Hyperfiniteness of boundary actions of relatively hyperbolic groups
Abstract
We show that if is a finitely generated group hyperbolic relative to a finite collection of subgroups , then the natural action of on the geodesic boundary of the associated relative Cayley graph induces a hyperfinite equivalence relation. As a corollary of this, we obtain that the natural action of on its Bowditch boundary also induces a hyperfinite equivalence relation. This strengthens a result of Ozawa obtained for consisting of amenable subgroups and uses a recent work of Marquis and Sabok.
1 Introduction
This paper studies equivalence relations induced by boundary actions of relatively hyperbolic groups. The study of boundary actions began with the work of Connes, Feldman and Weiss in [5] and Vershik in [21] who studied the actions of free groups on their boundaries. They showed that for a free group, its action on the Gromov boundary is -hyperfinite for every Borel quasi-invariant probability measure on the boundary. Adams [1] later generalized this result to all hyperbolic groups.
Relatively hyperbolic groups were introduced by Gromov [10]; see also the monograph of Osin [17]. Given a relatively hyperbolic group with a collection of parabolic subgroups there is a natural boundary called the Bowditch boundary, denoted , which is a compact metrizable space on which acts naturally by homeomorphisms.
In [18], Ozawa generalized the work of Adams [1] to the actions of relatively hyperbolic groups on their Bowditch boundary under the assumptions that the parabolic subgroups are exact. When the parabolic subgroups of in are amenable, Ozawa [18] proved that the action of on is topologically amenable, and, more generally, when the parabolic subgroups are exact, Ozawa [18] proved that the group is exact. Alternative proofs of the exactness of the group were given by Osin [16] who worked with parabolic subgroups with finite asymptotic dimension and by Dadarlat and Guentner [6] who worked with parabolic subgroups that are uniformly embeddable into a Hilbert space.
In [22], Zimmer introduced the notion of amenability of equivalence relations; see also the work of Connes, Feldman and Weiss [5]. By [2, Theorem 5.1], a measurable action of a countable group on a standard probability space is -amenable if and only if -almost all stabilizers are amenable and the orbit equivalence relation is -amenable.
In this paper we generalize the result of Ozawa and work with relatively hyperbolic groups without any assumptions on the parabolic subgroups. In fact, we consider boundary actions from the Borel perspective. A countable Borel equivalence relation is called hyperfinite if it is a countable increasing union of finite Borel sub-equivalence relations. Dougherty, Jackson and Kechris showed in [8, Corollary 8.2] that the boundary action of any free group induces a hyperfinite orbit equivalence relation. The result of Dougherty, Jackson and Kechris was generalized to cubulated hyperbolic groups by Huang, Sabok and Shinko in [12], and later to all hyperbolic groups by Marquis and Sabok in [15]. In this paper, we prove the following:
Theorem A.
Let be a finitely generated group hyperbolic relative to a finite collection of subgroups and let be the associated relative Cayley graph. Then the natural action of on the geodesic boundary induces a hyperfinite orbit equivalence relation.
Corollary B.
Let be a finitely generated group hyperbolic relative to a finite collection of subgroups . Then the natural action of on the Bowditch boundary induces a hyperfinite orbit equivalence relation.
Corollary B in particular strengthens the result of Ozawa [18] in case the parabolic subgroups are amenable. Indeed, hyperfiniteness implies -amenability for every invariant Borel probability measure and by [3, Theorem 3.3.7], an action of a countable group on a locally compact space by homeomorphisms is topologically amenable if and only if it is -amenable for every invariant Borel probability measure .
We proceed by following a similar approach to [12] and [15], studying geodesic ray bundles in relative Cayley graphs (Definition 2.2). For the case of a cubulating hyperbolic group studied in [12], the crucial property from which the hyperfinitess of the boundary action of follows is the finite symmetric difference of geodesic ray bundles: for any and any , is finite (see [12, Theorem 1.4]). In [20], Touikan showed that this symmetric difference need not be finite in Cayley graphs of general hyperbolic groups, although in [14], Marquis provides many examples of groups acting geometrically on locally finite hyperbolic graphs where this finite symmetric difference property does hold. In [15], Marquis and Sabok define a modified version of the geodesic ray bundle, denoted for and (see [15, Definition 5.5] and Definition 2.6 in our paper) and show ([15, Theorem 5.9]) that these modified geodesic ray bundles satisfy a finite symmetric difference property: for each and for each . Marquis and Sabok then deduce hyperfiniteness of the boundary action as a consequence of this finite symmetric difference property of the modified bundles (see [15, Section 6]).
Local finiteness of the Cayley graph plays a crucial role in establishing the finite symmetric difference property of the bundles in [15]. However, relative Cayley graphs of relatively hyperbolic groups are not locally finite. To make up for this loss of local finiteness, we rely on finiteness results about relative Cayley graphs of relatively hyperbolic groups from [17] (namely, [17, Theorem 3.26]).
We note also that the hyperfiniteness of boundary actions has been studied beyond relatively hyperbolic groups. Przytycki and Sabok have recently established the hyperfiniteness of the actions of a mapping class group of an oriented surface of finite type on the boundaries of the arc graph ([19, Theorem 1.1]) and the curve graph ([19, Corollary 1.2]) of the surface.
Acknowledgement: I owe great thanks to my advisor Marcin Sabok for his continuous support, patience and guidance throughout the production of this work.
2 Preliminaries
In this paper, for a hyperbolic metric space , will denote the geodesic boundary of . We will also denote the horoboundary of (see [15, Section 2.4] for a definition of the horoboundary).
2.1 Relatively hyperbolic groups
Relatively hyperbolic groups were first introduced by Gromov in his seminal paper [10] as a generalization of hyperbolic groups. The following definitions can be found in [17].
Let be a group generated by a finite set , let be a collection of subgroups of and let . The relative Cayley graph associated to and is the Cayley graph with respect to the generating set . This graph can be identified with the coned-off Cayley graph obtained by starting with the Cayley graph of with respect to , adjoining to a vertex for each left coset and connecting each vertex of in to by an edge of length . The notation and refer to the word metrics with respect to the generating sets and , respectively. We will use the notation to denote the closed ball of radius in the metric about the point .
A finitely generated group is hyperbolic relative to a collection of subgroups if there exists a finite generating set of such that the associated relative Cayley graph is hyperbolic and satisfies the bounded coset penetration property (BCP) (see [17, Definition 6.5] for the definition of the BCP; we will not need to use the definition of BCP, so we do not define it here). Relative hyperbolicity is invariant under change of finite generating set by [17, Proposition 2.8].
For a finitely generated group hyperbolic relative to a finite collection of subgroups, there is a natural compact metrizable space on which acts naturally by homeomorphisms, denoted and called the Bowditch boundary (see [4, Section 4] for the construction of the Bowditch boundary). The following theorem is the main ingredient in establishing Corollary B as a result of Theorem A.
Theorem 2.1.
Let be hyperbolic relative to a finite collection of subgroups , with relative Cayley graph . Then embeds -equivariantly and homeomorphically into with countable complement.
Proof.
In [7, Proposition 1, Section A.2], it is shown that the coned-off Cayley graph witnesses the relative hyperbolicity of with respect to according to Definition 2 of relative hyperbolicity from [4]. Therefore, by [4, Proposition 8.5] and [4, Proposition 9.1], embeds -equivariantly and homeomorphically into and has countable complement in .
∎
2.2 Combinatorial Geodesic Ray Bundles
Let be a hyperbolic graph equipped with its natural combinatorial metric (assigning edges length 1), and denote the vertex set of by . We present some definitions and terminology used in [15] that we will use in our paper. We refer the reader to Sections 3 and 4 of [15] for a further study of the objects we define in this section.
Definition 2.2.
For and , define to be the set of all combinatorial geodesic rays (CGRs) based at and define the combinatorial geodesic ray bundle to be the set of all vertices on CGRs in .
By [15, Lemma 3.2], every CGR converges to some . We denote such limit .
Definition 2.3.
Fixing a basepoint , for define the limit set .
By [15, Lemma 3.1] (which says that we can move the basepoint of any geodesic ray to any other basepoint to obtain a geodesic with the same tail), the definition of is independent of the basepoint (i.e. for any and , we have for some if and only if for some ).
Definition 2.4.
For and , define the combinatorial sector .
Definition 2.5.
For , a vertex is -special if contains a CGR . The set of all -special vertices is denoted .
By [15, Lemma 4.7], if , then there exists a unique such that . We denote such by .
Our main objects of interest will be the following modified geodesic ray bundles, first defined in [15, Definition 5.5].
Definition 2.6.
Let and . For , let be the set of -special vertices with at minimal distance to . Put
3 Geodesic Ray Bundles in Relatively Hyperbolic Groups
In this section, we examine modified geodesic ray bundles in the relative Cayley graph and prove that these modified bundles have finite symmetric difference for a fixed boundary point. This section generalizes [15, Theorem 5.9].
We begin by showing that is uniformly bounded for each , each and each , which is a well-known property in any uniformly locally finite hyperbolic graph.
We will make use of the following result, which states that geodesic triangles in the relative Cayley graph are slim with respect to the metric for some finite generating set .
Theorem 3.1.
Let be a finitely generated group hyperbolic relative to a collection of subgroups . There exists a finite generating set of such that the following holds. There exists a constant such that for any geodesic triangle in the relative Cayley graph and any vertex on , there exists a vertex on such that .
Proof.
Here is the main result of this section.
Theorem 3.2.
Let be a finitely generated group hyperbolic relative to a collection of subgroups . There exists a finite generating set of such that the following holds. Let be the associated relative Cayley graph. Then there is a constant such that for any , any , and each , we have
Proof.
Take the finite generating set to be as in Theorem 3.1. Let . Let be the constant from Theorem 3.1. Note that is -hyperbolic. Fix any and let . We will show that for each , there exists a vertex on with and such that .
Let be arbitrary. Begin by joining and with a geodesic (see Figure 1). By -hyperbolicity of , we have that , so has length at most .
Letting denote the restriction of a geodesic to , we apply Theorem 3.1 to the geodesic triangle with sides and , By Theorem 3.1, there exists a vertex on or on such that . We cannot have on because then we would have (since ), which would imply by the triangle inequality that
contradicting our choice of . Therefore, we must have that is on .
Lastly, let us show that . By -hyperbolicity, we have , and note that implies , so by the triangle inequality,
We conclude that for each and each , must be -close in to a vertex on with . There are at most such vertices on , so we obtain that . Thus, we set .

∎
As a corollary of Theorem 3.2, we obtain the following:
Theorem 3.3.
Let be a finitely generated group hyperbolic relative to a collection of subgroups . There exists a finite generating set of such that the following holds. If is the associated relative Cayley graph, then is finite for each and each .
Proof.
Let be as in Theorem 3.2. By Theorem 3.2, we have that is uniformly locally finite for each and . In [15, Theorem 5.9], it is proved that if a hyperbolic graph has the property that is uniformly locally finite for each vertex and each , then is finite for each pair of vertices and each . Therefore, is finite for each and each . ∎
Remark 3.4.
Note that Theorem 3.2 implies that if a relatively hyperbolic group is generated by a finite set as in Theorem 3.3, and if the set of ends of the associated relative Cayley graph is the same as , then the ends of have uniformly bounded degree (see [11, Section 2] for the definition of ends and the degree of an end). This appears to not have been known for relative Cayley graphs of relatively hyperbolic groups.
4 Hyperfiniteness of the boundary action
In this section, we establish the hyperfiniteness of the boundary actions of relatively hyperbolic groups as a consequence of Theorem 3.3. Our arguments follow [15, Section 6]. The main difference here is in our coding of labels of geodesics. In this section, we fix a finite generating set for as in Theorem 3.3 and let denote the associated relative Cayley graph of with respect to and .
First, we give a binary coding to the symmetrized generating set . Using that is countably infinite, we fix a bijection from to the set of all finite binary sequences (which we can identify with the set of all finitely supported, infinite binary strings). The label of a geodesic ray is then coded as an element of , the set of all infinite sequences of finite binary strings.
We will need to order elements of (i.e. the set of length sequences of length binary strings) for each . Following [8, Section 7], for each , each and for each with , we put , where is the restriction of the length binary sequence to the first entries. Similarly, if , we put . If we visualize as an matrix, then is an submatrix of the matrix , starting at the top left corner of .
For each , we fix a total order on as in [8, Section 7] such that for all , . Given , we define to be its coded label. Therefore, according to above, denotes the restricted label. Now, analogously to [15, Definition 6.1], we put:
Definition 4.1.
For , define:
Definition 4.2.
An in occurs in if for some . An in occurs infinitely often in if for infinitely many .
Note that for each , there exists which occurs infinitely often in because taking any , by [15, Proposition 5.8], is finite, so there exists some such that for all , . Then and for each . Since is finite, by the Pigeonhole Principle, some must repeat infinitely often in , that is, for infinitely many . For each , we can therefore choose the minimal (in the order defined above) such occuring infinitely often in . We shall denote this element by .
Proposition 4.3.
For each , we have that .
Proof.
Since appears infinitely often in , so does , so or . If , then since there are only finitely many extensions of to an element of and since appears infinitely often in , there would exist such that and appears infinitely often in . Since , we obtain that , contradicting the minimality of . Therefore, . ∎
We now fix a total order on the group such that (for instance, fixing a total order on , we can define to be lexicographic order on elements of as words over , where we choose for each element of the lexicographically least word over representing it). Using the same notation as in [15, Section 6], we put:
Definition 4.4.
For each and , put and put (where the minimum is with respect to the above total order on ). Put for each .
Note that exists because and is locally finite by Theorem 3.2. By definition of and since for each , we have that is a non-increasing sequence of sets and therefore the sequence is a non-decreasing sequence of natural numbers.
We shall now generalize the results of [15, Section 6], which were stated for Cayley graphs of hyperbolic groups. Recall that we fix our hyperbolicity constant to be from Theorem 3.1. We recall that the topology on is the discrete topology induced by the relative metric , the topology on is the canonical topology on the geodesic boundary, having countable neighbourhood base for each , each and each basepoint , has the product topology and has the topology of pointwise convergence.
Let us establish a link between the topology of and sequences of CGRs in . The condition in the following proposition is often used as the definition of the topology on when is a proper hyperbolic space, but in general does not give the same topology on that we work with here.
Proposition 4.5.
Suppose that in . Then for any , there exists a sequence of CGRs such that for each and such that every subsequence of itself has a subsequence which converges to some CGR .
Proof.
Since , by definition of the topology on , we have that for each , there exists a CGR and such that for every . Fixing any , we obtain that for every and every , since for all and hence are close for each . We claim that every subsequence of has a convergent subsequence. First, let us argue as in the proof of Theorem 3.2 to show that for each , is finite.
Given , set . For each , we have . Let denote a geodesic between and (see Figure 2).

Then arguing as in the proof of Theorem 3.2, there exists a vertex on with . It follows that is finite, and therefore that is finite. Therefore, is locally finite. Since is locally finite, by Kőnig’s lemma it follows that every subsequence of has a convergent subsequence. The limit CGR of this subsequence is in because for each , for all but finitely many , so for all .
∎
We now generalize the claims of [15, Section 6] to relatively hyperbolic groups. We begin by generalizing Claim 1 of [15]. In Claim 1 in [15], the set below is proved to be compact, while here it is only closed.
Claim 4.6.
The set is closed. Furthermore, for any and any , the set is compact.
Proof.
Let be a sequence of elements of converging to some . We claim that is a geodesic. Indeed, since , for each , there exists such that for all , we have . In particular, it follows that is a geodesic, since is a geodesic for each . Thus, is a geodesic ray based at and is hence a CGR, so . Therefore, is closed.
The "furthermore" statement follows immediately from Kőnig’s lemma, since is locally finite (by Theorem 3.2).
∎
The next claims are the exact relatively hyperbolic analogues of claims from [15] and their proofs are almost identical (most proofs are completely identical), however, we present all proofs for completeness.
Claim 4.7.
The set is closed in .
Proof.
Suppose that for all and that . Then , in (so that is eventually equal to , by discreteness of ) and in , so that for some (by Claim 4.6). We will show that .
As , by Proposition 4.5, there exists a sequence with which has a subsequence that converges to some . Choose large enough such that all equal , so that . We then have that for each . Taking , we obtain that for all , and therefore that . Thus, with , so and so is closed.
∎
Claim 4.8.
The set is Borel in .
Proof.
Claim 4.9.
The set is Borel in .
Proof.
We follow a similar proof to the proof of [15, Claim 4]. We will show that is both analytic and coanalytic, hence Borel by [13, Theorem 14.11]. By definition of , we have that if and only if
We also have that
which gives a Borel definition of the set . Thus, from Claim 4.7 we have that is analytic. To show that is coanalytic, we will show the following, denoting the -neighbourhood of a subset of :
This formula defines a coanalytic set since there is a single universal quantifier ranging over an uncountable standard Borel space .
For the forward direction, if , then there exists converging to . We simply take (the restriction of from 0 to ) for each . Then for each , we have for each , so for each . Furthermore, since converges to , we have that for all , there exists such that for all , we have .
For the reverse direction, let . Then there exists a sequence of geodesic paths starting at , each contained in and such that . For each , fix and using -hyperbolicity, choose an sufficiently large such that for all , we have
for all . Arguing as in the proof of Theorem 3.2, we have that is finite, so that is finite for each . Therefore, by Kőnig’s lemma, has a subsequence converging to some CGR based at , and , so . From as , we have that . Since , we conclude that .
∎
By [15, Proposition 5.2], for each , we have that the section is finite, having cardinality bounded above by the constant from Theorem 3.2. Since is Borel and has finite sections of size at most , by the Lusin-Novikov theorem we have Borel functions such that is the union of the graphs of the .
Claim 4.10.
For each , is Borel in .
Proof.
By [15, Lemma 4.2], for , denoting the union of all geodesic paths in from to , we have that for some, equivalently any, converging to . From this, we obtain that:
This yields the analyticity (from the above) and coanalyticity (from the above) of , hence Borelness of .
∎
Claim 4.11.
The set is Borel in .
Proof.
We have that if and only if:
Indeed, if , then contains a CGR , so we can take (the restriction from 0 to ) for all to satisfy the above condition.
Conversely, if the above condition holds, then by local finiteness of , the sequence with will have a subsequence converging to some and the above condition yields that , so that .
Since and are Borel, we conclude that is Borel.
∎
Claim 4.12.
The set is Borel in .
Proof.
We have if and only if:
Since is Borel (Claim 4.11), is Borel (as is Borel), and is Borel (Claim 4.10), the above yields that is Borel.
∎
Claim 4.13.
The set is Borel in .
Proof.
We have that if and only if and is the closest element to (in the metric ) such that and . Thus, by Claims 4.9, 4.10, 4.12, is Borel (note that for some , so is Borel in by Claim 4.10).
∎
Claim 4.14.
The set is Borel in .
Proof.
We have that if and only if such that and and . Since and are Borel, it follows that is Borel.
∎
Claim 4.15.
The set is Borel in .
Proof.
Since , we have:
Therefore, is the projection of onto . By Claim 4.14, is Borel. Also, the sections of are finite by Theorem 3.2 and [15, Proposition 5.2]. Therefore, by the Lusin-Novikov theorem, is Borel.
∎
Claim 4.16.
The set is Borel in .
Proof.
We have that if and only if and . By Claim 4.15, is Borel in . Also, is Borel by Claim 4.8. Therefore, is Borel.
∎
Claim 4.17.
For each , the set is Borel in .
Proof.
We have that if and only if is the -minimal element in for which the following holds:
Note that the the "only if" holds by local finiteness of . Thus, is Borel by Claim 4.16.
∎
Now let denote the orbit equivalence relation of the action of on .
Definition 4.18.
Let .
Since is locally finite (as is locally finite and ), we have that is the set of all such that there exists belonging to for all , i.e. for which there exists with label .
Lemma 4.19.
The map given by is a Borel reduction.
Proof.
We argue as in [15]. First, let us show that for each , each and each . If there are infinitely many pairs , then since the left action of on preserves labels of geodesics, there are infinitely many pairs , where for some and where (using [15, Lemma 5.10] in the last line).
By [15, Theorem 5.9], the symmetric difference between and is finite and so there are infinitely many pairs . Hence, there are infinitely many pairs . Thus, as is least in the order that appears infinitely often in , we have that . As for each , we have .
This implies that is constant on -orbits. Indeed, suppose for some , . We have that maps the boundary point to the boundary point . Note that and , because has label and left multiplication preserves labels of geodesics. On the other hand, maps to . We have that and . But by above, . Therefore, and both start at and have the same label. Therefore, they are the same geodesic. Hence, i.e. .
It follows that is reduction to on . Indeed, the above shows that . Conversely, if , then , so , and therefore .
It remains to show that is Borel. To show this, let us first show that the set is Borel. We have if and only if for each , so is Borel by Claim 4.17 (note that the map is continuous, hence Borel, for each ).
Now the Borelness of implies the Borelness of the graph of . Indeed, note that for and , denoting the continuous map , we have:
Putting , we have that is Borel because and are Borel (see Claim 4.7 for the Borelness of ). By above, the graph of is the projection of onto the first two coordinates . For each , the section is finite, being either a singleton or the empty set (because a geodesic ray is uniquely determined by its basepoint and label). Therefore, by the Lusin-Novikov theorem, we have that is Borel. Thus, the graph of is Borel, so is Borel.
∎
Lemma 4.20.
is smooth on the saturation .
Proof.
By Lemma 4.19, is smooth on , yielding the claim. ∎
Definition 4.21.
Let . For each , define by . Let be the equivalence relation on which is the restriction of the shift action of on to .
The following lemma is a generalization of [15, Lemma 6.7].
Lemma 4.22.
There exists a constant such that for each , each equivalence class of has size at most .
Proof.
Note that by Thereom 3.2, we have that each closed ball of radius in has cardinality at most , where is the constant from Theorem 3.2. We will show that we can take .
Let and suppose that . By the proof of [15, Lemma 6.7] (which only relies on the hyperbolicity of the Cayley graph and local finiteness of geodesic ray bundles and so holds in our context when applied to ), we have . For completeness, let us reproduce this proof.
By defiinition, (resp. ) is an infinite subset of (resp. ). Since is locally finite, this means that (resp. ) uniquely determines (resp. ). From , we have and since and determine their boundary points, this implies that . Let us denote by the common boundary point .

We have that , so there exists passing through and passing through . Write and for some . Note that by -hyperbolicity, we have . Also, we have . Indeed, since , we have:
where holds by -minimality of in .
Similarly, from , we have , and so there exists passing through and passing through . Write and for some (see Figure 3). By -hyperbolicity, we have and because and so by the -minimality of in , we have that:
Let us now consider the sub-CGRs of and starting at . Using -hyperbolicity, since , there exists such that . Then by the triangle inequality and our above estimates, we have:
Therefore,
where we have because .
Thus, implies that is in the ball of radius about in , which has cardinality at most . Thus, -classes have cardinality at most .
∎
The following remaining results have the same proof as in [15].
Lemma 4.23.
Let . Then the map is Borel and so is analytic.
Proof.
Using [15, Lemma 2.3], there exists a finite Borel equivalence relation on with . Since is finite, Borel, there exists a Borel reduction from to for each , using which we define by . Put , i.e. .
Lemma 4.24.
The equivalence relation is a hyperfinite countable Borel equivalence relation.
Proof.
Since is Borel, we have that is Borel. We also have that is hypersmooth by definition, and so it is hyperfinite by [9, Theorem 8.1.5]. We follow the same proof as the proof of [15, Lemma 6.9], to show that is countable.
For each , define the relation on by if for all . Each is countable because if , then , and is countable-to-one since is countable-to-one because if , then and is countable, and is finite-to-one since is finite. Therefore, there are only countably many choices for such that once is fixed. Thus, is countable. Noting that , we obtain that is countable.
∎
Lemma 4.25.
is a homomorphism from to .
Proof.
Suppose are -related, as witnessed by (so ). By [15, Theorem 5.9] and [15, Lemma 3.10], we have that and differ by a finite set. By local finiteness of and since , we have that there exists such that for all . By the proof of Lemma 4.19, we have , which gives, together with , that . This then yields for all . Thus, we have and so for all since . Therefore, for all and so .
∎
Proof of Theorem A.
Note that is a sub-relation of . Indeed, if and , then by Lemma 4.25, we have that , which implies . By Lemma 4.24, we have that is hyperfinite, so is hyperfinite, since a sub-relation of a hyperfinite equivalence relation is hyperfinite. On , is smooth by Lemma 4.20, and hence hyperfinite. Therefore, is hyperfinite on .
Recall that we worked with a fixed a finite generating set in this section. If we use a different finite generating set for , then the relative Cayley graph corresponding to is -equivariantly quasi-isometric to (via the identity map on ), so is -equivariantly homeomorphic to . It follows that the orbit equivalence relation of on is also hyperfinite.
As a corollary, we obtain Corollary B on the hyperfiniteness of the action of on , where is the collection of parabolic subgroups.
Proof of Corollary B.
By Theorem 2.1, embeds -equivariantly and topologically into with countable complement. Therefore, the orbit equivalence relation of on is a subrelation of the orbit equivalence relation of on . Since the orbit equivalence relation of on is hyperfinite (by Theorem A) and since is countable, it follows that the orbit equivalence relation of on is also hyperfinite.
References
- [1] S. Adams. Boundary amenability for word hyperbolic groups and an application to smooth dynamics of simple groups. Topology, 33(4):765–783, 1994.
- [2] S. Adams, G. Elliott, and T. Giordano. Amenable actions of groups. Trans. Amer. Math. Soc., 344(2):803–822, 1994.
- [3] Claire Anantharaman-Delaroche and Jean Renault. Amenable groupoids. Monographies de L’Enseignement Mathematique [Monographs of L’Enseignement Mathematique], 36, 2000.
- [4] B. Bowditch. Relatively hyperbolic groups. International Journal of Algebra and Computation, 22(03):1250016, 2012.
- [5] A. Connes, J. Feldman, and B. Weiss. An amenable equivalence relation is generated by a single transformation. Ergodic Theory and Dynamical Systems, 1(4):431–450, 1981.
- [6] M. Dadarlat and E. Guentner. Uniform embeddability of relatively hyperbolic groups. Journal für die reine und angewandte Mathematik, 612(2007):1–15, 2007.
- [7] François Dahmani. Les groupes relativement hyperboliques et leurs bords. Prépublication de l’Institut de Recherche Mathématique Avancée [Prepublication of the Institute of Advanced Mathematical Research], 2003.
- [8] R. Dougherty, S. Jackson, and A. S. Kechris. The structure of hyperfinite borel equivalence relations. Trans. Amer. Math. Soc., 341(1), 1994.
- [9] Su Gao. Invariant Descriptive Set Theory, volume 293 of Pure and Applied Mathematics. CRC Press, 2009.
- [10] M. Gromov. Hyperbolic groups. Essays in group theory, Math. Sci. Res. Inst. Publ., 8:75–263, 1987.
- [11] Matthias Hamann, Florian Lehner, Babak Miraftab, and Tim Ruhmann. A Stallings’ type theorem for quasi-transitive graphs. J. Combin. Theory, Series B, 157:40–69, 2022.
- [12] Jingyin Huang, Marcin Sabok, and Forte Shinko. Hyperfiniteness of boundary actions of cubulated hyperbolic groups. Ergodic Theory and Dynamical Systems, 40(9):2453–2466, Mar 2019.
- [13] Alexander Kechris. Classical Descriptive Set Theory, volume 156 of Graduate Texts in Mathematics. Springer-Verlag, 1995.
- [14] Timothée Marquis. On geodesic ray bundles in buildings. Geometriae Dedicata, 202(1):27–43, Oct 2018.
- [15] Timothée Marquis and Marcin Sabok. Hyperfiniteness of boundary actions of hyperbolic groups. Mathematische Annalen, 377(3-4):1129–1153, Jun 2020.
- [16] Denis Osin. Asymptotic dimension of relatively hyperbolic groups. International Mathematics Research Notices, 35(2005):2143–2161, 2005.
- [17] Denis Osin. Relatively hyperbolic groups: Intrinsic geometry, algebraic properties, and algorithmic problems, volume 179. Memoirs American Mathematical Society, 2006.
- [18] Narutaka Ozawa. Boundary amenability of relatively hyperbolic groups. Topology and its Applications, 153:2624–2630, 2006.
- [19] Piotr Przytycki and Marcin Sabok. Unicorn paths and hyperfiniteness for the mapping class group. Forum of Mathematics, Sigma, 9:e36, 2021.
- [20] Nicholas W. M. Touikan. On geodesic ray bundles in hyperbolic groups. Proceedings of the American Mathematical Society, 2018.
- [21] A. Vershik. The action of in is approximable. Uspehi Mat. Nauk, 199(1):209–210, 1978.
- [22] R. Zimmer. Amenable ergodic group actions and an application to poisson boundaries of random walks. J. Functional Analysis, 27(3):350–372, 1978.