Bounded cohomology and binate groups
Abstract.
A group is boundedly acyclic if its bounded cohomology with trivial real coefficients vanishes in all positive degrees. Amenable groups are boundedly acyclic, while the first non-amenable examples were the group of compactly supported homeomorphisms of (Matsumoto–Morita) and mitotic groups (Löh). We prove that binate (alias pseudo-mitotic) groups are boundedly acyclic, which provides a unifying approach to the aforementioned results. Moreover, we show that binate groups are universally boundedly acyclic.
We obtain several new examples of boundedly acyclic groups as well as computations of the bounded cohomology of certain groups acting on the circle. In particular, we discuss how these results suggest that the bounded cohomology of the Thompson groups , , and is as simple as possible.
Key words and phrases:
bounded cohomology, boundedly acyclic groups, binate groups, pseudo-mitotic groups, Thompson groups2020 Mathematics Subject Classification:
Primary: 18G901. Introduction
Bounded cohomology is defined via the topological dual of the simplicial resolution. It was introduced by Johnson and Trauber in the context of Banach algebras [35], then extended by Gromov to topological spaces [28]. Since then it has become a fundamental tool in several fields, including the geometry of manifolds [28], rigidity theory [13], the dynamics of circle actions [25], and stable commutator length [14].
Despite a good understanding in degree and a partial understanding in degree , the full bounded cohomology of a group seems to be hard to compute [22, Section 7]. Therefore, it is fundamental to produce alternative resolutions that compute the bounded cohomology of a group. In this respect, amenable groups played a fundamental role in the approach of Ivanov [32]. One can also exploit a larger class of groups for computing bounded cohomology, namely the class of boundedly acyclic groups [48, 33]:
Definition 1.1.
Let . A group is said to be -boundedly acyclic if for all . The group is boundedly acyclic if it is -boundedly acyclic for every .
Amenable groups are boundedly acyclic [35, 28]. The first non-amenable example, due to Matsumoto and Morita [42] is the group of compactly supported homeomorphisms of . Their proof relies on the acyclicity of this group, which for the purposes of this paper will always be intended with respect to the integers:
Definition 1.2.
A group is said to be acyclic if for all .
It was shown by Mather that is acyclic [40], and thus the proof of bounded acyclicity reduces to the proof of injectivity of the comparison map from bounded to ordinary cohomology. The same approach was employed by Löh to prove that mitotic groups are boundedly acyclic [39]. This class was introduced by Baumslag, Dyer, and Heller [1] to produce embedding results into finitely generated acyclic groups. Bounded acyclicity of mitotic groups, together with co-amenability of ascending HNN extensions, eventually led to finitely generated and finitely presented examples of non-amenable boundedly acyclic groups [22].
The two bounded acyclicity results mentioned above are similar in spirit but independent of one another, since is not mitotic [49]. On the other hand, there is a larger framework that includes both and mitotic groups: binate groups (see Section 3 for the definition). This class was introduced by Berrick [4] and independently by Varadarajan [55], under the name pseudo-mitotic. They proved that binate groups are acyclic, thus providing a unified approach to the proofs of Mather and Baumslag–Dyer–Heller, as well as several new and interesting examples of acyclic groups, mainly among groups of homeomorphisms (see Section 3.1). We adapt this unification to bounded cohomology:
Theorem 1.3 (Theorem 3.5).
All binate groups are boundedly acyclic.
We remark that in general binate groups are non-amenable, since they typically contain free subgroups. However, there are a few exceptions (Section 3.1.4).
Binate groups reflect enough of group theory to serve as a faithful testing class for open problems such as the Bass Conjecture, a modified version of the Baum–Connes Conjecture or the Kervaire Conjecture [6]. By Theorem 1.3, boundedly acyclic groups also serve as conjecture testers. For instance, if the Bass Conjecture holds for all boundedly acyclic groups, then it holds for all groups. This is especially interesting since amenable groups, which serve as the prototypical example of boundedly acyclic groups, are known to satisfy the Bass Conjecture [7].
The bounded acyclicity of binate groups is a phenomenon that is not strictly linked to real coefficients. Indeed we prove:
Theorem 1.4 (Corollary 5.4).
Binate groups are universally boundedly acyclic: If is a binate group, then for every complete valued field and every , we have .
More generally, we characterize universally boundedly acyclic groups as those groups that are simultaneously acyclic and boundedly acyclic (Theorem 5.2). In this sense, Theorem 1.4 is a combination of the acyclicity result of Berrick and Varadarajan, together with Theorem 1.3, but it also contains both results in its statement.
Hereditary properties of boundedly acyclic groups
By analogy with the amenable case, it is interesting to check which group-theoretic constructions preserve bounded acyclicity. It is known that extensions, as well as quotients with boundedly acyclic kernels, do [48]. It is therefore natural to wonder whether these two results extend to a -out-of- property for bounded acyclicity and group extensions.
Using the fact that every group embeds -step subnormally into a binate group, we show that this cannot hold. More precisely:
Theorem 1.5 (Theorem 4.5).
There exists a boundedly acyclic group with a normal subgroup such that is boundedly acyclic, but is continuum-dimensional for every .
Application to Thompson groups
The advantage of Theorem 1.3 is that the class of binate groups is flexible enough that one can construct several concrete examples of boundedly acyclic groups. We use this to study the bounded cohomology of certain analogues of the classical Thompson groups , and . The amenability question for is one of the most influential open questions in modern group theory. It is therefore natural to wonder whether is at least boundedly acyclic. It is known that is -boundedly acyclic, but nothing seems to be known in higher degrees. Using Theorem 1.3, we show that a countably singular analogue of the Thompson group is boundedly acyclic (Proposition 6.7).
Moreover, we prove that if is -boundedly acyclic, then the bounded cohomology of is generated by the real Euler class and its cup powers, up to degree (Corollary 6.12). In particular, we obtain:
Theorem 1.6 (Corollary 6.12/6.13).
If the Thompson group is boundedly acyclic, then (with the cup-product structure) is isomorphic to the polynomial ring with and the bounded Euler class of is a polynomial generator of . Moreover, the canonical semi-norm on then is a norm.
Therefore, the bounded acyclicity of would make into the first group of type that is not boundedly acyclic, and whose bounded cohomology ring can be completely and explicitly be computed. Similarly, if is boundedly acyclic, then the bounded cohomology ring of is a polynomial ring in generators of degree (Corollary 6.13).
Independently, Monod and Nariman recently established analogous results for the bounded cohomology of the group of orientation-preserving homeomorphisms of [46].
A note from the future
Organisation of this article
We recall the definition of bounded cohomology and the uniform boundary condition in Section 2. Binate groups are surveyed in Section 3. In Section 4, we study hereditary properties of boundedly acyclic groups. Section 5 is devoted to universal bounded acyclicity. The applications to Thompson groups are discussed in Section 6. Finally, Appendix A contains the proof of Theorem 1.3.
Acknowledgements
We wish to thank Jon Berrick, Jonathan Bowden, Amir Khodayan Karim, Kevin Li, Yash Lodha, Antonio López Neumann, Nicolas Monod, Sam Nariman and George Raptis for helpful discussions.
2. Bounded cohomology
We quickly recall basic notions concerning bounded cohomology.
2.1. Definition of bounded cohomology
Let be a group and let be the bounded simplicial -resolution of . More generally, if is a normed -module, we consider the complex and set
The bounded cohomology of with coefficients in is defined as
The norm on induces a semi-norm on , the so-called canonical semi-norm.
The canonical inclusion induces a natural transformation between bounded cohomology and ordinary cohomology, the comparison map
Further information on the bounded cohomology of groups (and spaces) can be found in the literature [28, 32, 24].
In Section 5, we will also be dealing with bounded cohomology over different valued fields. Recall that an absolute value on a field is a multiplicative map such that if and only if ; and the triangle inequality holds: . We say that is a complete valued field if the metric induced by the absolute value is complete. One can then define bounded cohomology over in exactly the same way.
If the strong triangle inequality holds for all , then is said to be non-Archimedean. Concerning the bounded cohomology over non-Archimedean fields [20], we will only use the following result:
Lemma 2.1 ([20, Corollary 9.38]).
Let be a group, let , and suppose that is finitely generated. Then the comparison map
is injective.
2.2. The uniform boundary condition
We recall the uniform boundary condition, originally due to Matsumoto and Morita [42], and some of its variations [39].
Definition 2.2 (Uniform boundary condition).
Let and let . A group satisfies the -uniform boundary condition, or simply -, if for every there exists a chain with
A group satisfies - if it has - for some .
The uniform boundary condition can lead to bounded acyclicity:
Theorem 2.3 ([42, Theorem 2.8]).
Let be a group and let . Then, the following are equivalent:
-
(1)
The group satisfies -;
-
(2)
The comparison map is injective.
In particular: Every acyclic group that satisfies in all positive degrees is boundedly acyclic.
In the proof of Theorem 1.3, it will be useful to extend the definition of from groups to group homomorphisms [39, Definition 4.5].
Definition 2.4 ( for homomorphisms).
Let and let . A group homomorphism satisfies the -uniform boundary condition, or simply -, if there exits a linear map
with
Here is the operator norm of with respect to the -norms.
The uniform boundary condition will be more systematically reviewed in a forthcoming paper [38].
3. Binate groups (alias pseudo-mitotic groups)
We recall basic notions, properties, and examples of binate (alias pseudo-mitotic) groups. We begin with the original definition given by Berrick [4]:
Definition 3.1 (Binate).
Let be a group. We say that is binate if for every finitely generated subgroup there exists a homomorphism and an element such that for every , we have
We will rather work with the equivalent notion of pseudo-mitotic groups, introduced by Varadajan [55]: Here an extra homomorphism is taken as part of the structure, which leads to more transparent proofs. We refer the reader to the literature [8, Remark 2.3] for a proof of the equivalence, and point out that the terminology binate is more commonly used.
Definition 3.2 (Pseudo-mitosis).
Let be a a group and let be a subgroup. We say that has a pseudo-mitosis in if there exist homomorphisms , and an element such that
-
(1)
For every , we have ;
-
(2)
For all , we have ;
-
(3)
For every , we have .
Here is what the definition intuitively means. There exists a homomorphism whose image commutes with . This induces a homomorphism
Precomposing it with the diagonal inclusion , we get a homomorphism . In terms of acyclicity the crucial condition is the third item: and are conjugate inside .
Definition 3.3 (Pseudo-mitotic group).
A group is said to be pseudo-mitotic if all finitely generated subgroups of admit a pseudo-mitosis in .
Theorem 3.4 ([55, Theorem 1.7]).
All pseudo-mitotic groups are acyclic.
In the present article, we prove that pseudo-mitotic groups are also examples of boundedly acyclic groups (Theorem 1.3):
Theorem 3.5.
All pseudo-mitotic groups are boundedly acyclic.
Since the proof is rather technical and it closely follows the ones of Matsumoto–Morita [42] and Löh [39], we postpone it to Appendix A.
Remark 3.6.
It is an easy consequence of Theorem 3.4 that pseudo-mitotic groups are -boundedly acyclic, namely that if is a pseudo-mitotic group, then . Indeed, if is a pseudo-mitotic group and , by definition there exist homomorphisms and an element such that
In fact, this commutator expression is the one appearing in the definition of binate groups (Definition 3.1). Hence, every element in a pseudo-mitotic group is a commutator and so the second comparison map is injective [2]. This shows that embeds into , which vanishes by Theorem 3.4.
3.1. Examples
We present several examples of pseudo-mitotic groups. A more detailed discussion of these examples can be found in Berrick’s work [5].
We start with a combinatorial construction of pseudo-mitotic groups containing a given group.
Example 3.7 (Binate tower).
Let be a group. Set , and construct inductively by performing an HNN-extension of so that the embedding is a pseudo-mitosis. More precisely, if
then is a pseudo-mitotic embedding of in .
By construction, the direct limit of the is pseudo-mitotic. It is the initial object in a category of pseudo-mitotic groups containing [4].
This example shows that every group embeds into a pseudo-mitotic group. We will see in the next section that a less canonical construction leads to embeddings with more special properties (Proposition 4.4).
The following allows to construct new binate groups from old ones:
Example 3.8.
Let be a family of binate groups. Then their direct product is binate [49, Proposition 1.7].
We will soon see that is binate. Therefore the previous example shows that is binate, whence boundedly acyclic. A direct proof of bounded acyclicity is given by Monod and Nariman [46].
For comparison, note that an arbitrary direct product of amenable groups need not be amenable. For instance, if is a non-amenable residually finite group, such as a non-abelian free group, then embeds into the direct product of its finite quotients, which is therefore not amenable.
3.1.1. Dissipated groups
Let us move to more concrete examples. Varadarajan proved that the group of compactly supported homeomorphisms of is pseudo-mitotic [55, Theorem 2.2]. Following Berrick [5], we show here that this is just an instance of the behaviour of a larger class of groups: Dissipated boundedly supported transformation groups.
Definition 3.9 (Boundedly supported group).
Let be a group acting faithfully on a set , which is expressed as a directed union of subsets . For each , let . We say that is boundedly supported if is the directed union of the .
The key property that makes certain boundedly supported groups pseudo-mitotic is the following:
Definition 3.10 (Dissipators).
Let and be as in Definition 3.9. Let . A dissipator for is an element such that
-
(1)
for all .
-
(2)
For all , the bijection of defined by
() is in .
If for each there exists a dissipator for , we say that is dissipated.
In order for to be a dissipator, the element needs to belong to , and the boundedly supported hypothesis implies that there exists such that for all . Figure 1 illustrates this situation.
left: the subsets ; right: the action of .
The presence of dissipators is enough to ensure that the group is pseudo-mitotic:
Proposition 3.11 ([5, Section 3.1.6]).
Dissipated groups are pseudo-mitotic.
Proof.
Let and be as in the definition of a dissipated group (Definition 3.10). Let be a finitely generated subgroup. Since is boundedly supported, there exists an such that . Notice that commutes with (as defined in Equation ( ‣ 2)) since their supports are disjoint in . Hence, if we define as , it is immediate to check that is a homomorphism and that for all . We then set and . By construction, this implies that for all . Hence, and are the witnesses of a pseudo-mitosis of in . ∎
A more topological version of this criterion is described by Sankaran and Varadarajan [50, Theorem 1.5]. Many boundedly supported groups are dissipated, and quite surprisingly this is usually easy to check. We list some examples for which dissipators can be computed directly. More details and further constructions can be found in Berrick’s paper [5, Section 3.1.6] and the references therein, as well as in the one of Sankaran–Varadarajan [50].
Example 3.12 (Dissipated groups).
The following groups are dissipated:
-
(1)
The group of compactly supported homeomorphisms of is dissipated. This is already contained in a paper of Schreier and Ulam [52], where they study this phenomenon for the (isomorphic) group of homeomorphisms of the -ball in fixing a neighbourhood of the boundary. Acyclicity was shown by Mather [40], and the proof serves as a model for the proof of acyclicity of pseudo-mitotic groups [55].
- (2)
-
(3)
Let be the standard Cantor set, embedded in . Then the group of homeomorphisms of that are the identity in a neighbourhood of and is dissipated [50, Theorem 2.4].
-
(4)
Let be endowed with the topology as subspace of . Then, the group of homeomorphisms of having support contained in some interval with is dissipated. The same holds for the space of irrational numbers [50, Theorem 1.13].
-
(5)
Forgetting the topology, denote by the group of bijections of whose support is contained in some interval with . Then is dissipated [49, Theorem 3.2]. The same holds for groups of bijections of infinite sets with similar properties.
3.1.2. Flabby groups
Another source of examples are flabby groups:
Definition 3.13 (Flabby group).
A group is flabby if there exist homomorphisms and such that the following holds: For every finitely generated subgroup there exist such that for all :
-
(1)
;
-
(2)
;
-
(3)
.
Flabby groups were defined by Wagoner [56], who proved that they are acyclic. In fact, the following stronger result is true:
Lemma 3.14 ([4, Section 3.3]).
Flabby groups are pseudo-mitotic.
Proof.
Let be a flabby group and a finitely generated subgroup. Let be as in the definition of flabby group. We define . Then, since commutes with , we have for all . Let for every . Then:
for every . By setting , we get the thesis. ∎
The definition of a flabby group is more restrictive than that of a pseudo-mitotic group, since the homomorphisms and impose some uniformity in the choices of the homomorphisms and . Still, there are several examples of flabby groups in the literature.
Example 3.15 (Flabby groups).
The following groups are flabby:
-
(1)
To study the algebraic theory of a ring , Wagoner [56] embeds into another ring called the cone over . Then the direct limit general linear group is shown to be flabby, whence acyclic.
-
(2)
Building on the work of Wagoner, several other examples of flabby groups are exhibited by de la Harpe and McDuff [17], and they all have the following flavour. Let be an (infinite-dimensional) Hilbert space, and let
be a chain of closed subspaces such that is isomorphic to for all . Let be the group of continuous linear isomorphisms of , and let . Then the direct limit of the ’s is flabby. The same holds if one restricts to unitary operators.
-
(3)
Certain groups of automorphisms of measure spaces fall into the framework of the previous item, and thus are flabby [17].
3.1.3. Mitotic groups
Pseudo-mitotic groups were introduced by Varadarajan [55] as a generalization of a more restricted class, that of mitotic groups, introduced by Baumslag, Dyer and Heller [1]. Let us recall the definition:
Definition 3.16 (Mitosis).
Let be a group and let be a subgroup. We say that has a mitosis in if there exist elements such that
-
(1’)
For every , we have ;
-
(2’)
For all , we have .
Definition 3.17 (Mitotic).
A group is said to be mitotic if all finitely generated subgroups of admit a mitosis in .
The main examples of mitotic groups are algebraically closed groups [1, Theorem 4.3]; moreover, a functorial embedding analogous to Example 3.7 is also possible for mitotic groups.
If has a mitosis in , then also has a pseudo-mitosis in : Indeed, we can choose to be conjugation by , to be conjugation by , and . Therefore, every mitotic group is pseudo-mitotic. On the other hand, the class of pseudo-mitotic groups is strictly larger as proved by Sankaran and Varadarajan [49]. Since Theorem 1.3 generalizes the bounded acyclicity of mitotic groups [39] to the class of pseudo-mitotic groups, let us show an explicit example of a group that is pseudo-mitotic but not mitotic:
Lemma 3.18.
The group is not mitotic.
Proof.
Let be a finitely generated group acting minimally on : For concreteness, one could take to be the Thompson group . We embed inside by letting act trivially on . Let be an element such that commutes with . Note that is supported on and also acts minimally. So in order to commute, must be disjoint from .
It follows that the diagonal group is supported on the disconnected set , and therefore cannot be conjugate to by an element in . This shows that is not mitotic. ∎
3.1.4. Amenable examples
Most of the examples that we have seen so far are non-amenable, since they contain free subgroups. In particular, their bounded acyclicity does not follow from the classical result for amenable groups. However, there are some exceptions. The first one is due to Berrick:
Example 3.19.
Locally finite groups cannot be dissipated, since by construction the dissipators must have infinite order. In the next example, we construct a dissipated amenable group.
Example 3.20.
We start with , with the action given by left translation. Of course, is amenable.
Next, let be the disjoint union of countably many copies of indexed by , which contains a distinguished copy of , indexed by . The direct product acts on coordinate-wise. Let be the bijection of shifting the copies of . We set to be the group generated by the direct product and . Note that splits as a semidirect product , so it is -step-solvable. Moreover, given , the element from Definition 3.10 is just with for and otherwise.
By induction, if and have been constructed, we construct as the disjoint union of -many copies of , we let be the shift, and define as the group generated by the direct product of the and . Then is -step-solvable. Moreover, given , the element again belongs to .
The directed union of the acts on the directed union of the . This action is boundedly supported by definition, and the are dissipators by construction. Finally, is a directed union of solvable groups, so it is amenable.
Remark 3.21.
In the construction, one cannot pick any amenable group , since a direct power of amenable groups need not be amenable in general. Indeed, we strongly use the fact that a direct power of an -step-solvable group is still -step-solvable. By the same argument, we could have started with any group satisfying an amenable law.
4. Hereditary properties of boundedly acyclic groups
In this section, we discuss the stability of bounded acyclicity under certain operations. We will present new results concerning normal subgroups and directed unions. The case of normal subgroups will make use of pseudo-mitotic groups, showcasing their versatility compared to mitotic groups.
4.1. Extensions
We start with the operation of taking extensions of boundedly acyclic groups. This behaves particularly well:
Theorem 4.1 ([48, Corollary 4.2.2]).
Let be an exact sequence of groups, where denotes the quotient map, and let . Suppose that is -boundedly acyclic. Then is -boundedly acyclic if and only if is -boundedly acyclic.
In particular, the class of -boundedly acyclic groups is closed under extensions. This generalizes the classical results for extensions with amenable kernel [28] and amenable quotient [35, 47].
A natural question is then whether a -out-of- property holds. Namely:
Question 4.2.
In an extension , suppose that and are -boundedly acyclic. Is necessarily -boundedly acyclic?
4.2. Normal subgroups
Pseudo-mitotic groups are varied enough that they allow for several strong embedding constructions. The following is the most relevant one:
Example 4.3 (Cone over a group).
If is a group, let be the group of functions from to that map all numbers outside some finite interval to the neutral element. The group , where (defined in Example 3.12) acts on by shifting the coordinates, is pseudo-mitotic, in fact dissipated [4, Section 3.5]. The group is called the cone over , and was introduced by Kan and Thurston [36] as a key step in the proof of their celebrated theorem.
Proposition 4.4 ([4, Section 3.5]).
Every group embeds -step subnormally in a pseudo-mitotic group. More precisely, for every group there exists a group such that is pseudo-mitotic.
Proof.
By Example 4.3, it suffices to show that , for some group . The group embeds normally in as the subgroup of functions that map every non-zero rational to the identity. We then set to be the subgroup of functions that map to . ∎
In particular, every group embeds -step subnormally in a boundedly acyclic group. Embeddings into boundedly acyclic groups have been considered before [39, 22], but Proposition 4.4 goes one step further, and provides a strong negative answer to Question 4.2:
Theorem 4.5.
There exists a boundedly acyclic group with a normal subgroup such that is boundedly acyclic, but is continuum-dimensional for every .
Groups such as above are said to have large bounded cohomology: Countable [39] and even finitely generated [22] examples are known to exist.
Proof.
Even without the additional hypothesis on the quotient, it seems that Theorem 4.5 gives the first example of a non-boundedly acyclic normal subgroup of a boundedly acyclic group. Indeed, subgroups of amenable groups are amenable, and several of the non-amenable examples of boundedly acyclic groups available in the literature are simple, so they cannot provide counterexamples. For instance, is simple, as are many other groups of boundedly supported homeomorphisms [18].
4.3. Quotients
An intriguing open problem is whether boundedly acyclic groups are closed under passage to quotients [22, Section 3.2]. One of the main difficulties about this problem is that mitotic groups behave extremely well with respect to quotients:
Lemma 4.6 ([1, Page 16]).
Mitotic groups are closed under passage to quotients.
On the other hand, the same behaviour does not hold for pseudo-mitotic groups:
Lemma 4.7 ([49, Theorem 3.3]).
Pseudo-mitotic groups are not closed under passage to quotients.
This suggests that pseudo-mitotic groups might be useful for constructing counterexamples for the problem above. However, the example considered by Sankaran and Varadarajan [49, Theorem 3.3] still produces a boundedly acyclic quotient. Indeed, in this situation the kernel of the epimorphism is the group of finitely supported permutations of , which is locally finite and therefore amenable. Hence, by Theorem 4.1 (or simply by Gromov’s Mapping Theorem [28]), the quotient is a boundedly acyclic non-pseudo-mitotic group.
A more interesting situation arises from the context of algebraic -theory. Indeed, following Berrick [3] given a (unital, associative) ring one can embed it into its cone as a two-sided ideal. This leads to a short exact sequence
where is the direct general linear group over the suspension of , usually denoted by [3, page 85]. As discussed in Example 3.15.1, the group is a flabby group, whence pseudo-mitotic (in fact this is also a dissipated group as proved by Berrick [3, pages 84–85]). On the other hand, since one can compute the -groups of the original ring in terms of the plus construction over , in general is far from being acyclic [3]. Hence, a natural question is the following:
Question 4.8.
Let be a ring. Is the group boundedly acyclic?
A negative answer to this question would show that boundedly acyclic groups are not closed under passage to quotients. On the other hand, here we prove the following:
Proposition 4.9.
Let be a ring. Then, the group is -boundedly acyclic.
Proof.
Recall that a group extension provides an exact sequence in bounded cohomology in low degrees [44, Corollary 12.4.1 and Example 12.4.3], which in this case gives
Using the fact that is pseudo-mitotic, whence boundedly acyclic (Theorem 1.3), we then have
and
We show that , whence the thesis.
It suffices to show that has commuting conjugates [21]; that is, for every finitely generated subgroup there exists such that and commute. Now let be finitely generated. Then there exists some such that . Let be a permutation matrix that swaps the basis vectors with . Then acts trivially on the span of , and acts trivially on the span of ; therefore, these subgroups commute. We conclude that has commuting conjugates, and so [21]. This finishes the proof. ∎
4.4. Directed unions
The operations we have looked at so far are known to preserve amenability. This is not surprising since amenable groups are the most illustrious examples of boundedly acyclic groups. One further operation that preserves amenability is that of directed unions. Here we study the behaviour of bounded acyclicity under directed unions, and show that it is preserved under an additional technical requirement.
To proceed with the proof, it is convenient to consider the following dual version of UBC:
Definition 4.10.
Let , and let be a group such that . We define the -th vanishing modulus of as the minimal such that the following holds:
For each there exists such that
Example 4.11.
Every amenable group has an -th vanishing modulus of , for all . Indeed, the proof of bounded acyclicity of amenable groups [24, Theorem 3.6] exhibits a contracting chain homotopy for the cochain complex , which has norm in every degree. Hence, given , we can just set and obtain:
In our definition the vanishing modulus takes values in . It turns out that only finite values are possible.
Lemma 4.12.
Let and let be such that . Then the -th vanishing modulus of is finite.
Proof.
This is implicit in the work of Matsumoto and Morita [42]: Because of , we have . Hence, the bounded linear map has closed range; by the Open Mapping Theorem, induces a Banach space isomorphism
Let be the inverse of . If , then the definition of the quotient norm on shows that there exists a with
Thus, the constant is a finite upper bound for the -th vanishing modulus. ∎
Proposition 4.13.
Let be a group that is the directed union of a directed family of subgroups. Moreover, let and suppose that for all , and that there is a uniform, finite upper bound for the -th vanishing moduli of all the ’s. Then .
Proof.
Let be a common upper bound for the -th vanishing moduli of the ’s. We show that the -th vanishing modulus of is at most . Let be a bounded cocycle. For each , we set
It suffices to show that . To this end, we use the Banach–Alaoglu Theorem: By construction, each is a bounded weak-closed subset of and for all with . Moreover, : By hypothesis, there exists with and . We now extend by ; this extension lies in .
Because the system is directed, the family satisfies the finite intersection property; by the Banach–Alaoglu Theorem, therefore the whole intersection is non-empty. ∎
The following special case will be used when studying the Thompson group (Lemma 6.5):
Corollary 4.14.
Let be a group that is the directed union of a directed family of subgroups. Suppose that the ’s are pairwise isomorphic and -boundedly acyclic. Then is -boundedly acyclic.
In degree , we may get rid of the uniformity condition in Proposition 4.13, thanks to the following surprising fact:
Proposition 4.15.
Let be a -boundedly acyclic group. Then the second vanishing modulus of is .
Proof.
This is essentially a dual version of a result by Matsumoto and Morita [42, Corollary 2.7]. First, the map is injective, since the only bounded homomorphism is the trivial one. We consider the map
where we use the abbreviation , and claim that is the inverse of . By definition, ; moreover, because sends constant functions to constant functions.
We are left to prove the claim. Let , say . We need to show that : Using -invariance, we obtain for all :
Note that all series involved are absolutely convergent because is bounded, which is what allows us to change the order of summation. ∎
Corollary 4.16.
A directed union of -boundedly acyclic groups is -boundedly acyclic. ∎
Proposition 4.15 is essentially equivalent to the fact that the canonical semi-norm in degree is always a norm [42, Corollary 2.7]. This fails already in degree [53, 23], but such examples also have large bounded cohomology and so are difficult to control. Therefore we ask:
Question 4.17.
Does the analogue of Proposition 4.15 hold in higher degrees?
5. Universal bounded acyclicity
In this section, we show that the bounded acyclicity of pseudo-mitotic groups is not a phenomenon confined to real coefficients. Since several different coefficients are involved in this section, we will be explicit and talk about -acyclic groups (Definition 1.2) and -boundedly acyclic groups (Definition 1.1).
Definition 5.1.
Let be a complete valued field, and let be a group. We say that is -boundedly acyclic if for all . If this holds for all complete valued fields , we say that is universally boundedly acyclic.
We can characterize universal bounded acyclicity in very simple terms:
Theorem 5.2.
Let be a group. Then is universally boundedly acyclic if and only if it is -boundedly acyclic and -acyclic.
Remark 5.3.
In fact, Theorem 5.2 even holds degree-wise. More precisely, for a group and an integer the following are equivalent:
-
(1)
and for all ;
-
(2)
for every complete valued field and all .
This will be apparent from the proof, but we prefer to state the theorem in global terms to simplify the notation.
Before giving the proof, we note the following consequence:
Corollary 5.4.
Pseudo-mitotic groups are universally boundedly acyclic.
Proof.
The proof of Theorem 5.2 will be carried out in two steps: the Archimedean and the non-Archimedean case.
Lemma 5.5.
Let be a group. Then is -boundedly acyclic if and only if is -boundedly acyclic.
Proof.
Because as normed -vector spaces, the cochain complex splits as the direct sum . Therefore, we obtain the isomorphism (over ). The claim easily follows. ∎
Lemma 5.6.
Let be a group. Then is -boundedly acyclic for every non-Archimedean field if and only if it is -acyclic.
Proof.
Suppose that is -boundedy acyclic for every non-Archimedean field . Endowing an arbitrary field with the trivial norm, we deduce that for every field . It then follows from the Universal Coefficient Theorem [10, Chapter I] that is -acyclic for every field , that is, for all . In particular, is -acyclic and -acyclic for every prime ; so is -acyclic [30, Corollary 3A.7].
Conversely, let us suppose that is -acyclic, and let be a non-Archimedean field . By Lemma 2.1, the comparison map is injective. So it suffices to show that . This follows immediately from the Universal Coefficient Theorem; hence, is -boundedly acyclic. ∎
Proof of Theorem 5.2.
Remark 5.7.
Corollary 5.4 provides many examples of groups that are universally boundedly acyclic. One could ask whether something similar could be said for the stronger notion of universal amenability, defined analogously using the general notion of -amenability for valued fields defined by Shikhof [51]. However, it turns out that if is -amenable in the sense of Shikhof for every prime , then is trivial [20, Example 5.5, Theorem 6.2]. The same holds for the weaker notion of normed -amenability [20], which also implies bounded -acyclicity [20, Theorem 1.3].
Corollary 5.8.
Let be a universally boundedly acyclic group. Then, for all , we have , with the standard absolute value on .
Proof.
The short exact sequence induces a long exact sequence [24, proof of Proposition 2.13]
By Theorem 5.2, the group is -boundedly acyclic and -acyclic. The Universal Coefficient Theorem and -acyclicity give for all . Therefore, the long exact sequence and surjectivitiy of the induced map show that for all . ∎
6. Thompson groups and their siblings
The groups , , and were introduced by Richard Thompson in 1965; they are some of the most important groups in geometric and dynamical group theory. These groups can be realized as groups of homeomorphisms of the interval, the circle, and the Cantor set respectively; these realizations exhibit inclusions . We refer the reader to the literature [15] for a detailed discussion.
The groups , , and are finitely presented, even of type . Moreover, and are simple (in fact, they were the first examples of infinite finitely presented simple groups). On the other hand, has abelianization , but its derived subgroup is simple, and infinitely generated.
The rational cohomology [26, 9] and, with the exception of , the integral cohomology [11, 26, 54] of these groups has been computed. However, little is known about their real bounded cohomology. We formulate one question for each group:
Question 6.1.
Is the Thompson group boundedly acyclic?
Question 6.2.
Does the following hold?
The bounded cohomology of the Thompson group is given by
where the non-trivial classes are spanned by cup powers of the bounded real Euler class.
Question 6.3.
Is the Thompson group boundedly acyclic?
The rest of this section is devoted to discussing these three questions, how they relate to each other, and provide some evidence towards positive answers.
One may also formulate corresponding questions for every degree, namely whether the previous descriptions hold up to degree . We will see that all three questions have a positive answer up to degree , while to our knowledge nothing is known from degree onwards.
6.1. On the bounded cohomology of
We recall the definition of .
Definition 6.4.
The Thompson group is the group of orientation-preserving piecewise linear homeomorphisms of the interval with the following properties:
-
(1)
has finitely many breakpoints, all of which lie in ;
-
(2)
Away from the breakpoints, the slope of is a power of .
The map , where is the slope of at , is a surjective homomorphism, called the germ at . Similarly, there is a germ at , leading to a surjective homomorphism . This is the abelianization of , so the derived subgroup coincides with the subgroup of homeomorphisms that are compactly supported in .
The most important open question about is whether is amenable or not. Since amenable groups are boundedly acyclic, a negative answer to Question 6.1 would disprove its amenability. The general philosophy is that is very close to being amenable, and so it is likely to satisfy most properties that are somewhat weaker than amenability.
For example: The group is -boundedly acyclic. This can be deduced from the explicit description of its rational cohomology [26], by using arguments analogous to those of Heuer and Löh for the computation of the second bounded cohomology of [31] (although a direct approach is possible [21]). To our knowledge, nothing is known about the bounded cohomology of with trivial real coefficients in higher degrees, although vanishing is known in every degree with mixing coefficients [45].
The connection between pseudo-mitotic groups and is more transparent when passing to the derived subgroup. The following equivalent formulation will be relevant:
Lemma 6.5.
Let . Then the Thompson group is -boundedly acyclic if and only if is -boundedly acyclic.
Proof.
If is -boundedly acyclic, then is also boundedly acyclic by Theorem 4.1, or more simply by coamenability [47].
Conversely, let us suppose that is -boundedly acyclic. Let and be sequences of dyadic rationals in that converge to and respectively. Then may be expressed as the directed union of the subgroups consisting of elements supported in . Since each group is isomorphic to , the group is a directed union of pairwise isomorphic -boundedly acyclic groups. It follows from Corollary 4.14 that is -boundedly acyclic. ∎
The derived subgroup is a group of boundedly supported homeomorphisms of the interval. In analogy with Example 3.12, one may ask whether is pseudo-mitotic. This is not the case, because is not acyclic [26]. Intuitively, cannot be dissipated, since a dissipator could not possibly have finitely many breakpoints. However, a countably singular analogue of is dissipated:
Definition 6.6.
Let be the group of orientation-preserving homeomorphisms of the interval with the following properties:
-
(1)
There exists a closed and countable set such that is linear on each component of ;
-
(2)
Away from , the slope of is a power of .
Since the set of breakpoints of each element is contained in , the germs at and are still defined, and is the subgroup of homeomorphisms that are compactly supported in .
Proposition 6.7.
The groups and are boundedly acyclic.
Proof.
Once again, the bounded acyclicity of follows from that of by Theorem 4.1.
We show that is dissipated. Let and be sequences of dyadic rationals in converging to and , respectively. For every , let be the subgroup consisting of homeomorphisms supported in . We show that there exists a dissipator for , that is
-
(1)
For every , we have ;
-
(2)
For every , the element
is in .
To this end, let us set and . We then pick a dyadic rational in such that
is a power of . Moreover, given a dyadic rational , we can extend to a sequence of dyadic rationals converging to and such that for every the ratio
is a power of .
Now we define piecewise as follows:
and let be the unique affine isomorphism on each of these pieces. Notice that is supported in , and the set of breakpoints is closed, countable, and consists only of dyadic rationals. Since all the slopes are powers of , this implies that .
We claim that is a dissipator for . First, notice that by construction and the definition of and , we have
for every . This shows that satisfies property (1).
Finally, for every , the support of the homeomorphism is contained in . Moreover, the set of breakpoints of is still a closed, countable set consisting only of dyadic rationals. Hence, also satisfies property (2).
We believe that a careful study of the embedding could lead to some understanding of the bounded cohomology of .
6.2. On the bounded cohomology of
We recall the definition of :
Definition 6.8.
The Thompson group is the group of orientation-preserving piecewise linear homeomorphisms of the circle with the following properties:
-
(1)
has finitely many breakpoints, all of which lie in ;
-
(2)
Away from the breakpoints, the slope of is a power of ;
-
(3)
preserves .
The stabilizer of for the canonical -action on the circle is canonically isomorphic to the Thompson group .
Since acts minimally on the circle, it admits a second bounded cohomology class, namely the real Euler class [41, 12]. The bounded Euler class is a refinement of the classical Euler class. All cup powers of the classical Euler class are non-trivial in cohomology [26]; thus, also the cup-powers of the bounded Euler class are non-trivial in ; this was first noticed by Burger and Monod [13]. Therefore, Question 6.2 is asking whether these are the only bounded cohomology classes. In degree , this is known to be true [31], but again, to our knowledge nothing is known in higher degrees.
The main goal of this section is to show that a positive answer to Question 6.1 implies a positive answer to Question 6.2, and this implication holds degree-wise. In order to do this, we prove the following general criterion for computing the bounded cohomology of groups acting highly transitively on the circle with boundedly acyclic stabilizers:
Proposition 6.9.
Let . Let be a group acting orientation-preservingly on the circle, let be an orbit of with . Suppose that the following holds:
-
(1)
For all , the action of on the set of circularly ordered -tuples in is transitive;
-
(2)
For all , the stabilizer of a circularly ordered -tuple is -boundedly acyclic.
Then is generated by the bounded Euler class of this circle action of and
for all , generated by the cup-powers of Euler class.
Recall that a -tuple in is circularly ordered if there exists a point such that is an ordered -tuple in the interval. We follow the convention that circularly ordered tuples are non-degenerate, i.e., they consist of pairwise distinct entries.
For the proof, we follow the general principle of computing bounded cohomology through boundedly acyclic actions. Boundedly acyclic stabilizers lead to boundedly acyclic modules:
Lemma 6.10.
Let be a group and let be an action of on a set that has only finitely many orbits . Let . If each of the orbits has -boundedly acyclic stabilizer, then we have for all that
Proof.
Let . Because is finite, we have and
We show that each of the summands is trivial. Let and let be the stabilizer of a point in . Then, by the Eckmann–Shapiro Lemma in bounded cohomology [44, Proposition 10.13], we obtain
the last term is trivial, because is -boundedly acyclic by hypothesis. ∎
The effect of boundedly acyclic stabilizers is studied more systematically in a forthcoming article on boundedly acyclic covers and relative simplicial volume [38].
Proof of Proposition 6.9.
The given -action on gives a simplicial -resolution [24, Lemma 4.21].
Claim A.
The -resolution is boundedly acylic up to degree , i.e., for all and all , we have
Proof of Claim A.
The -space consists only of finitely many -orbits: Indeed, every tuple can be permuted to be circularly ordered (possibly with repetitions), and only finitely many permutations and repetition patterns are possible. Moreover, acts transitively on circularly ordered tuples of every given size .
The stabilizer groups of the -space are all -boundedly acyclic by hypothesis. Thus, Lemma 6.10 shows the claim. ∎
Therefore, we can apply the fact that boundedly acyclic resolutions compute bounded cohomology [48, Proposition 2.5.4] and symmetrisation [24, Section 4.10] to conclude that
(1) |
for all . Here, denotes the subcomplex of alternating cochains, i.e., functions with
for all and all permutations of .
Claim B.
Let .
-
(1)
If is odd, then .
-
(2)
If is even, then , generated by the function constructed in the proof below.
Proof of Claim B.
Let . We first show that is determined by the value on a single circularly ordered tuple: Indeed, vanishes on tuples with a repetition; all other tuples may be permuted to be circularly ordered. Moreover, since acts transitively on the set of circularly ordered tuples, is constant on the set of all circularly ordered tuples. In particular, .
As , there exists a circularly ordered tuple .
Let be odd. It suffices to show that . With also is circularly ordered. Because is odd, these two tuples differ by an odd permutation. As is both constant on all circulary ordered tuples and alternating, we obtain
and thus . Therefore, .
Let be even. We define as follows: On tuples with a repetition, we define to vanish. If has no repetition, we set
where is a permutation such that is circularly ordered; this permutation is only unique up to a -cycle, but since is even, is well-defined. Because acts orientation-preservingly and because there exists at least one circularly ordered -tuple, this gives a well-defined non-trivial element in . Therefore, . ∎
In view of Claim B, the cochain complex is (up to degree ) isomorphic to the cochain complex
(whose coboundary operator is necessarily trivial) and if is even, then is non-trivial in . In particular, we obtain
for all .
As for degree , under our assumptions we cannot show that follows the same periodic pattern. However we still have:
Claim C.
The differential is trivial.
Proof.
This is obvious if is odd, since then by Claim B.
Suppose instead that is even, and let be a function that takes the constant value on circularly ordered tuples. Then, if is a circularly ordered tuple, we have
since is even. ∎
Therefore
as well. It remains to deal with the bounded Euler class and its powers:
Claim D.
The bounded Euler class is non-trivial.
Proof of Claim D.
We make the isomorphism in (1) more explicit. Let . For , we consider the map
Then is a degree-wise bounded -cochain map that extends the identity on the resolved module . Because the resolution is strong [24, Lemma 4.21], induces an isomorphism for all [48, Proposition 2.5.4, Remark 2.5.5].
As the inclusion is a -cochain map that induces an isomorphism , we conclude that induces an isomorphism .
By construction, gives the orientation cocycle of the -action. Because of , we know that the bounded Euler class
is non-zero in . ∎
Because of Claim D and , we conclude that is generated by the bounded Euler class.
Claim E.
For each , the cup-power is non-trivial.
Proof of Claim E.
In view of the relation between and (proof of claim D) and the above description of , it suffices to show that is non-trivial, where denotes the standard cup-product on the cochain level (notice that even if is alternating, the non-trivial cup-product is not so). Indeed, we have
This completes the proof of Proposition 6.9. ∎
Remark 6.11.
The second hypothesis in Proposition 6.9 was used to show that the modules are -boundedly acyclic, which in turn is used to apply the computation of bounded cohomology through acyclic resolutions [48, Proposition 2.5.4]. Note however that this result does not require -bounded acyclicity of all stabilizers. Indeed, it is enough to ask that the stabilizer of a circularly ordered -tuple is boundedly acyclic, for . To keep notation simple, we chose to state Proposition 6.9 with the stronger hypothesis.
We apply this to , to show that if is -boundedly acyclic, then Question 6.2 has a positive answer up to degree :
Corollary 6.12.
If is boundedly acyclic, then (with the cup-product structure) is isomorphic to the polynomial ring with , and the bounded Euler class of is a polynomial generator of .
Proof.
For each , the group acts transitively on the set of circularly ordered -tuples in ; the stabilizers of this action are isomorphic to direct powers of [15]. In particular, the stabilizers are boundedly acyclic by Theorem 4.1. Therefore, Proposition 6.9 is applicable and we obtain that is isomorphic as a graded -algebra to with corresponding to the bounded Euler class .
Alternatively, in this case, the non-triviality of the powers of the bounded Euler class is already known through the computations of Ghys–Sergiescu and Burger–Monod [13]. ∎
Corollary 6.12 also holds in a range up to (with the same proof).
Corollary 6.13.
If is boundedly acyclic and , then
defines an isomorphism of graded -algebras; here, carries the cup-product structure, , and denotes the projection onto the -th factor for each . Moreover, the canonical semi-norm on then is a norm.
Proof.
We combine Corollary 6.12 with suitable Künneth arguments. As usual in bounded cohomology, some care is necessary to execute this.
We first show that the polynomial ring embeds into : The -algebra homomorphism given by
is injective [26]. Therefore, the Künneth Theorem shows that
yields an injective -algebra homomorphism . In combination with the Universal Coefficient Theorem, we obtain: For every polynomial , there exists a class with
These non-trivial evaluations show that also the bounded version
is injective; even more, for each , we have
and thus . So far, we did not use the postulated bounded acyclicity of .
It remains to show that is surjective. To this end, it suffices to inductively (on ) establish that holds for all .
The base case is handled in Corollary 6.12; moreover, the evaluation argument above shows that the canonical semi-norm on indeed is a norm.
For the induction step, let us assume that the claim holds for . We recall that for group extensions there is a Hochschild–Serre Spectral Sequence
in bounded cohomology, whenever the canonical semi-norm on is a norm [44, Proposition 12.2.1]. Applying this spectral sequence to the trivial product extension
shows that the degree-wise dimensions of are at most the degree-wise dimensions of (with for all ). Hence, is surjective. In particular, again by the evaluation argument above, the canonical semi-norm on all of is a norm. ∎
Remark 6.14.
Analogous results are obtained by Monod and Nariman [46], who computed the full bounded cohomology of the groups of orientation-preserving homeomorphisms of the circle and the -disc. In fact, in Proposition 6.9 one can replace orbits by fat orbits (that is, orbits of fat points [46]). This allows to compute the full bounded cohomology of from the bounded acyclicity of . Moreover, again using fat orbits, one can deduce from Proposition 6.7 a positive answer to Question 6.2 for a natural countably singular analogue of Thompson’s group .
6.3. On the bounded cohomology of
We recall the definition of :
Definition 6.15.
The Thompson group is the group of piecewise linear right-continuous bijections of the circle with the following properties:
-
(1)
has finitely many breakpoints, all of which lie in ;
-
(2)
Away from the breakpoints, is orientation-preserving and the slope of is a power of ;
-
(3)
preserves .
It was recently proved that is acyclic [54]: This had been conjectured by Brown [9], who already proved that is rationally acyclic. Moreover, is uniformly perfect [15]; so, using the same argument as in Remark 3.6, we deduce that is -boundedly acyclic.
While the proof of acyclicity of is involved, the proof of rational acyclicity is much simpler and only relies on standard arguments in equivariant homology. A bounded analogue of equivariant homology theory has recently been developed [37], but it does not seem possible to directly translate Brown’s proof to bounded cohomology.
Appendix A Pseudo-mitotic groups are boundedly acyclic
We prove Theorem 1.3. The proof is an adaption of the original proofs for ordinary cohomology [4, 55] to the setting of bounded cohomology. In bounded cohomology, we keep additional control on primitives, similarly to Matsumoto and Morita for certain homeomorphism groups [42] and similarly to the case of mitotic groups [39].
The main ingredient in the proof is the following proposition which is an adaptation for the pseudo-mitotic setting of the mitotic case [39, Proposition 4.6]:
Proposition A.1.
Let and let . Let
be a chain of group homomorphisms such that
-
(1)
The homomorphism is a pseudomitosis of in ;
-
(2)
For every we have ;
-
(3)
For every the homomorphisms and satisfy -.
Then, for all we have
Moreover, there exists a constant (depending only on and ) such that the composition satisfies - for every .
We give the proof of Proposition A.1 in Section A.1. Following the mitotic case [39], we show first how to deduce Theorem 1.3 from this result.
Proof of Theorem 1.3.
Let be a pseudo-mitotic group. According to Theorems 2.3 and 3.4, it is sufficient to show that for every the group satisfies -.
Let and let be a boundary, i.e., there exists a chain such that . Since both and involve only finitely many elements of , there exists a finitely generated subgroup of such that and We show that the inclusion satisfies -, where only depends on . This condition readily implies that satisfies -, whence the thesis.
Since is pseudo-mitotic, has a mitotis into . Let and be witnesses of such a pseudo-mitosis. Then, also admits a pseudo-mitosis into the following finitely generated subgroup of :
By iterating this construction we get a sequence of finitely generated groups such that at each step the inclusion is a pseudo-mitosis.
Following verbatim the proof of the mitotic case [39, Theorem 1.2], by induction on and using Proposition A.1, one can show that the inclusion of into a sufficiently large satisfies -, where only depends on .
Using the fact that , this implies that there exists with
This shows that satisfies - for all positive degrees ; whence, is boundedly acyclic (Theorem 2.3). ∎
A.1. Proof of Proposition A.1
This section is devoted to the proof of Proposition A.1. The proof is based on a refinement of Varadarajan’s proof [55, Proposition 1.4], additionally taking the norm of the morphisms involved into account. Our approach will closely follow the mitotic case [39, Appendix A].
Proof of Proposition A.1.
We prove the statement in degree . For convenience we write
Since for every , also in the same degrees. The fact that for every was already proved by Varadarajan [55, Proposition 1.4]. In order to adapt the mitosis proof [39, Appendix A] to the case of pseudo-mitoses, it is convenient to recall Varadarajan’s argument.
Let and be witnesses of the pseudo-mitosis . Then, we define the map
Notice that is a group homomorphism by Condition (2) of the definition of pseudo-mitosis [55, proof of Proposition 1.4].
Let and be the inclusions into the first and the second factor, respectively. Let denote the conjugation with respect to , i.e., for every . Moreover, let denote the diagonal homomorphism. Then, for every , we have:
On the other hand, we also have
Hence, the Künneth Formula (and its naturality) together with the assumption that for all imply that the following diagram commutes (similarly to the mitotic case [39, p. 729]):
where denote the projections onto the two factors and the inclusions of the factors. The commutativity of the previous diagram leads to
Since the conjugation is trivial in homology, i.e., , we obtain
We are thus reduced to show that the previous construction can be controlled in such a way that satisfies the required condition.
Let . We will construct a controlled -primitive for . Following the mitotic case [39, p. 731], we have
(2) |
on the chain level, where is bounded and admits a bound that only depends on the given and (the proof uses hypothesis (3) in the statement).
To complete the construction of a controlled -primitive for we consider the following chain homotopy
between and the identity. That the previous map is in fact such a chain homotopy and that is proved as in the mitotic case [39, Lemma A.2] (notice that for convenience we changed the sign of ). We then have:
where we moved from the first line to the second one using the formula (2) and the last equality holds because is a cycle. Moreover, using the fact that group homomorphisms induce norm non-increasing chain maps, we have that the norm
is bounded from above by a quantity depending only on and (since this is true for and ). This shows that satisfies -, as claimed. ∎
Appendix B Computation of cup-powers
In the following, we give the combinatorial part of the proof of Claim E in the proof of Proposition 6.9. We use the notation established in the proof of Proposition 6.9.
Lemma B.1.
For all , we have
Proof.
Proceeding inductively, it suffices to show that
This is a purely combinatorial statement. Let be a circularly ordered -tuple over . In order to simplify notation, we will write for , etc. In this notation, since is circularly ordered, it suffices to show that is equal to
By definition of and the cup-product on simplicial cochains, we have
Because is even, we can fix one position and obtain via cyclic permutations that
Flipping and changes both the sign of and of . Therefore, we obtain
Here, we use the following notation: ; the permutation/tuple on is obtained by using , in the first two positions and then filling up with , etc.
Let be the sequence of elements in , in order. Then
for all , ; here, we set of to if is even, and to if is odd. We distinguish two cases:
-
•
If is odd, then is zero, because there are equally many even and odd numbers in .
-
•
If is even, then contains one more odd number than even numbers, whence
Therefore, since there are even numbers inside , we obtain
as claimed. ∎
Similar computations can also be found in the literature [34].
References
- [1] G. Baumslag, E. Dyer, and A. Heller. The topology of discrete groups. J. Pure Appl. Algebra, 16(1):1–47, 1980.
- [2] C. Bavard. Longueur stable des commutateurs. Enseign. Math. (2), 37(1-2):109–150, 1991.
- [3] A. J. Berrick. An approach to algebraic -theory, volume 56 of Research Notices in Mathematics. Pitman Advanced Publishing Program, 1982.
- [4] A. J. Berrick. Universal groups, binate groups and acyclicity. In Group theory (Singapore, 1987), pages 253–266. de Gruyter, Berlin, 1989.
- [5] A. J. Berrick. A topologist’s view of perfect and acyclic groups. In Invitations to geometry and topology, volume 7 of Oxf. Grad. Texts Math., pages 1–28. Oxford Univ. Press, Oxford, 2002.
- [6] A. J. Berrick. The acyclic group dichotomy. J. Algebra, 326:47–58, 2011.
- [7] A. J. Berrick, I. Chatterji, and G. Mislin. From acyclic groups to the bass conjecture for amenable groups. Math. Ann., 329(4):597–621, 2004.
- [8] A. J. Berrick and K. Varadarajan. Binate towers of groups. Arch. Math. (Basel), 62(2):97–111, 1994.
- [9] K. S. Brown. The geometry of finitely presented infinite simple groups. In Algorithms and classification in combinatorial group theory, pages 121–136. Springer, 1992.
- [10] K. S. Brown. Cohomology of groups, volume 87 of Graduate Texts in Mathematics. Springer Science & Business Media, 2012.
- [11] K. S. Brown and R. Geoghegan. An infinite-dimensional torsion-free group. Invent. Math., 77(2):367–381, 1984.
- [12] M. Bucher, R. Frigerio, and T. Hartnick. A note on semi-conjugacy for circle actions. Enseign. Math., 62(3):317–360, 2017.
- [13] M. Burger and N. Monod. Continuous bounded cohomology and applications to rigidity theory. Geom. Funct. Anal., 12(2):219–280, 2002.
- [14] D. Calegari. scl, volume 20 of MSJ Memoirs. Mathematical Society of Japan, Tokyo, 2009.
- [15] J. W. Cannon, W. J. Floyd, and W. R. Parry. Introductory notes on Richard Thompson’s groups. Enseign. Math., 42:215–256, 1996.
- [16] J. W. S. Cassels. Local fields, volume 3 of London Mathematical Society Student Texts. Cambridge University Press Cambridge, 1986.
- [17] P. de La Harpe and D. McDuff. Acyclic groups of automorphisms. Comment. Math. Helv., 58(1):48–71, 1983.
- [18] D. B. Epstein. The simplicity of certain groups of homeomorphisms. Compos. Math., 22(2):165–173, 1970.
- [19] G. M. Fisher. On the group of all homeomorphisms of a manifold. Trans. Amer. Math. Soc., 97(2):193–212, 1960.
- [20] F. Fournier-Facio. Normed amenability and bounded cohomology over non-Archimedean fields. arXiv:2011.04075 To appear in Mem. Amer. Math. Soc., 2020.
- [21] F. Fournier-Facio and Y. Lodha. Second bounded cohomology of groups acting on -manifolds and applications to spectrum problems. arXiv preprint arXiv:2111.07931, 2021.
- [22] F. Fournier-Facio, C. Löh, and M. Moraschini. Bounded cohomology of finitely presented groups: Vanishing, non-vanishing, and computability. arXiv preprint arXiv:2106.13567, 2021.
- [23] F. Franceschini, R. Frigerio, M. B. Pozzetti, and A. Sisto. The zero norm subspace of bounded cohomology of acylindrically hyperbolic groups. Comment. Math. Helv., 94(1):89–139, 2019.
- [24] R. Frigerio. Bounded cohomology of discrete groups, volume 227 of Mathematical Surveys and Monographs. American Mathematical Soc., 2017.
- [25] E. Ghys. Groupes d’homéomorphismes du cercle et cohomologie bornée. In The Lefschetz centennial conference, Part III (Mexico City, 1984), volume 58 of Contemp. Math., pages 81–106. Amer. Math. Soc., Providence, RI, 1987.
- [26] É. Ghys and V. Sergiescu. Sur un groupe remarquable de difféomorphismes du cercle. Comment. Math. Helv., 62(1):185–239, 1987.
- [27] R. I. Grigorchuk. Some results on bounded cohomology. In Combinatorial and geometric group theory (Edinburgh, 1993), volume 204 of London Math. Soc. Lecture Note Ser., pages 111–163. Cambridge Univ. Press, Cambridge, 1995.
- [28] M. Gromov. Volume and bounded cohomology. Inst. Hautes Études Sci. Publ. Math., 56:5–99 (1983), 1982.
- [29] P. Hall. Some constructions for locally finite groups. J. London Math. Soc., 1(3):305–319, 1959.
- [30] A. Hatcher. Algebraic Topology. Cambridge University Press, 2002.
- [31] N. Heuer and C. Löh. The spectrum of simplicial volume. Invent. Math., 223(1):103–148, 2021.
- [32] N. V. Ivanov. Foundations of the theory of bounded cohomology. volume 143, pages 69–109, 177–178. 1985. Studies in topology, V.
- [33] N. V. Ivanov. Leray theorems in bounded cohomology theory. arXiv preprint arXiv:2012.08038, 2020.
- [34] S. Jekel. Powers of the Euler class. Adv. Math., 229(3):1949–1975, 2012.
- [35] B. E. Johnson. Cohomology in Banach algebras. Mem. Amer. Math. Soc., 127, 1972.
- [36] D. M. Kan and W. P. Thurston. Every connected space has the homology of a . Topology, 15(3):253–258, 1976.
- [37] K. Li. Bounded cohomology of classifying spaces for families of subgroups. arXiv:2105.05223 To appear in Algebr. Geom. Topol., 2021.
- [38] K. Li, C. Löh, and M. Moraschini. Bounded acyclicity and relative simplicial volume. arXiv preprint: arXiv:2202.05606.
- [39] C. Löh. A note on bounded-cohomological dimension of discrete groups. J. Math. Soc. Japan (JMSJ), 69(2):715–734, 2017.
- [40] J. Mather. The vanishing of the homology of certain groups of homeomorphisms. Topology, 10:297–298, 1971.
- [41] S. Matsumoto. Numerical invariants for semiconjugacy of homeomorphisms of the circle. Proc. Amer. Math. Soc., 98(1):163–168, 1986.
- [42] S. Matsumoto and S. Morita. Bounded cohomology of certain groups of homeomorphisms. Proc. Amer. Math. Soc., 94(3):539–544, 1985.
- [43] N. Monod. Lamplighters and the bounded cohomology of thompson’s group. arXiv preprint: arXiv:2112.13741.
- [44] N. Monod. Continuous bounded cohomology of locally compact groups, volume 1758 of Lecture Notes in Mathematics. Springer-Verlag, Berlin, 2001.
- [45] N. Monod. On the bounded cohomology of semi-simple groups, -arithmetic groups and products. J. Reine Angew. Math., 640:167–202, 2010.
- [46] N. Monod and S. Nariman. Bounded and unbounded cohomology of diffeomorphism groups. 2021. arXiv preprint arXiv:2111.04365.
- [47] N. Monod and S. Popa. On co-amenability for groups and von Neumann algebras. C. R. Acad. Sci. Canada, 25(3):82–87, 2003.
- [48] M. Moraschini and G. Raptis. Amenability and acyclicity in bounded cohomology theory. arXiv preprint arXiv:2105.02821, 2021.
- [49] P. Sankaran and K. Varadarajan. Some remarks on pseudo-mitotic groups. Indian J. Math., 29:283–293, 1987.
- [50] P. Sankaran and K. Varadarajan. Acyclicity of certain homeomorphism groups. Canadian J. Math., 42(1):80–94, 1990.
- [51] W. H. Schikhof. Non-archimedean invariant means. Compos. Math., 30(2):169–180, 1975.
- [52] J. Schreier and S. Ulam. Über topologische Abbildungen der euklidischen Sphären. Fund. Math., 23(1):102–118, 1934.
- [53] T. Soma. The zero-norm subspace of bounded cohomology. Comment. Math. Helv., 72(4):582–592, 1997.
- [54] M. Szymik and N. Wahl. The homology of the Higman–Thompson groups. Invent. Math., 216(2):445–518, 2019.
- [55] K. Varadarajan. Pseudo-mitotic groups. J. Pure Appl. Algebra, 37:205–213, 1985.
- [56] J. B. Wagoner. Delooping classifying spaces in algebraic -theory. Topology, 11(4):349–370, 1972.