or homotopy equivalent to a point. Then, we prove Theorem 1.3 as well as Corollary 1.4 and Corollary 1.5. Finally, we give examples of topological spaces satisfying that the groups and are not nilpotent in general.
2 Preliminaries
Definition 2.1.
Let and be topological spaces. A continuous function is said to be a weak homotopy equivalence if it induces isomorphisms on all the homotopy groups.
Definition 2.2.
An Alexandroff space is a topological space satisfying that the arbitrary intersection of open sets is open.
If is an Alexandroff space, then for every there exists a minimal open neighbourhood given by the intersection of every open set containing . denotes the set given by the intersection of every closed set containing . Trivially, every finite topological space is an Alexandroff space. An Alexandroff space is locally finite if for every the set is finite.
Let be a partially ordered set or poset. If , then we write () if and only if () and there is no such that (). We will denote by the maximum of if it exists. We denote by the cardinal numbers and . A set is called lower (upper) if for every and () we have .
It is not difficult to verify the following two properties:
-
•
For a partially ordered set the family of lower (upper) sets of is a topology on , that makes a Alexandroff space.
-
•
For a Alexandroff space, the relation if and only if () is a partial order on .
In addition, for a set , the Alexandroff space topologies on are in bijective correspondence with the partial orders on .
From now on, every Alexandroff space satisfies the separation axiom. The following results can be found, for instance, in [3, 14].
Proposition 2.3.
If is a map between Alexandroff spaces, then is continuous if and only if preserves the order.
From this and previous properties, the following result can be deduced.
Theorem 2.4.
The category of Alexandroff spaces is isomorphic to the category of partially ordered sets.
Hence, partially ordered sets and Alexandroff spaces can be treated as the same object.
Proposition 2.5.
Let be continuous maps between Alexandroff spaces. If () for every , then and are homotopic.
Remark 2.6.
If is a Alexandroff space with a minimum (maximum) , then is contractible to . This follows from the previous proposition and the fact that the constant map given by satisfies that () for every .
Definition 2.7.
Given a finite poset , the height of is one less than the maximum number of elements in a chain of . The height of a point in a locally finite Alexandroff space is given by . For a general Alexandroff space , the height of a point is defined as if contains a chain without a minimum and otherwise.
Example 2.8.
Let us consider the real numbers with the usual order. For every the height of is because the chain does not have a minimum. Moreover, since there is no satisfying that or . Let us consider , where we consider the partial order defined as follows: for every . It is clear that is not a finite set but the height of is . Furthermore, .
Proposition 2.9.
Let and be Alexandroff spaces. If is a homeomorphism and , then (). Furthermore, for every the height of is equal to the height of and .
The following results provide a combinatorial way to study the homotopy and weak homotopy type of finite topological spaces.
Definition 2.10.
Let be an Alexandroff space. A point in is a down beat point (resp. up beat point) if has a maximum (resp. has a minimum). A finite topological space is a minimal finite space if it has no beat points. A core of a finite topological space is a strong deformation retract which is a minimal finite space.
Proposition 2.11 ([17]).
Let be an Alexandroff space and let be a beat point. Then is a strong deformation retract of .
If is a finite topological space, then has a core. We only need to remove beat points one by one to obtain a minimal finite space.
Theorem 2.12 ([17]).
If is a minimal finite space, then is a homeomorphism if and only if is a homotopy equivalence.
Corollary 2.13.
If is a minimal finite space, then is isomorphic to .
Remark 2.14.
The result of Corollary 2.13 can be stated in a stronger way. Let be a minimal finite space. It is easy to check that . For every in there is only one element in the class . We can identify every homeomorphism with its homotopy class.
Remark 2.15.
Corollary 2.13 can be generalized to Alexandroff spaces. To do this, consider the notion of being locally a core introduced in [13]. This notion generalizes the notion of minimal finite space, that is, every minimal finite space is locally a core. Let be locally a core. Then a continuous map is a homeomorphism if and only if is a homotopy equivalence. Moreover, we get that .
Definition 2.16.
Let be a finite topological space. A point in is a down weak beat point (resp. up weak beat point) if is contractible (resp. is contractible).
Proposition 2.17 ([3]).
Let be a finite topological space and let be a weak beat point. Then the inclusion is a weak homotopy equivalence.
Definition 2.18 ([3]).
Let be a finite topological space and let . It is said that collapses to by an elementary collapse if is obtained from by removing a weak beat point. Given two finite topological spaces and , collapses to if there is a sequence of finite topological spaces such that for each , collapse to by an elementary collapse.
Lemma 2.19.
Let and be Alexandroff spaces and let be a homeomorphism. Then is a down (up) weak beat point if and only if is a down (up) weak beat point.
Proof.
There is no loss of generality in assuming that is a down weak beat point. If is an up weak beat point, then the argument is similar. It is easy to see that . Therefore, and we get the desired result.
∎
We recall the notion of a Hasse diagram for a locally finite Alexandroff space . The Hasse diagram of is a directed graph. The vertices of are the points of . There is an edge between two vertices and if and only if and the orientation of the edge is from the lower element to the upper element. We omit the orientation of the subsequent Hasse diagrams and we assume an upward orientation.
Remark 2.20.
It is easy to identify beat points of a finite topological space by looking at its Hasse diagram. A vertex is a down beat point (resp. up beat point) if there is only one edge that enters (exits) it, i.e., , where ().
The homotopy and singular homology groups of Alexandroff spaces were studied in [15].
Definition 2.21.
Let be an Alexandroff space. Its McCord complex or order complex is the simplicial complex whose simplices are the non-empty chains of . Let be a simplicial complex. The face poset of , denoted by , is defined to be the poset of simplices of ordered by inclusion.
Remark 2.22.
A finite topological space is said to be collapsible if it collapses to a point. If is a collapsible finite topological space, then is also collapsible.
The geometric realization of a simplicial complex is denoted by .
Theorem 2.23.
[15] Given an Alexandroff space , there exists a weak homotopy equivalence .
Theorem 2.24.
[15] Given a simplicial complex , there exists a weak homotopy equivalence .
Finally, we recall some remarks and a definition. For a more complete treatment we refer to the reader to [3].
Definition 2.25.
The non-Hausdorff join of two Alexandroff spaces and is the disjoint union keeping the given ordering within and and setting for every and .
Remark 2.26.
Given two Alexandroff spaces and , , where denotes the usual join of simplicial complexes. If and are finite topological spaces and one of them is collapsible, then is collapsible.
Proposition 2.27.
Let and be two Alexandroff spaces. Then .
3 Examples and motivation of the proof of Lemma 1.1
We present an example to illustrate the idea of the construction given in the proof of Lemma 1.1.
Example 3.1.
Let us consider the Klein four-group , where we denote and , and the cyclic group of two elements , where we denote and . We also denote and for simplicity. Moreover, let , be generating sets of respectively. We declare . Our goal is to find a finite topological space such that and .
By [5, 7], there exists a finite topological space satisfying that is isomorphic to . In Figure 1 we have represented in blue the Hasse diagram of . It is clear that adding to a minimum , i.e., with for every , we get that is trivial since is contractible. On the other hand, if , then we have that because is a minimum. Thus we deduce that is isomorphic to . Our next goal is to find a topological space satisfying that is trivial and is isomorphic to . Again, by [7], there exists a finite topological space satisfying that . In Figure 1, the Hasse diagram of corresponds to the red and black parts of the diagram on the right. We modify in order to reduce the number of self-homeomorphisms without changing the number of self-homotopy equivalences. For this purpose we add some points to . We consider , where we have the following relations: and . The Hasse diagram of can be seen on the right on Figure 1, where the new points are pictured in orange. It is easy to check that . The new points are beat points so we can remove them without changing the homotopy type of . Hence, we have that and have the same homotopy type. We prove that is trivial. If , then and , where and denote the set of maximal and minimal elements of respectively. From this, using Proposition 2.9 and the fact that preserves heights, we deduce that is the identity.

Combining with we obtain . We identify the point of and . We also extend the partial order of the two previous posets using transitivity, that is, if and , then we have that if and only if . It is not difficult to check that satisfies the properties required at the beginning. and have the same homotopy type because we can collapse to . We thus get that . Since is the only point with height and , it follows that is a fixed point for every . By the continuity of , it is easily seen that and . From this, we conclude that is isomorphic to .

We can also consider what we call the dual case, that is, , where we have that and . We can now proceed analogously to the previous arguments, i.e., we find and . In Figure 3 we have the Hasse diagrams of and .

Again, combining with we get , that is, the finite topological space given by the Hasse diagram of Figure 4. It is trivial to verify that is not homeomorphic to because of their different cardinality. Furthermore, and are not homotopy equivalent since is not isomorphic to . Another way to prove the last assertion is the following. After removing one by one the beat points of we get ; after removing one by one the beat points of we get . However, is not homeomorphic to because of their different cardinality. By Theorem 2.12 we conclude that and are not homotopy equivalent. Moreover, studying their McCord complexes it can be shown that and are not weak homotopy equivalent.

4 Proof of Lemma 1.1 and remarks
Proof of Lemma 1.1.
We follow the same strategy as in Example 3.1. Firstly, we find a topological space such that and is trivial. Secondly, we find a topological space such that and is trivial. Finally, we combine properly both topological spaces to obtain a topological space satisfying that and .
We first assume that and are non-trivial groups. The trivial case will be considered later.
Construction of and properties. We consider a set of non-trivial generators for . Without loss of generality we can consider a well-order on satisfying that if exists and , then there exists satisfying . If exists and there is no satisfying , then the well-order defined on can be modified as follows: for every and the rest of the relations defined on unaltered. It is obvious that the new partial order defined on is indeed a well-order and does not exist. We consider , where we assume that , and extend the well-order defined on to as follows: for every . We consider
where we have the following relations:
-
•
if , where and .
-
•
, where and .
-
•
, where .
The rest of the relations can be deduced from the above relations using transitivity. It is easy to check that is a partially ordered set.
We prove that and is the trivial group. We have that is the trivial group because is contractible to , which is a minimum. Since is a minimum, it follows that every self-homeomorphism must fix this point. From this we deduce that . In addition, is the same topological space considered in [7, Section 3] and denoted by . Hence, we know that , where is given by and is an isomorphism of groups.
Construction of and properties. We consider a set of non-trivial generators for . There is no loss of generality in assuming that there exists a well-order on satisfying that if exists and , then there exists satisfying . We repeat the same construction made before, that is, we consider , where we assume that , and extend the well-order defined on to as follows: for every .
For every we take a well-ordered non-empty set such that is isomorphic to if and only if . For every let denote the first element or minimum of . We consider
where
and we have the following relations:
-
1.
if , where and .
-
2.
, where and .
-
3.
; and , where and .
-
4.
; ; and , where and .
-
5.
, where .
-
6.
We extend the partial order defined on to declaring that , where .
The remaining relations can be deduced from the above using transitivity. It is routine to verify that with the previous relations is a partially ordered set.

We proceed to show that is the trivial group and . It is clear that and have the same homotopy type. We define given by
It is trivial to show that is continuous and satisfies that for every , where denotes the identity map. This implies that is a strong deformation retract of . On the other hand, is the same topological space considered in [7, Section 3] and denoted by . Therefore we know that , where given by and is an isomorphism of groups.
The task is now to prove that is the trivial group. Let us take . We consider for some . Since has a minimum , it follows that is an up beat point. By Lemma 2.19, we know that is also an up beat point. Therefore, is of the form for some . By Proposition 2.9 we get that . It follows from the continuity of that . Since is a homeomorphism, we have that is also a homeomorphism. Therefore we get that ; otherwise we would get a contradiction since is homeomorphic to if and only if . Using Proposition 2.9 it is easy to verify that fixes for every . On the other hand, [7, Remark 4.2] says that if a homeomorphism coincides at one point with the identity map, then is the identity map. Thus, we can deduce that is the identity map and is the trivial group.
Construction of . We consider , where we are identifying the point of both topological spaces, i.e., the partial order of preserves the relations defined on and :
-
•
If and , then is smaller than if and only if and .
-
•
If , then is smaller (greater) than if and only if is smaller (greater) than with the partial order defined on .
-
•
If , then is smaller (greater) than if and only if is smaller (greater) than with the partial order defined on .
It is evident that is isomorphic to because is contractible to and is homotopy equivalent to . It suffices to show that is isomorphic to . We verify that every satisfies that for every and for every . Firstly, we show that is a fixed point for every homeomorphism . We have . Since for every the height of is at least or different from , it follows that . The only elements of that have height one are of the form or or , , for some and . We can discard the maximal elements, otherwise, would send a maximal element to a non-maximal element. If , then we get a contradiction since is an up beat point. If for some , then we get that by Proposition 2.9. By Lemma 2.19 we have that for every because is not a down beat point. Hence, the only possibility is . Finally, by the continuity of , we get that for every . This implies that is isomorphic to .
We prove the remaining case. If is the trivial group, then it suffices to consider to conclude. If is the trivial group, then satisfies the desired properties. ∎
Remark 4.1.
Proposition 4.2.
Let and be groups. Then the Alexandroff space obtained in the proof of Lemma 1.1 has the weak homotopy type of the wedge sum of circles when is a finite group and the wedge sum of circles when is a non-finite countable set.
Proof.
We have that has the same homotopy type of . Repeating the same arguments used in [7, Proposition 6.1] the desired result follows. ∎
Remark 4.3.
Let and be finite groups and let be the finite topological space obtained in the proof of Lemma 1.1. We can remove from . The resulting poset also satisfies that its group of homeomorphisms is isomorphic to and its group of homotopy classes of self-homotopy equivalences is isomorphic to . This finite topological space has points. The first term corresponds to , the second term corresponds to , the third term corresponds to the sets , the fourth term corresponds to the points of the sets and the last term corresponds to the point .
We can change the sets from the proof of Lemma 1.1 by as in [7, Section 5], where with the following relations:
(1) | |||
(2) |
We consider (1) for odd and (2) for even. We denote this poset by .
Corollary 4.4.
Given two finite groups and , there are infinitely many (non-homotopy-equivalent) topological spaces such that is isomorphic to and is isomorphic to for every .
5 Examples, remarks and proof of Theorem 1.2
The idea of this section is to modify the topological space obtained in Lemma 1.1 to prove Theorem 1.2. Given a homomorphism of groups , we slightly modify the topological space defined in the proof of Lemma 1.1 to get . Adding new relations to we can control the homomorphism of groups given by .
Example 5.1.
Let us consider the cyclic group of two elements and the group of integer numbers . We consider the homomorphism of groups given by . We consider the topological space obtained in the proof of Lemma 1.1. We remove and from it. The resulting poset corresponds to the Hasse diagram shown in black in Figure 6.

Now we add the following relations to : if and if . In Figure 6 we have represented these relations in blue. It is easy to verify that satisfies that , and .
Example 5.2.
Let be the homomorphism of groups given by and . In Figure 7 we have the Hasse diagram of , where we use the same notation introduced in Example 3.1. We have that , and .

Proof of Theorem 1.2.
Suppose and are not trivial groups since otherwise the result follows from Lemma 1.1. Let be the topological space obtained in the proof of Lemma 1.1. We consider . We keep the same relations defined on as a subspace of and add the following relations:
-
•
if , where and .
-
•
if .
It is easy to check that is a partially ordered set with the above relations. The task is now to show that . We consider given by
We have that preserves the order so it is a continuous map. It is simple to verify that for every , where denotes the identity map. From this it follows that is homotopy equivalent to . On the other hand, repeating the same arguments used in [7], it can be proved that is locally a core [13] or a minimal finite space in case is a finite group, which implies that . Since and , it follows that every homeomorphism fixes . Hence, the group of homeomorphisms of as a subspace of is isomorphic to the group of homeomorphisms of the topological space obtained in the proof of Lemma 1.1. Thus, .
We proceed to show that . We consider the following auxiliary sets: , where , and , where . If , then every homeomorphism satisfies that . We prove the last assertion. We know that does not contain beat points. On the other hand, for every we have that is a beat point of height . Using Proposition 2.9 and the notion of continuity we deduce that for every and , where and , we have for some .
We consider given by if and , if and , defined in the natural way if and and . We prove that is well-defined. We verify the continuity of . Suppose for some and . By hypothesis, . Therefore,
It is easy to check that preserves the remaining relations. The inverse of is given by . Hence, is well-defined. By construction, is a monomorphism of groups. Suppose . Proposition 2.9, Remark 4.1 and the fact that imply that every satisfies and for some and , where and . We consider for some and . We get for some and for some . By Remark 4.1, the proof of Lemma 1.1 and the fact that , there exists such that , where . Hence, . By Proposition 2.9, , we have . Thus, because of Remark 4.1 and the fact that . By construction, for every the equality holds. Since every can be seen as for some , it follows that , where . ∎
Remark 5.3.
Proposition 5.4.
Let and be groups. If are homomorphisms of groups, then is homotopy equivalent to .
Proof.
The result is an immediate consequence of the construction. We have that is homotopy equivalent to for every homomorphism of groups . Therefore, the homotopy type of the topological space obtained in the proof of Theorem 1.2 does not depend on the homomorphism chosen to construct it. Thus we deduce the desired result. ∎
Proposition 5.5.
Let and be groups and let be homomorphisms of groups. Then if and only if is homeomorphic to .
Proof.
One of the implications is trivial. It suffices to show that if is homeomorphic to , then . Since is homeomorphic to , it follows that there exists a homeomorphism . From the construction of and in the proof of Theorem 1.2 it can be easily deduced that and . This is due to the fact that contains beat points while does not have beat points. Therefore, can be related to the action of an element and can be related to the action of an element . We have , where denotes the identity element in , and we also have
which implies that . Thus, because is a homomorphism of groups. In addition, for every , we know that there exists a relation in of the following form . We have
By the construction of we get . Earlier we prove that , which implies that for every . ∎
6 Groups of homology, homotopy and automorphisms
We first prove that the groups studied previously do not determine neither the homotopy type nor the topological type of a topological space in general. To do this we provide an example. However, if the topological space satisfies some properties, namely, is compact and a locally Euclidean manifold with or without boundary, then its group of homeomorphisms determines its topological type of it, see [19] for more details.
Example 6.1.
Let us consider the Alexandroff space given by the Hasse diagram of Figure 8. It is the union of and , where we are identifying the point of for . For simplicity, denotes . The topological space was introduced in [16, Figure 2] and has the weak homotopy type of a point. It is proved in [8] that is the smallest finite topological space having the same weak homotopy type of a point but not contractible. It is clear that does not have beat points, so is isomorphic to . On the other hand, it is easy to show that is the trivial group. Since only for and the heights of these points are different, it follows that every homeomorphism must fix and . Proposition 2.9 leads to the desired result. Furthermore, the homotopy and singular homology groups of are trivial. We can study the weak homotopy type of studying the McCord complex or removing beat and weak beat points. We have that is a weak beat point. After removing this point, and are up beat points. If we remove them, then it is easy to check that the remaining space is homotopy equivalent to the space given by the points . We continue in this fashion. We have is a weak beat point. After removing this point, and are down beat points. Thus, has the same weak homotopy type of a point. Therefore we have that is a topological space satisfying that for every but it is not homeomorphic nor homotopy equivalent to a point. We can generalize this topological space by taking more copies of the topological space introduced in [16]. For instance, we can define just as , where and we are identifying the point of and . It is easy to prove that has the weak homotopy type of a point for every because with is a weak beat point.

One possible consequence of the previous construction is the following result.
Proposition 6.2.
Let and be finite groups. There exists a topological space such that is isomorphic to , is isomorphic to and is weak homotopy equivalent to a point.
Proof.
We consider the topological space obtained in the proof of Lemma 1.1 and the finite topological space given in Example 6.1. Let denote . By Proposition 2.27, Example 6.1 and the proof of Lemma 1.1 we have that . We get that is homotopy equivalent to by removing its beat points one by one. Since does not contain beat points by Corollary 2.13, it follows that . In addition, since is collapsible, it follows that has the weak homotopy type of a point by Remark 2.26 ∎
We motivate the proof of Theorem 1.3 with one example.
Example 6.3.
Let us consider and . We consider the minimal finite model of the 2-dimensional sphere , that is, , where and . Then is homeomorphic to .
We want to find a finite topological space such that and are isomorphic to and respectively for every non-negative integer , is isomorphic to and is isomorphic to . The idea is to modify to obtain a new space satisfying that its group of homeomorphisms is trivial. We enumerate the points of . For each with , we add to as in Example 6.1. The Hasse diagram of the new topological space, denoted by , can be seen in Figure 9. The Hasse diagram of is painted black whereas the new part is blue and purple. In purple we have the weak beat points that are not beat points. It is clear that does not have beat points so is isomorphic to . A homeomorphism sends weak beat points to weak beat points by Lemma 2.19. It is easy to deduce from this that is the trivial group. On the other hand, the new structure added can be removed without changing the weak homotopy type of the space, see Example 6.1. Therefore, we have that is isomorphic to and is isomorphic to for every non-negative integer .
Finally, we add a new point that connects to , where is the space obtained the proof of Lemma 1.1 for finite groups. In Figure 9 we have the Hasse diagram of the new topological space . The relations with the point are shown in green. The Hasse diagram of is painted read and orange. It is easy to check that satisfies the desired properties.

Proof of Theorem 1.3.
If has the same homotopy type of a point, then the result can be deduced from Proposition 6.2. Therefore we can assume that does not have the same homotopy type of a point. The idea of the proof is to follow techniques similar to the ones used in the proof of Lemma 1.1. From the simplicial approximation to CW-complexes, [11, Theorem 2C.5.] we get that there exists a finite simplicial complex that is homotopy equivalent to . By abuse of notation we continue to write for the finite simplicial complex. We apply the McCord functor to in order to obtain a finite topological space such that is weak homotopy equivalent to . We can suppose that does not have beat points or weak beat points; otherwise we can remove them one by one until there are none left. We denote by and label the points in , that is, . For each we consider , where is the topological space obtained in Example 6.1. We consider , where we are identifying the point with for every . We define the partial order on extending the already existing partial orders. To do that we use transitivity, i.e., for , if and only if one of the following situations is satisfied:
-
•
and is greater than with the partial order defined on .
-
•
for some and is greater than with the partial order defined on .
-
•
, for some and .
Consider , where is the space obtained in the proof of Lemma 1.1 and the space given in Example 6.1. We extend the partial order defined on and to by declaring that for some , where is a minimal point in and . We prove that restricted to is the identity. In there are weak beat points that we will denote by with . In fact, we have that the only weak beat points that are not beat points or do not have a bigger beat point are in . Hence, if is a weak beat point, we have for some . By Proposition 2.9 and Lemma 2.19, we have that . But is homeomorphic to if and only if . By the continuity of , , so , as we wanted. It is easy to check that is also a fixed point for every homeomorphism since and . We get . By Proposition 2.27, Example 6.1 and the Proof of Lemma 1.1, . On the other hand, , but does not contain beat points. Therefore, by Corollary 2.13, . From here, repeating similar arguments than the ones used before, it can be deduced that .
Finally, is clearly the wedge sum of and . From Remark 2.26 we obtain that is homotopy equivalent to a point since is collapsible, which implies that is also collapsible, and is homotopy equivalent to . We also get that is homotopy equivalent to because every can be removed following the steps of Example 6.1 without changing the weak homotopy type of . Therefore, for every , we have and . ∎
Proof of Corollary 1.4.
Proof of Corollary 1.5.
We only need to use the beginning of the construction of Eilenberg-Maclane spaces to obtain a compact CW-complex with and possibly non-trivial higher homotopy groups. Therefore, the result is an immediate consequence of Theorem 1.3. ∎
Remark 6.4.
For a compact -complex , and are nilpotent groups, see for instance [10, Section 4]. () can be seen as the kernel of a homomorphism of groups. We consider the functor between and the category of groups given by (). It is easy to check that induces a homomorphism of groups, . This sends each self-homotopy equivalence to its induced morphism in the homotopy groups (homology groups), where denotes here the group of automorphisms of a group in the category of groups. Then, () can be seen as the kernel of and so it is a normal subgroup of . With the following example we prove that for a general topological space we cannot expect the same result.
Example 6.5.
Applying the construction obtained in the proof of Theorem 1.3 we can get a topological space such that is trivial, and is weak homotopy equivalent to a circle, where denotes the symmetric group on a set of elements. In Figure 10 we present the Hasse diagram of . By construction, every self-homotopy equivalence of fixes the blue, black, green and orange parts of the Hasse diagram. On the other hand, the only part that contributes to the homotopy groups or homology groups is the black one. Therefore we can deduce that , which implies that and are not nilpotent groups.

Remark 6.6.
Acknowledgment. We thank Jesús Antonio Álvarez López for posing the question that led to one of the main results of this paper. We wish to express our gratitude to Jonathan A. Barmak for his useful comments and suggestions that substantially improved this manuscript. The authors gratefully acknowledges the many helpful corrections of Andrei Martínez Finkelshtein, Elena Castilla and Jaime J. Sánchez-Gabites during the revision of the paper.
References
- [1] P. S. Alexandroff. Diskrete Räume. Mathematiceskii Sbornik (N.S.), 2(3):501–519, 1937.
- [2] L. Babai. Finite digraphs with given regular automorphism groups. Period. Math. Hung., 1:257,270, 1980.
- [3] J. A. Barmak. Algebraic topology of finite topological spaces and applications, volume 2032. Springer, 2011.
- [4] J. A. Barmak. Automorphism groups of finite posets II. Preprint. arXiv:2008.04997, 2020.
- [5] J. A. Barmak and E. G. Minian. Automorphism groups of finite posets. Discrete Math., 309(10):3424–3426, 2009.
- [6] G. Birkhoff. On groups of automorphisms. Rev. Un. Mat. Argentina, 11:155–157, 1946.
- [7] P. J. Chocano, M. A. Morón, and F. R. Ruiz del Portal. Topological realizations of groups in alexandroff spaces. Rev. R. Acad. Cien. Serie A. Mat., DOI: 10.1007/s13398-020-00964-7, 2021.
- [8] N. Cianci and M. Ottina. Smallest weakly contractible non-contractible topological spaces. Proc. Edinburgh Math. Soc., 63(1):263–274, 2020.
- [9] C. Costoya and A. Viruel. Every finite group is the group of self-homotopy equivalences of an elliptic space. Acta Math., 213(1):49–62, 2014.
- [10] C. Costoya and A. Viruel. A primer on the group of self-homotopy equivalences: a Rational Homotopy Theory approach. Graduate J. Math., 5(1):76..87, 2020.
- [11] A. Hatcher. Algebraic topology. Cambridge Univ. Press, Cambridge, 2000.
- [12] D. W. Kahn. Realization problems for the group of homotopy classes of self-equivalences. Math. Ann., 220(1):37–46, 1976.
- [13] M. J. Kukieła. On homotopy types of Alexandroff spaces. Order, 27(1):9–21, 2010.
- [14] J. P. May. Finite spaces and larger contexts. Unpublished book, 2016.
- [15] M. C. McCord. Singular homology groups and homotopy groups of finite topological spaces. Duke Math. J., 33(3):465–474, 1966.
- [16] I. Rival. A fixed point theorem for finite partially ordered sets. J. Combin. Theory A 21, pages 309–318, 1976.
- [17] R. E. Stong. Finite topological spaces. Trans. Amer. Math. Soc., 123(2):325–340, 1966.
- [18] M. C. Thornton. Spaces with given homeomorphism groups. Proc. Amer. Math. Soc., 33(1):127–131, 1972.
- [19] J. V. Whittaker. On Isomorphic Groups and Homeomorphic Spaces. Ann. of Math., 78(1):74–91, 1963.
P.J. Chocano, Departamento de Álgebra, Geometría y Topología, Universidad Complutense de Madrid, Plaza de Ciencias 3, 28040 Madrid, Spain
E-mail address:pedrocho@ucm.es
M. A. Morón, Departamento de Álgebra, Geometría y Topología, Universidad Complutense de Madrid and Instituto de Matematica Interdisciplinar, Plaza de Ciencias 3, 28040 Madrid, Spain
E-mail address: ma_moron@mat.ucm.es
F. R. Ruiz del Portal, Departamento de Álgebra, Geometría y Topología, Universidad Complutense de Madrid and Instituto de Matematica Interdisciplinar , Plaza de Ciencias 3, 28040 Madrid, Spain
E-mail address: R_Portal@mat.ucm.es