Bi-exactness of relatively hyperbolic groups
Abstract
We prove that finitely generated relatively hyperbolic groups are bi-exact if and only if all peripheral subgroups are bi-exact. This is a generalization of Ozawa’s result which claims that finitely generated relatively hyperbolic groups are bi-exact if all peripheral subgroups are amenable.
1 Introduction
MSC Primary: 20F65. Secondary: 20F67, 20F99, 46L10.Key words and phrases: bi-exact groups, relatively hyperbolic groups, proper arrays.Bi-exactness is an analytic property of groups defined by Ozawa in [16] (by the name of class ). Recall that a group is exact if there exists a compact Hausdorff space on which acts topologically amenably and bi-exact if it is exact and there exists a map such that for every , we have
It is known that exactness is equivalent to Yu’s property A (cf. [13]).
The notion of bi-exactness is of fundamental importance to the study of operator algebras. Notably, Ozawa proved in [14] that the group von Neumann algebra of any non-amenable bi-exact icc group is prime (i.e. it cannot be decomposed as a tensor product of two factors) by showing that is solid. In [18], Ozawa and Popa proved a unique prime factorization theorem which states that if are non-amenable bi-exact icc groups and is decomposed as a tensor product of factors with i.e. , then we have and after permutation of indices and unitary conjugation, each is isomorphic to an amplification of . Subsequently, various rigidity results were proved for bi-exact groups. In particular, Sako showed measure equivalence rigidity for direct product (cf. [20]) and Chifan and Ioana showed rigidity result in von Neumann algebra sense for amalgamated free product (cf. [3]).
It is known that the class of bi-exact groups contains amenable groups, hyperbolic groups, discrete subgroups of connected simple Lie groups of rank one, and (cf. [2][16][17]). On the other hand, unlike exactness, bi-exactness is not preserved under some basic group theoretic constructions. For example, direct products and increasing unions of bi-exact groups are not necessarily bi-exact (e.g. and , where is a free group of rank 2 and is an infinite countable locally finite group, are exact but not bi-exact).
The standard way of proving bi-exactness is using topologically amenable action on a special compact space called ‘boundary small at infinity’. Using this method, Ozawa proved in [15] that any finitely generated group hyperbolic relative to amenable peripheral subgroups is bi-exact. In this paper, we use a different approach and generalize Ozawa’s result by proving the following.
Theorem 1.1.
Suppose that is a finitely generated group hyperbolic relative to a collection of subgroups of . Then, is bi-exact if and only if all subgroups are bi-exact.
The ‘Only if’ direction is obvious and the whole paper is devoted to proving the ‘if’ direction. In relation to this result, it is worth mentioning the following folklore conjecture.
Conjecture 1.2.
Suppose that a finitely generated group is hyperbolic relative to a collection of subgroups of . If all subgroups are exact, then is bi-exact relative to .
For the definition of relative bi-exactness, the reader is referred to Definition 15.1.2 of [2]. Note that the ‘if’ direction of Theorem 1.1 is not a weak version of Conjecture 1.2. Indeed, while the assumption of Theorem 1.1 is stronger, the conclusion is also stronger.
Our proof of Theorem 1.1 is based on a characterization of bi-exactness using exactness and the existence of a proper array (cf. Proposition 2.2) and uses two technologies related to relatively hyperbolic groups. The first one is a bicombing of fine hyperbolic graphs, which was constructed by Mineyev and Yaman in [10] using the same idea as in [9]. The second one is based on the notion of separating cosets of hyperbolically embedded subgroups, which was introduced by Hull and Osin in [8] and developed further by Osin in [12]. We will construct two arrays using each of these techniques and combine them to make a proper array. It is worth noting that the Mineyev-Yaman’s bicombing of relatively hyperbolic groups alone is not sufficient to derive Conjecture 1.2 for the reason explained in Remark 3.7. Therefore, Conjecture 1.2 is still considered open.
The paper is organized as follows. In Section 2, we discuss the necessary definitions and known results about bi-exact groups and relatively hyperbolic groups. In Section 3.1, we give an outline of our proof. Section 3.2 and 3.3 discuss the constructions of arrays based on ideas of [10] and [8], respectively, and Section 3.4 provides the proof of Theorem 1.1.
Acknowledgment. I thank Denis Osin for introducing this topic to me, for sharing his ideas, and for many helpful discussions.
2 Preliminary
2.1 Bi-exact groups
In this section, we introduce some equivalent conditions of bi-exact groups. The definition of an array given below was suggested in [4].
Definition 2.1.
Suppose that is a group, a Hilbert space, and a unitary representation. A map is called an array on into , if satisfies (1) and (2) below. When there exists such , we say that admits an array into .
-
(1)
for all .
-
(2)
For every , we have .
If, in addition, satisfies (3) below, is called proper.
-
(3)
For any , is finite.
Conditions (2) and (3) in Proposition 2.2 are simplified versions of Proposition 2.3 of [4] and Proposition 2.7 of [19] respectively.
Proposition 2.2.
For any countable group , the following three conditions are equivalent.
-
(1)
is bi-exact.
-
(2)
is exact and admits a proper array into the left regular representation .
-
(3)
is exact, and there exist an orthogonal representation to a real Hilbert space that is weakly contained in the regular representation of and a map that is proper and satisfies
for all .
2.2 Relatively hyperbolic groups
There are several equivalent definitions of relatively hyperbolic groups due to Gromov, Bowditch, Farb, and Osin. We adopt Farb’s definition with the Bounded Coset Penetration property replaced by fineness of coned-off Cayley graphs. [7], [5, Appendix], and [11, Appendix] contain proofs of the equivalence of these definitions. To state our definition, we first prepare some terminologies. When we consider a graph without loops or multiple edges, the edge set is defined as a irreflexive symmetric subset of , where is the vertex set (see Section 2.3 for details).
Definition 2.3 (Coned-off Cayley graph).
Suppose that is a finitely generated group with a finite generating set and is a collection of subgroups of . Let be the Cayley graph of with respect to , in which we don’t allow loops nor multiple edges. More precisely, its vertex set and edge set are defined by
(1) |
We define a new graph whose vertex set and edge set are given by
The graph is called coned-off Cayley graph with respect to .
We set the length of any edge of to be 1. acts on naturally by graph automorphism. For a vertex , we denote its stabilizer by .
Definition 2.4 (Hyperbolic space).
A geodesic metric space is called hyperbolic if there exists such that for any , any geodesic path , and any point , there exists satisfying .
Definition 2.5 (Circuit).
A closed path in a graph is called circuit, if doesn’t have self-intersection except its initial and terminal vertices.
Definition 2.6 (Fine graph).
A graph is called fine, if for any and any edge of , the number of circuits of length containing is finite.
The following is the definition of relatively hyperbolic groups on which we work. The reader is referred to [5, Appendix] for the equivalence of Definition 2.7 and the other definitions.
Definition 2.7.
A finitely generated group is called hyperbolic relative to a collection of subgroups , if is finite and for some (equivalently, any) finite generating set of , the coned-off Cayley graph is fine and hyperbolic. A member of the collection of subgroups is called a peripheral subgroup.
Some definitions (e.g. Osin’s definition in [11]) can be stated without requiring to be finite. In Osin’s definition, finiteness of follows if is finitely generated.
2.3 Mineyev-Yaman’s bicombing
In [10], Mineyev and Yaman constructed a bicombing of a fine hyperbolic graph using the same idea as in [9]. We first discuss 1-chains on a graph and a unitary representation of a group acting on a graph in general and then introduce Mineyev-Yaman’s bicombing.
In this section, we consider graphs without loops nor multiple edges, following [10] and [1]. More precisely, for a graph with a vertex set , its edge set is a subset of such that , where we define for any . We call a subset of a set of positive edges if it satisfies . Choosing a set of positive edges means choosing directions of edges. Note that when we consider a group action on a graph, we allow inversion of edges. We begin with auxiliary notations.
For a graph , we denote by the set of 1-chains on over . More precisely, is defined as follows. Consider a direct product of vector spaces over . Note that this summation notation is just a formal sum. We define a quotient space by
Here, we denote by the element in corresponding to where . Note that for any , we have
Fix a set of positive edges . The map
is an isomorphism of vector spaces. The space of 1-chains is defined by
Note that is independent of the choice of . For and a path from to , we define a 1-chain
(2) |
Here, we use the same notation to denote a path and the corresponding 1-chain by abuse of notation.
For each (in fact, we use only cases ), define a subspace of by
and for each define a norm
becomes a Banach space isomorphic to . In particular, is a Hilbert space. Note that doesn’t depend on the choice of . We also have for any .
Definition 2.8.
Suppose that a group acts on by graph automorphisms, then acts on isometrically by
In particular, this action on becomes a unitary representation of and we denote this unitary representation by .
There exists a unique well-defined linear map
such that for any .
Definition 2.9.
Suppose that is a graph. A map is called a homological bicombing, if for any , we have . In particular, if all coefficients of are in for any , is called a -bicombing.
Definition 2.10 (2-vertex-connectivity).
A graph is called 2-vertex-connected, if is connected and for any , is connected, where is an induced subgraph of whose vertex set is .
Theorem 2.11.
Suppose that is a 2-vertex-connected fine hyperbolic graph and is a group acting on . If the number of -orbits of is finite and the edge stabilizer is finite for any , then there exists a -bicombing of satisfying the following conditions.
-
(1)
is -equivariant, i.e. for any and any .
-
(2)
is anti-symmetric, i.e. for any .
-
(3)
There exists a constant such that for any ,
-
(4)
There exist constants and such that for any ,
where is the graph metric of .
Remark 2.12.
Remark 2.13.
Theorem 2.11 was actually proved for what Mineyev and Yaman called a ‘hyperbolic tuple’ which was defined in their paper as a quadruple consisting of a graph, a group acting on it, a subset of the vertex set, and a set of vertex stabilizers (cf. [10, Definitions 20, 27, 29, 38]). However, the statement and the proof of Theorem 47 and Proposition 46 (3) in [10] does not require the whole quadruple, and we simplify their exposition.
2.4 Separating cosets of hyperbolically embedded subgroups
In this section, we define Hull-Osin’s separating cosets and explain their properties required for our discussion in Section 3.3. The notion of separating cosets of hyperbolically embedded subgroups was first introduced by Hull and Osin in [8] and further developed by Osin in [12]. There is one subtle difference in the definition of separationg cosets in [8] and in [12], which is explained in Remark 2.21, though other terminologies and related propositions are mostly the same between them. With regards to this difference, we follow definitions and notations of [12] in our discussion. In this section, suppose that is a group, is a subset of , and is a collection of subgroups of such that generates . Note that and are possibly infinite. We begin with defining some auxiliary concepts.
Let . Note that this union is disjoint as sets of labels, not as subsets of . Let be the Cayley graph of with respect to , which allows loops and multiple edges, that is, its vertex set is and its positive edge set is . We call the relative Cayley graph.
For each , we consider the Cayley graph , which is a subgraph of , and define a metric as follows. We say that a path in is -admissible, if doesn’t contain any edge of . Note that can contain an edge whose label is an element of (e.g. the case when the initial vertex of the edge is not in ) and also can pass vertices of . For , we define to be the minimum of lengths of all -admissible paths from to . If there is no -admissible path from to , then we define . For convenience, we extend to by defining if and otherwise.
Definition 2.15.
Suppose that is a group and is a subset of . The collection of subgroups of is said to be hyperbolically embedded in , if it satisfies the two conditions below.
-
(1)
generates and is hyperbolic.
-
(2)
For any , is locally finite, i.e. for any , is finite.
The following is a simplified version of Proposition 4.28 of [6].
Proposition 2.16.
Suppose that is a finitely generated group hyperbolic relative to a collection of subgroups of . Then, for any finite generating set of , is hyperbolically embedded in .
In the remainder of this section, suppose that is hyperbolically embedded in .
Definition 2.17.
[12, Definition 4.1] Suppose that is a path in the relative Cayley graph . A subpath of is called an -subpath if the labels of all edges of are in . In the case is a closed path, can be a subpath of any cyclic shift of . An -subpath of a path is called -component if is not contained in any longer -subpath of . In the case is a closed path, we require that is not contained in any longer -subpath of any cyclic shift of . Further, by a component, we mean an -component for some . Two -components of a path is called connected, if all vertices of and are in the same -coset. An -component of a path is called isolated, if is not connected to any other -component of .
Remark 2.18.
Note that all vertices of an -component lie in the same -coset.
The following proposition is important, which is a particular case of Proposition 4.13 of [6].
Proposition 2.19.
[8, Lemma 2.4] There exists a constant such that for any geodesic -gon in and any isolated -component of , we have
In what follows, we fix any constant with
(3) |
We can now define separating cosets.
Definition 2.20.
[12, Definition 4.3] A path in is said to penetrate a coset for some , if decomposes as , where are possibly trivial, is an -component, and . Note that if is a geodesic, penetrates any coset of at most once. In this case, is called the component of corresponding to and also the vertices and are called the entrance and exit points of and are denoted by and respectively. If in addition, we have , then we say that essentially penetrates . For and , if there exists a geodesic path from to in which essentially penetrates an -coset , then is called an -separating coset. We denote the set of -separating -cosets by .
Remark 2.21.
The following lemma is immediate from the above definition.
Lemma 2.22.
[8, Lemma 3.2] For any and any , the following holds.
-
(a)
.
-
(b)
.
We state some nice properties of separating cosets, all of which were proven in [8]. For , we denote by the set of all geodesic paths in from to .
Lemma 2.23.
[8, Lemma 3.3] For any , any , and any -separating coset , the following holds.
-
(a)
Every path in connecting to and composed of at most 2 geodesics penetrates .
-
(b)
For any , we have
Proof of Lemma 2.23 is the same as Lemma 3.3 of [8], though their statements are slightly different. Actually, in Lemma 2.23, we don’t need to assume . This difference comes from the difference of definitions of separating cosets, which was mentioned in Remark 2.21.
Corollary 2.24.
[8, Corollary 3.4] For any and any , we have , where we define to be the set of all -cosets which penetrates. In particular, we have , hence is finite.
The following lemma makes into a totally ordered set.
Lemma 2.25.
[8, Lemma 3.5] Suppose and that penetrates an -coset and decomposes as , where are possibly trivial and is an -component corresponding to . Then, we have .
Definition 2.26.
[8, Definition 3.6] Given any , a relation on the set is defined as follows.
Next, we define a set of pairs of entrance and exit points of a separating coset. That is, for , , and , we define
Note that because is a -separating coset, any geodesic from to penetrates by Lemma 2.23 (a).
Lemma 2.27.
[8, Lemma 3.8] For any , , and , the following holds.
-
(a)
.
-
(b)
.
-
(c)
.
Lemma 2.28.
For any and any , is either penetrated by all geodesics from to or penetrated by all geodesics from to .
Proof.
Suppose there exists a geodesic which doesn’t penetrate . For any geodesic , we can apply Lemma 2.23 (a) to a path and conclude penetrates . ∎
The following lemma is crucial in constructing an array that satisfies the bounded area condition.
Lemma 2.29.
[8, Lemma 3.9] For any and any , can be decomposed as where
-
(a)
and we have for any ,
-
(b)
and we have for any ,
-
(c)
.
3 Main theorem
3.1 Overview of proof
In this section, we explain the idea of the construction of a proper array in the proof of Proposition 3.20 (see (9) (10)) by considering two particular cases.
First, suppose that is a hyperbolic group i.e., the collection of peripheral subgroups is empty. We take a symmetric finite generating set of with and define
The Cayley graph , as defined by (1), has no loops or multiple edges and is a 2-vertex-connected locally finite hyperbolic graph. Let be the barycentric subdivision of . Note that is fine, acts on without inversion of edges, and all edge stabilizers are trivial. By Theorem 2.11, there exists a -equivariant anti-symmetric -bicombing of and a constant such that for any , we have
We define a map by (see Definition 3.4 for details of this definition)
It is straightforward to check that is an array into (cf. Definition 2.8) by using Lemma 3.5 (2). Also, we have
(see Remark 2.12, Lemma 3.3, and Lemma 3.5 (1)). Since is a barycentric subdivision of , we have for any , where is a word metric on with respect to . Hence, for any , if satisfies , we have
This implies that is proper. Combined with Proposition 2.2 (3), this gives another proof of the fact that hyperbolic groups are bi-exact. Here, note that is a direct sum of copies , because acts on without inversion of edges and its action on is free. Hence, it is weakly contained by . We can also use Mineyev’s bicombing in [9] instead of Theorem 2.11 in this case.
Second, suppose that is a free product of bi-exact groups and . In this case, are peripheral subgroups and we can construct a proper array on using the normal forms of elements of the free product and proper arrays on and as follows. Let be a proper array into the left regular representation for each . We can assume for any by changing values of on a finite subset of , if necessary. Also, we regard each as a map from to by composing it with the embedding . For , without loss of generality, let be the normal form of , where and . We define a map by
Here, we define . In the same way, we define a map from . It is not difficult to show that the map defined by
is a proper array into .
The case of general relatively hyperbolic groups is a combination of the above two cases. To every finitely generated relatively hyperbolic group, we associate two arrays. The first one is constructed in Section 3.2 starting with a fine hyperbolic graph as in the case of a hyperbolic group above. The second array is a generalized version of the array for . Hull-Osin’s separating cosets play the same role as syllables in the normal form in the free product case. One technical remark is that the condition ‘’ in the free product case is used to prevent the set from becoming infinite, but in the proof of Proposition 3.20, we don’t need this condition thanks to the first array.
3.2 First array
This section is basically a continuation of Section 2.3. Notation and terminology related to graphs follow those in Section 2.3. The goal of this section is to prove Proposition 3.1 by using Mineyev-Yaman’s bicombing.
Proposition 3.1.
Suppose that is a finitely generated group hyperbolic relative to a collection of subgroups of . Then, there exist a finite generating set of , a 2-vertex-connected fine hyperbolic graph on which acts without inversion of edges, and an array into that satisfy the following conditions.
-
(1)
The edge stabilizer is trivial for any edge in .
-
(2)
For any , we have
(4) where is the coned-off Cayley graph of with respect to .
Remark 3.2.
The unitary representation in Proposition 3.1 is weakly contained by the left regular representation . Indeed, since acts on without inversion of edges and all edge stabilizers are trivial, is a direct sum of copies of .
We first prove a few general lemmas about graphs. Lemma 3.3 can be proven for graphs with loops and multiple edges as well exactly in the same way, but we stick to our current setting.
Lemma 3.3.
Suppose that is a connected graph without loops or multiple edges. If are two vertices, are paths from to as 1-chains, and are complex numbers, then we have
Proof.
We have for any , where we define and . Hence, by defining
and
for each , we get the decomposition of edges . Also, we can choose directions of edges so that they point outward from , that is, there exists a set of positive edges such that
and
for all , where . For simplicity, we use the notation , , and . Given a 1-chain , we define linear maps by
By , we have for any ,
In particular, we have .
We will show that, for any with , we have
By identifying with as explained in Section 2.3, we define a linear map by . Note that the sums are finite and depends on the chosen set of positive edges. For any , we have
Here, for any path from to and any with , we have
Indeed, let as a sequence of edges, where and , and as its subsequence by abuse of notation, i.e. and as a 1-chain. We can see that is odd, i.e. with and also . This implies
Hence,
Thus, we finally get
∎
Definition 3.4.
Suppose that is a graph without loops or multiple edges and is a set of positive edges. For such that for any , we define an element by
Note that the above definition is independent of the choice of a set of positive edges.
Lemma 3.5.
For any whose coefficients are real, the following hold.
-
(1)
.
-
(2)
.
Proof.
(1) follows trivially from Definition 3.4. We will prove (2). For any , we have
Therefore, given
we have for any . ∎
Lemma 3.6.
If is a group and is a generating set of such that and , then the Cayley graph is 2-vertex-connected, where
Proof.
Note that generates since . Since acts transitively on by graph automorphisms, it’s enough to prove that is connected. Any two vertices are connected in by an edge having label . Also, for any vertex , if , then is connected to a vertex in by an edge of label . Hence, any two vertices adjacent to are connected in by a path of length at most 3. This implies that is connected. ∎
In the following proof, recall that our coned-off Cayley graph is a graph without loops or multiple edges (see Definition 2.3).
Proof of Proposition 3.1.
We take a symmetric finite generating set of with , and define . The Cayley graph is 2-vertex-connected by Lemma 3.6. Without loss of generality, we can assume that the subgroups are non-trivial. The coned-off Cayley graph with respect to is 2-vertex-connected, because is 2-vertex-connected and doesn’t contain the trivial subgroup. We denote by the barycentric subdivision of . acts on without inversion of edges. Since is a 2-vertex-connected fine hyperbolic graph, so is . Here, we used the fact that doesn’t have loops to ensure is 2-vertex-connected. Also, since the number of -orbits of is finite and the edge stabilizer is trivial for any edge in , the action of on satisfies these conditions as well. Therefore, by Theorem 2.11, there exist a -equivariant anti-symmetric -bicombing of and a constant such that for any , we have
We define a map by
For every , we have
because is -equivariant and anti-symmetric. Given any , Lemma 3.5 (2) implies
Hence, is an array into . By Remark 2.12, for any , the 1-chain is a convex combination of paths from to i.e. there exist paths from to and with such that
Hence, by Lemma 3.3 and Lemma 3.5 (1), we have
Since is a barycentric subdivision of , we have for any ,
hence
∎
Remark 3.7.
Note that is not even proper relative to . For example, if is a free product of infinite groups and , then by Theorem 2.11 (4), we can show that there exists some such that .
3.3 Second array
This section is a continuation of Section 2.4. The goal of this section is to prove Proposition 3.8. We will construct an array on from an array on a subgroup which is a member of a hyperbolicaly embedded collection of subgroups . The construction follows Section 4 of [8] and uses the notion of separating cosets explained in Section 2.4.
Proposition 3.8.
Suppose that is a group, is a subset of , and is a collection of subgroups hyperbolically embedded in . Then, for any and any array on into , there exists an array on into and a constant satisfying the following for any , any separating coset , and any geodesic path in the relative Cayley graph from to , we have
(5) |
The proof of Proposition 3.8 is essentially the same as the proof of Theorem 4.2 of [8], but because we deal with arrays instead of quasi-cocycles, we give full details with all necessary changes to make the proof self-contained.
Suppose that is an array on into the left regular representation . By the embedding , we can think of as a map . We define a map by
where is the left regular representation of .
Remark 3.9.
If are in the same coset of , i.e. there exists a -coset for some such that , then the support of is in .
Lemma 3.10.
For any , the following hold.
-
(1)
.
-
(2)
.
Proof.
(1) For any , if and only if . If , we have
(2) If , we have
∎
Lemma 3.11.
For any , we have
Proof.
In the following, we denote
(6) |
Remark 3.12.
For any , we have
Remark 3.13.
By Lemma 3.10 (1), , hence for any , we have
Lemma 3.14.
For any elements that are in the same coset of , we have
Proof.
For and , we define by
Remark 3.15.
Lemma 3.16.
For any and , the following holds.
-
(a)
.
-
(b)
.
For , we define a constant by
(7) |
where is defined by (6). Because is a locally finite metric space (cf. Definition 2.15 (2)), we have .
Lemma 3.17.
For any , any , and any , we have
Proof.
Finally, we define a map by
This implicitly means that if is empty, then .
Lemma 3.18.
For any , the following hold.
-
(a)
.
-
(b)
.
-
(c)
.
Proof.
For , we define
(8) |
The proof of the following lemma is similar to Lemma 4.7 of [8].
Lemma 3.19.
For any , we have
Proof.
For , suppose that . We define for each -coset . Note that we have
Let , , and be the decomposition in Lemma 2.29.
If , then we have by Remark 3.15.
If , we have
since and . We can argue similarly for , , , , . Hence, we have
Case 1 . Let be the -components of respectively, corresponding to , i.e. we have . Let be paths of length at most 1 connecting to , to , to respectively, whose labels are in . We claim that is isolated in the geodesic triangle . Indeed, if is connected to an -component of where we have , then there exists a path of length at most 1 connecting to since and are in the same -coset. Hence, the path conneting to is shorter than . This contradicts that is a geodesic. Similarly, is not connected to any -component of . Hence, we have by Proposition 2.19. Similarly, we also have , hence
where is defined by (8). Note that the inclusion follows from the inclusions and . Hence, by Lemma 3.14 (see also (7)), we have
Combining with Lemma 3.17 and Lemma 3.18 (c) (also note (3)), we obtain
Case 2 or . We can assume without loss of generality.
2 a) If penetrates , then as in Case 1, let be the -components of respectively, corresponding to and be paths of length at most 1 connecting to , to , to respectively, whose labels are in . In the same way as Case 1, we have . Also, by , we have , hence
2 b) If doesn’t penetrate , let be the -components of respectively, corresponding to , i.e. , and let be paths of length at most 1 connecting to , to respectively, whose labels are in . Because is isolated in the geodesic triangle , we have . Also, because doesn’t penetrate , is isolated in the geodesic 4-gon , hence we have by Proposition 2.19.
In both 2 a) and 2 b), we have and . Thus, by Lemma 3.14 and 3.17, we have
When , we can show in the same way.
Case 3 . Note that by and Lemma 2.23 (a), at least one of and penetrates .
3 a) If both and penetrate , let be as in Case 1. Then, we have in the same way as Case 1. Also, by , we have . Hence,
3 b) If only one of and penetrates , assume penetrates without loss of generality, and let be as in 2 b) of Case 2. Then, we have and . Also by , we have . hence,
In both 3 a) and 3 b), we have , hence by Lemma 3.14 and Remark 3.13, we have
Here, we used . When only penetrates , we can show in the same way.
Summarizing Case 1,2,3, if , we have
We can argue similarly for and as well. Also, by Lemma 2.29 (c), we have . Thus, for any we have
∎
Proof of Proposition 3.8.
3.4 Proof of main theorem
Proposition 3.20.
Suppose that is a finitely generated group hyperbolic relative to a collection of subgroups of . If all subgroups are bi-exact, then is also bi-exact.
Proof.
Note that is finite by definition. Because ’s are exact, is also exact by Corollary 3 of [15]. In the following, we will verify the condition of Proposition 2.2 (3). We take a finite generating set of , a unitary representation , and an array as in Proposition 3.1. Since every is bi-exact, there exists a proper array on into for each by Proposition 2.2 (2). By Proposition 3.8, for each , there exist an array on into and a constant satisfying (5). Here, we used the fact that is hyperbolically embedded in by Proposition 2.16. We define a Hilbert space and a unitary representation of by
(9) |
Since is a direct sum of copies of by Remark 3.2, is weakly contained by . Now, we define a map
(10) |
Because and ’s are arrays, is an array on into . Hence, for any , we have
where we denote for each ,
Hence, for any , we have
Finally, we will show that is proper. Let and satisfy . Since , we get and for any . Because the identity map is bi-Lipschitz, there exists such that for any . By (4), this implies
We denote for simplicity.
Let be the label of a geodesic path from to in , where is a word in the alphabet and . We have
where denotes the number of letters in . In particular, we have and for any .
On the other hand, for each , there exists such that . For simplicity, we denote
If , then we have by definition of separating cosets. If , then by (5), we have
In either case, we have , where
Note that is finite, because is a locally finite metric on and is proper. Therefore, we have
Since , ’s, and are finite, the set on the right-hand side above is finite, hence is proper. ∎
Lemma 3.21.
Subgroups of countable bi-exact groups are also bi-exact.
Proof.
Let be a countable bi-exact group and be a subgroup of . is exact, because is exact and subgroups of exact groups are exact (cf. [2]). By Proposition 2.2 (2), there exists a proper array on into . Note that the restriction of to , that is, is unitarily isomorphic to , hence we have . Also, it is straightforward to show that is a proper array on into . Thus, by Proposition 2.2 (3), is bi-exact. ∎
References
- [1] B. H. Bowditch, Relatively hyperbolic groups, Internat. J. Algebra Comput. 22 (2012), no. 3, 1250016, 66. MR 2922380
- [2] N. P. Brown and N. Ozawa, -algebras and finite-dimensional approximations, Graduate Studies in Mathematics, vol. 88, American Mathematical Society, Providence, RI, 2008. MR 2391387
- [3] I. Chifan and A. Ioana, Amalgamated free product rigidity for group von Neumann algebras, Adv. Math. 329 (2018), 819–850. MR 3783429
- [4] I. Chifan, T. Sinclair, and B. Udrea, On the structural theory of factors of negatively curved groups, II: Actions by product groups, Adv. Math. 245 (2013), 208–236. MR 3084428
- [5] F. Dahmani, Les groupes relativement hyperboliques et leurs bords thèse de doctorat, université louis pasteur, strasbourg I. (2003), https://www-fourier.ujf-grenoble.fr/~dahmani/Files/These.pdf.
- [6] F. Dahmani, V. Guirardel, and D. Osin, Hyperbolically embedded subgroups and rotating families in groups acting on hyperbolic spaces, Mem. Amer. Math. Soc. 245 (2017), no. 1156, v+152. MR 3589159
- [7] G. C. Hruska, Relative hyperbolicity and relative quasiconvexity for countable groups, Algebr. Geom. Topol. 10 (2010), no. 3, 1807–1856. MR 2684983
- [8] M. Hull and D. Osin, Induced quasicocycles on groups with hyperbolically embedded subgroups, Algebr. Geom. Topol. 13 (2013), no. 5, 2635–2665. MR 3116299
- [9] I. Mineyev, Straightening and bounded cohomology of hyperbolic groups, Geom. Funct. Anal. 11 (2001), no. 4, 807–839. MR 1866802
- [10] I. Mineyev and A. Yaman, Relative hyperbolicity and bounded cohomology, https://faculty.math.illinois.edu/~mineyev/math/art/rel-hyp.pdf.
- [11] D. V. Osin, Relatively hyperbolic groups: intrinsic geometry, algebraic properties, and algorithmic problems, Mem. Amer. Math. Soc. 179 (2006), no. 843, vi+100. MR 2182268
- [12] , Acylindrically hyperbolic groups, Trans. Amer. Math. Soc. 368 (2016), no. 2, 851–888. MR 3430352
- [13] N. Ozawa, Amenable actions and exactness for discrete groups, C. R. Acad. Sci. Paris Sér. I Math. 330 (2000), no. 8, 691–695. MR 1763912
- [14] , Solid von Neumann algebras, Acta Math. 192 (2004), no. 1, 111–117. MR 2079600
- [15] , Boundary amenability of relatively hyperbolic groups, Topology Appl. 153 (2006), no. 14, 2624–2630. MR 2243738
- [16] , A Kurosh-type theorem for type factors, Int. Math. Res. Not. (2006), Art. ID 97560, 21. MR 2211141
- [17] , An example of a solid von Neumann algebra, Hokkaido Math. J. 38 (2009), no. 3, 557–561. MR 2548235
- [18] N. Ozawa and S. Popa, Some prime factorization results for type factors, Invent. Math. 156 (2004), no. 2, 223–234. MR 2052608
- [19] S. Popa and S. Vaes, Unique Cartan decomposition for factors arising from arbitrary actions of hyperbolic groups, J. Reine Angew. Math. 694 (2014), 215–239. MR 3259044
- [20] H. Sako, Measure equivalence rigidity and bi-exactness of groups, J. Funct. Anal. 257 (2009), no. 10, 3167–3202. MR 2568688
Department of Mathematics, Vanderbilt University, Nashville 37240, U.S.A.
E-mail: koichi.oyakawa@vanderbilt.edu