This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Hyperfiniteness of boundary actions of relatively hyperbolic groups

Chris Karpinski111McGill University. Email: [email protected].
Abstract

We show that if GG is a finitely generated group hyperbolic relative to a finite collection of subgroups 𝒫\mathcal{P}, then the natural action of GG 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 GG on its Bowditch boundary (G,𝒫)\partial(G,\mathcal{P}) also induces a hyperfinite equivalence relation. This strengthens a result of Ozawa obtained for 𝒫\mathcal{P} 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 μ\mu-hyperfinite for every Borel quasi-invariant probability measure μ\mu 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 GG with a collection of parabolic subgroups 𝒫\mathcal{P} there is a natural boundary called the Bowditch boundary, denoted (G,𝒫)\partial(G,\mathcal{P}), which is a compact metrizable space on which GG 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 GG in 𝒫\mathcal{P} are amenable, Ozawa [18] proved that the action of GG on (G,𝒫)\partial(G,\mathcal{P}) is topologically amenable, and, more generally, when the parabolic subgroups are exact, Ozawa [18] proved that the group GG 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 GG on a standard probability space (X,μ)(X,\mu) is μ\mu-amenable if and only if μ\mu-almost all stabilizers are amenable and the orbit equivalence relation is μ\mu-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 GG be a finitely generated group hyperbolic relative to a finite collection of subgroups 𝒫\mathcal{P} and let Γ^\hat{\Gamma} be the associated relative Cayley graph. Then the natural action of GG on the geodesic boundary Γ^\partial\hat{\Gamma} induces a hyperfinite orbit equivalence relation.

Corollary B.

Let GG be a finitely generated group hyperbolic relative to a finite collection of subgroups 𝒫\mathcal{P}. Then the natural action of GG on the Bowditch boundary (G,𝒫)\partial(G,\mathcal{P}) 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 μ\mu-amenability for every invariant Borel probability measure μ\mu 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 μ\mu-amenable for every invariant Borel probability measure μ\mu.

We proceed by following a similar approach to [12] and [15], studying geodesic ray bundles Geo(x,η)\text{Geo}(x,\eta) in relative Cayley graphs (Definition 2.2). For the case of a cubulating hyperbolic group GG studied in [12], the crucial property from which the hyperfinitess of the boundary action of GG follows is the finite symmetric difference of geodesic ray bundles: for any x,yGx,y\in G and any ηG\eta\in\partial G, Geo(x,η)Geo(y,η)\text{Geo}(x,\eta)\triangle\text{Geo}(y,\eta) 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 Geo1(x,η)\text{Geo}_{1}(x,\eta) for xGx\in G and ηG\eta\in\partial G (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: |Geo1(x,η)Geo1(y,η)|<|\text{Geo}_{1}(x,\eta)\triangle\text{Geo}_{1}(y,\eta)|<\infty for each x,yGx,y\in G and for each ηG\eta\in\partial G. 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 Geo1\text{Geo}_{1} 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 XX, X\partial X will denote the geodesic boundary of XX. We will also denote Chb(X)C_{hb}(X) the horoboundary of XX (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 GG be a group generated by a finite set XX, let 𝒫={H1,,Hn}\mathcal{P}=\{H_{1},...,H_{n}\} be a collection of subgroups of GG and let =𝒫\mathcal{H}=\bigcup\mathcal{P}. The relative Cayley graph associated to XX and 𝒫\mathcal{P} is the Cayley graph Γ^\hat{\Gamma} with respect to the generating set XX\cup\mathcal{H}. This graph can be identified with the coned-off Cayley graph obtained by starting with the Cayley graph Γ\Gamma of GG with respect to XX, adjoining to Γ\Gamma a vertex vgHiv_{gH_{i}} for each left coset gHigH_{i} and connecting each vertex of gHigH_{i} in Γ\Gamma to vgHiv_{gH_{i}} by an edge of length 12\frac{1}{2}. The notation dXd_{X} and dd refer to the word metrics with respect to the generating sets XX and XX\cup\mathcal{H}, respectively. We will use the notation BrX(x)B_{r}^{X}(x) to denote the closed ball of radius rr in the metric dXd_{X} about the point xGx\in G.

A finitely generated group GG is hyperbolic relative to a collection of subgroups 𝒫={H1,,Hn}\mathcal{P}=\{H_{1},...,H_{n}\} if there exists a finite generating set XX of GG 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 GG hyperbolic relative to a finite collection 𝒫\mathcal{P} of subgroups, there is a natural compact metrizable space on which GG acts naturally by homeomorphisms, denoted (G,𝒫)\partial(G,\mathcal{P}) 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 GG be hyperbolic relative to a finite collection of subgroups 𝒫\mathcal{P}, with relative Cayley graph Γ^\hat{\Gamma}. Then Γ^\partial\hat{\Gamma} embeds GG-equivariantly and homeomorphically into (G,𝒫)\partial(G,\mathcal{P}) with countable complement.

Proof.

In [7, Proposition 1, Section A.2], it is shown that the coned-off Cayley graph Γ^\hat{\Gamma} witnesses the relative hyperbolicity of GG with respect to 𝒫\mathcal{P} according to Definition 2 of relative hyperbolicity from [4]. Therefore, by [4, Proposition 8.5] and [4, Proposition 9.1], Γ^\partial\hat{\Gamma} embeds GG-equivariantly and homeomorphically into (G,𝒫)\partial(G,\mathcal{P}) and Γ^\partial\hat{\Gamma} has countable complement in (G,𝒫)\partial(G,\mathcal{P}).

2.2 Combinatorial Geodesic Ray Bundles

Let XX be a hyperbolic graph equipped with its natural combinatorial metric (assigning edges length 1), and denote the vertex set of XX by X(0)X^{(0)}. 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 xX(0)x\in X^{(0)} and ηX\eta\in\partial X, define CGR(x,η)CGR(x,\eta) to be the set of all combinatorial geodesic rays (CGRs) based at xx and define the combinatorial geodesic ray bundle Geo(x,η)=CGR(x,η)\text{Geo}(x,\eta)=\bigcup\text{CGR}(x,\eta) to be the set of all vertices on CGRs in CGR(x,η)\text{CGR}(x,\eta).

By [15, Lemma 3.2], every CGR γ=(xn)n\gamma=(x_{n})_{n} converges to some ξChb(Γ^)\xi\in C_{hb}(\hat{\Gamma}). We denote such limit ξ=ξγ\xi=\xi_{\gamma}.

Definition 2.3.

Fixing a basepoint zX(0)z\in X^{(0)}, for ηΓ^\eta\in\partial\hat{\Gamma} define the limit set Ξ(η)={ξγ:γCGR(z,η)}\Xi(\eta)=\{\xi_{\gamma}:\gamma\in\text{CGR}(z,\eta)\}.

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 Ξ(η)\Xi(\eta) is independent of the basepoint (i.e. for any z1,z2X(0)z_{1},z_{2}\in X^{(0)} and ξΞ(η)\xi\in\Xi(\eta), we have ξ=ξγ\xi=\xi_{\gamma} for some γCGR(z1,η)\gamma\in\text{CGR}(z_{1},\eta) if and only if ξ=ξγ\xi=\xi_{\gamma^{\prime}} for some γCGR(z2,η)\gamma^{\prime}\in\text{CGR}(z_{2},\eta)).

Definition 2.4.

For xX(0),ηXx\in X^{(0)},\eta\in\partial X and ξΞ(η)\xi\in\Xi(\eta), define the combinatorial sector Q(x,ξ)={yX(0):yγ for some γCGR(x,η) with ξγ=ξ}Q(x,\xi)=\{y\in X^{(0)}:y\in\gamma\text{ for some }\gamma\in\text{CGR}(x,\eta)\text{ with }\xi_{\gamma}=\xi\}.

Definition 2.5.

For ηX\eta\in\partial X, a vertex xX(0)x\in X^{(0)} is η\mathbf{\eta}-special if ξΞ(η)Q(x,ξ)\bigcap_{\xi\in\Xi(\eta)}Q(x,\xi) contains a CGR γ\gamma. The set of all η\mathbf{\eta}-special vertices is denoted Xs,ηX_{s,\eta}.

By [15, Lemma 4.7], if xXs,ηx\in X_{s,\eta}, then there exists a unique ξΞ(η)\xi^{\prime}\in\Xi(\eta) such that ξΞ(η)Q(x,ξ)=Q(x,ξ)\bigcap_{\xi\in\Xi(\eta)}Q(x,\xi)=Q(x,\xi^{\prime}). We denote such ξ\xi^{\prime} by ξ=ξx,η\xi^{\prime}=\xi_{x,\eta}.

Our main objects of interest will be the following modified geodesic ray bundles, first defined in [15, Definition 5.5].

Definition 2.6.

Let xX(0)x\in X^{(0)} and ηX\eta\in\partial X. For ξΞ(η)\xi\in\Xi(\eta), let Y(x,ξ)Y(x,\xi) be the set of η\eta-special vertices yGeo(x,η)y\in\text{Geo}(x,\eta) with ξy,η=ξ\xi_{y,\eta}=\xi at minimal distance to xx. Put

Geo1(x,η)=ξΞ(η)yY(x,ξ)Q(y,ξ)\text{Geo}_{1}(x,\eta)=\bigcup_{\xi\in\Xi(\eta)}\bigcup_{y\in Y(x,\xi)}Q(y,\xi)

3 Geodesic Ray Bundles in Relatively Hyperbolic Groups

In this section, we examine modified geodesic ray bundles in the relative Cayley graph Γ^\hat{\Gamma} 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 |{γ(i):γCGR(x,η)}||\{\gamma(i):\gamma\in\text{CGR}(x,\eta)\}| is uniformly bounded for each ii, each xGx\in G and each ηΓ^\eta\in\partial\hat{\Gamma}, 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 dXd_{X} for some finite generating set XX.

Theorem 3.1.

Let GG be a finitely generated group hyperbolic relative to a collection of subgroups {H1,,Hn}\{H_{1},...,H_{n}\}. There exists a finite generating set XX of GG such that the following holds. There exists a constant ν\nu such that for any geodesic triangle pqrpqr in the relative Cayley graph Γ^\hat{\Gamma} and any vertex uu on pp, there exists a vertex vv on qrq\cup r such that dX(u,v)νd_{X}(u,v)\leq\nu.

Proof.

The finite generating set XX is constructed in the proof of [17, Lemma 3.1] and it is shown in the proof of [17, Theorem 3.26] that XX satisfies the stated property. ∎

Here is the main result of this section.

Theorem 3.2.

Let GG be a finitely generated group hyperbolic relative to a collection of subgroups {H1,,Hn}\{H_{1},...,H_{n}\}. There exists a finite generating set XX of GG such that the following holds. Let Γ^\hat{\Gamma} be the associated relative Cayley graph. Then there is a constant BB such that for any xGx\in G, any ηΓ^\eta\in\partial\hat{\Gamma}, and each ii\in\mathbb{N}, we have

|{γ(i):γCGR(x,η)}|B|\{\gamma(i):\gamma\in\text{CGR}(x,\eta)\}|\leq B
Proof.

Take the finite generating set XX to be as in Theorem 3.1. Let ii\in\mathbb{N}. Let ν\nu be the constant from Theorem 3.1. Note that Γ^\hat{\Gamma} is ν\nu-hyperbolic. Fix any γ0CGR(x,η)\gamma_{0}\in CGR(x,\eta) and let k=i+3ν+1k=i+3\nu+1. We will show that for each γCGR(x,η)\gamma\in\text{CGR}(x,\eta), there exists a vertex vv on γ0\gamma_{0} with d(v,γ0(i))3νd(v,\gamma_{0}(i))\leq 3\nu and such that dX(γ(i),v)νd_{X}(\gamma(i),v)\leq\nu.

Let γCGR(x,η)\gamma\in CGR(x,\eta) be arbitrary. Begin by joining γ(k)\gamma(k) and γ0(k)\gamma_{0}(k) with a geodesic α\alpha (see Figure 1). By ν\nu-hyperbolicity of Γ^\hat{\Gamma}, we have that d(γ(k),γ0(k))2νd(\gamma(k),\gamma_{0}(k))\leq 2\nu, so α\alpha has length (α)\ell(\alpha) at most 2ν2\nu.

Letting |k|_{k} denote the restriction of a geodesic to {0,1,,k}\{0,1,...,k\}, we apply Theorem 3.1 to the geodesic triangle with sides γ0|k,α\gamma_{0}|_{k},\alpha and γ|k\gamma|_{k}, By Theorem 3.1, there exists a vertex vv on γ0|k\gamma_{0}|_{k} or on α\alpha such that dX(γ(i),v)νd_{X}(\gamma(i),v)\leq\nu. We cannot have vv on α\alpha because then we would have d(γ(i),v)νd(\gamma(i),v)\leq\nu (since ddXd\leq d_{X}), which would imply by the triangle inequality that

ki=d(γ(i),γ(k))d(γ(i),v)+d(v,γ(k))d(γ(i),v)+(α)ν+2ν=3νk-i=d(\gamma(i),\gamma(k))\leq d(\gamma(i),v)+d(v,\gamma(k))\leq d(\gamma(i),v)+\ell(\alpha)\leq\nu+2\nu=3\nu

contradicting our choice of kk. Therefore, we must have that vv is on γ0|k\gamma_{0}|_{k}.

Lastly, let us show that d(v,γ0(i))3νd(v,\gamma_{0}(i))\leq 3\nu. By ν\nu-hyperbolicity, we have d(γ(i),γ0(i))2νd(\gamma(i),\gamma_{0}(i))\leq 2\nu, and note that dX(γ(i),v)νd_{X}(\gamma(i),v)\leq\nu implies d(γ(i),v)νd(\gamma(i),v)\leq\nu, so by the triangle inequality,

d(v,γ0(i))d(v,γ(i))+d(γ(i),γ0(i))ν+2ν=3νd(v,\gamma_{0}(i))\leq d(v,\gamma(i))+d(\gamma(i),\gamma_{0}(i))\leq\nu+2\nu=3\nu

We conclude that for each ii\in\mathbb{N} and each γCGR(x,η)\gamma\in\text{CGR}(x,\eta), γ(i)\gamma(i) must be ν\nu-close in dXd_{X} to a vertex vv on γ0\gamma_{0} with d(v,γ0(i))3νd(v,\gamma_{0}(i))\leq 3\nu. There are at most 6ν+16\nu+1 such vertices on γ0\gamma_{0}, so we obtain that |{γ(i):γCGR(x,η)}|(6ν+1)|BXν(1)||\{\gamma(i):\gamma\in\text{CGR}(x,\eta)\}|\leq(6\nu+1)|B_{X}^{\nu}(1)|. Thus, we set B=(6ν+1)|BXν(1)|B=(6\nu+1)|B_{X}^{\nu}(1)|.

Refer to caption
Figure 1: The arrangement of geodesics in the proof of Theorem 3.2.

As a corollary of Theorem 3.2, we obtain the following:

Theorem 3.3.

Let GG be a finitely generated group hyperbolic relative to a collection of subgroups {H1,,Hn}\{H_{1},...,H_{n}\}. There exists a finite generating set XX of GG such that the following holds. If Γ^\hat{\Gamma} is the associated relative Cayley graph, then Geo1(x,η)Geo1(y,η)\text{Geo}_{1}(x,\eta)\triangle\text{Geo}_{1}(y,\eta) is finite for each x,yGx,y\in G and each ηΓ^\eta\in\partial\hat{\Gamma}.

Proof.

Let XX be as in Theorem 3.2. By Theorem 3.2, we have that Geo(x,η)\text{Geo}(x,\eta) is uniformly locally finite for each xGx\in G and ηΓ^\eta\in\partial\hat{\Gamma}. In [15, Theorem 5.9], it is proved that if a hyperbolic graph Γ\Gamma has the property that Geo(x,η)\text{Geo}(x,\eta) is uniformly locally finite for each vertex xx and each ηΓ\eta\in\partial\Gamma, then Geo1(x,η)Geo1(y,η)\text{Geo}_{1}(x,\eta)\triangle\text{Geo}_{1}(y,\eta) is finite for each pair of vertices x,yx,y and each ηΓ\eta\in\partial\Gamma. Therefore, Geo1(x,η)Geo1(y,η)\text{Geo}_{1}(x,\eta)\triangle\text{Geo}_{1}(y,\eta) is finite for each x,yGx,y\in G and each ηΓ^\eta\in\partial\hat{\Gamma}. ∎

Remark 3.4.

Note that Theorem 3.2 implies that if a relatively hyperbolic group GG is generated by a finite set XX as in Theorem 3.3, and if the set of ends of the associated relative Cayley graph is the same as Γ^\partial\hat{\Gamma}, then the ends of Γ^\hat{\Gamma} 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 XX for GG as in Theorem 3.3 and let Γ^\hat{\Gamma} denote the associated relative Cayley graph of GG with respect to {H1,,Hn}\{H_{1},...,H_{n}\} and XX.

First, we give a binary coding to the symmetrized generating set S:=(X)±S:=(X\cup\mathcal{H})^{\pm}. Using that SS is countably infinite, we fix a bijection f:S2<:=n2nf:S\to 2^{<\mathbb{N}}:=\bigcup_{n\in\mathbb{N}}2^{n} from SS to the set 2<2^{<\mathbb{N}} 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 (2<)(2^{<\mathbb{N}})^{\mathbb{N}}, the set of all infinite sequences of finite binary strings.

We will need to order elements of (2n)n(2^{n})^{n} (i.e. the set of length nn sequences of length nn binary strings) for each nn. Following [8, Section 7], for each m1,m2m_{1},m_{2}\in\mathbb{N}, each w=(w0,w1,,wn1)(2m1)m2w=(w_{0},w_{1},...,w_{n-1})\in(2^{m_{1}})^{m_{2}} and for each nn\in\mathbb{N} with nm1,m2n\leq m_{1},m_{2}, we put w|n=((w0)|n,(w1)|n,,(wn1)|n)w|_{n}=((w_{0})|_{n},(w_{1})|_{n},...,(w_{n-1})|_{n}), where (wj)|n(w_{j})|_{n} is the restriction of the length m1m_{1} binary sequence wjw_{j} to the first nn entries. Similarly, if w(2)w\in(2^{\mathbb{N}})^{\mathbb{N}}, we put w|n=((w0)|n,(w1)|n,,(wn1)|n)w|_{n}=((w_{0})|_{n},(w_{1})|_{n},...,(w_{n-1})|_{n}). If we visualize w(2n)nw\in(2^{n})^{n} as an n×nn\times n matrix, then w|iw|_{i} is an i×ii\times i submatrix of the n×nn\times n matrix ww, starting at the top left corner of ww.

For each nn\in\mathbb{N}, we fix a total order <n<_{n} on (2n)n(2^{n})^{n} as in [8, Section 7] such that for all w,v(2n+1)n+1w,v\in(2^{n+1})^{n+1}, w|n<nv|nw<n+1vw|_{n}<_{n}v|_{n}\implies w<_{n+1}v. Given γCGR(g,η)\gamma\in\text{CGR}(g,\eta), we define lab(γ)(2<)\text{lab}(\gamma)\in(2^{<\mathbb{N}})^{\mathbb{N}} to be its coded label. Therefore, according to above, lab(γ)|n(2n)n\text{lab}(\gamma)|_{n}\in(2^{n})^{n} denotes the restricted label. Now, analogously to [15, Definition 6.1], we put:

Definition 4.1.

For ηΓ^\eta\in\partial\hat{\Gamma}, define:

Cη={(g,lab(γ)|n)G×(2n)n:gGeo1(e,η),γCGR(g,η),n}C^{\eta}=\{(g,\text{lab}(\gamma)|_{n})\in G\times(2^{n})^{n}:g\in Geo_{1}(e,\eta),\gamma\in CGR(g,\eta),n\in\mathbb{N}\}
Definition 4.2.

An ss in (2n)n(2^{n})^{n} occurs in CηC^{\eta} if (g,s)Cη(g,s)\in C^{\eta} for some gGeo1(e,η)g\in Geo_{1}(e,\eta). An ss in (2n)n(2^{n})^{n} occurs infinitely often in CηC^{\eta} if (g,s)Cη(g,s)\in C^{\eta} for infinitely many gGeo1(e,η)g\in Geo_{1}(e,\eta).

Note that for each nn\in\mathbb{N}, there exists s(2n)ns\in(2^{n})^{n} which occurs infinitely often in CηC^{\eta} because taking any γCGR(e,η)\gamma\in\text{CGR}(e,\eta), by [15, Proposition 5.8], γGeo1(e,η)\gamma\setminus\text{Geo}_{1}(e,\eta) is finite, so there exists some NN such that for all kNk\geq N, γ(k)Geo1(e,η)\gamma(k)\in\text{Geo}_{1}(e,\eta). Then (γ(k),lab((γ(i))ik)|n)Cη(\gamma(k),\text{lab}((\gamma(i))_{i\geq k})|_{n})\in C^{\eta} and lab((γ(i))ik)|n(2n)n\text{lab}((\gamma(i))_{i\geq k})|_{n}\in(2^{n})^{n} for each kNk\geq N. Since (2n)n(2^{n})^{n} is finite, by the Pigeonhole Principle, some s(2n)ns\in(2^{n})^{n} must repeat infinitely often in CηC^{\eta}, that is, (γ(k),s)Cη(\gamma(k),s)\in C^{\eta} for infinitely many kNk\geq N. For each nn\in\mathbb{N}, we can therefore choose the minimal (in the order <n<_{n} defined above) such s(2n)ns\in(2^{n})^{n} occuring infinitely often in CηC^{\eta}. We shall denote this element by snηs_{n}^{\eta}.

Proposition 4.3.

For each nn\in\mathbb{N}, we have that (sn+1η)|n=snη(s_{n+1}^{\eta})|_{n}=s_{n}^{\eta}.

Proof.

Since sn+1ηs_{n+1}^{\eta} appears infinitely often in CηC^{\eta}, so does (sn+1η)|n(s_{n+1}^{\eta})|_{n}, so snη<n(sn+1η)|ns_{n}^{\eta}<_{n}(s_{n+1}^{\eta})|_{n} or snη=(sn+1η)|ns_{n}^{\eta}=(s_{n+1}^{\eta})|_{n}. If snη<n(sn+1η)|ns_{n}^{\eta}<_{n}(s_{n+1}^{\eta})|_{n}, then since there are only finitely many extensions of snηs_{n}^{\eta} to an element of (2n+1)n+1(2^{n+1})^{n+1} and since snηs_{n}^{\eta} appears infinitely often in CηC^{\eta}, there would exist s(2n+1)n+1s\in(2^{n+1})^{n+1} such that s|n=snηs|_{n}=s_{n}^{\eta} and ss appears infinitely often in CηC^{\eta}. Since s|n<n(sn+1η)|ns|_{n}<_{n}(s_{n+1}^{\eta})|_{n}, we obtain that s<n+1sn+1ηs<_{n+1}s_{n+1}^{\eta}, contradicting the minimality of sn+1ηs_{n+1}^{\eta}. Therefore, snη=(sn+1η)|ns_{n}^{\eta}=(s_{n+1}^{\eta})|_{n}. ∎

We now fix a total order \leq on the group GG such that ghd(e,g)d(e,h)g\leq h\implies d(e,g)\leq d(e,h) (for instance, fixing a total order on SS, we can define \leq to be lexicographic order on elements of GG as words over SS, where we choose for each element of GG the lexicographically least word over SS representing it). Using the same notation as in [15, Section 6], we put:

Definition 4.4.

For each nn\in\mathbb{N} and ηΓ^\eta\in\partial\hat{\Gamma}, put Tnη={gGeo1(e,η):(g,snη)Cη}T_{n}^{\eta}=\{g\in Geo_{1}(e,\eta):(g,s_{n}^{\eta})\in C^{\eta}\} and put gnη=minTnηg_{n}^{\eta}=\min T_{n}^{\eta} (where the minimum is with respect to the above total order on GG). Put knη=d(e,gnη)k_{n}^{\eta}=d(e,g_{n}^{\eta}) for each nn\in\mathbb{N}.

Note that minTnη\min T_{n}^{\eta} exists because TnηGeo(e,η)T_{n}^{\eta}\subseteq\text{Geo}(e,\eta) and Geo(e,η)\text{Geo}(e,\eta) is locally finite by Theorem 3.2. By definition of \leq and since snη=(sn+1η)|ns_{n}^{\eta}=(s_{n+1}^{\eta})|_{n} for each nn, we have that (Tnη)n(T_{n}^{\eta})_{n} is a non-increasing sequence of sets and therefore the sequence (knη)n(k_{n}^{\eta})_{n\in\mathbb{N}} 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 ν\nu from Theorem 3.1. We recall that the topology on GG is the discrete topology induced by the relative metric dd, the topology on Γ^\partial\hat{\Gamma} is the canonical topology on the geodesic boundary, having countable neighbourhood base V(η,m)g={μΓ^:γCGR(g,μ) and λCGR(g,η) with d(γ(t),λ(t))2ν for each tm}V(\eta,m)^{g}=\{\mu\in\partial\hat{\Gamma}:\exists\gamma\in\text{CGR}(g,\mu)\text{ and }\lambda\in\text{CGR}(g,\eta)\text{ with }d(\gamma(t),\lambda(t))\leq 2\nu\text{ for each }t\leq m\} for each mm\in\mathbb{N}, each ηΓ^\eta\in\partial\hat{\Gamma} and each basepoint gGg\in G, GG^{\mathbb{N}} has the product topology and Chb(Γ^)C_{hb}(\hat{\Gamma}) has the topology of pointwise convergence.

Let us establish a link between the topology of Γ^\partial\hat{\Gamma} and sequences of CGRs in Γ^\hat{\Gamma}. The condition in the following proposition is often used as the definition of the topology on X\partial X when XX is a proper hyperbolic space, but in general does not give the same topology on X\partial X that we work with here.

Proposition 4.5.

Suppose that ηnη\eta_{n}\to\eta in Γ^\partial\hat{\Gamma}. Then for any gGg\in G, there exists a sequence of CGRs (γn)n(\gamma_{n})_{n} such that γnCGR(g,ηn)\gamma_{n}\in CGR(g,\eta_{n}) for each nn and such that every subsequence of (γn)n(\gamma_{n})_{n} itself has a subsequence which converges to some CGR γCGR(g,η)\gamma\in CGR(g,\eta).

Proof.

Since ηnη\eta_{n}\to\eta, by definition of the topology on Γ^\partial\hat{\Gamma}, we have that for each mm\in\mathbb{N}, there exists a CGR γmCGR(g,ηm)\gamma_{m}\in\text{CGR}(g,\eta_{m}) and λmCGR(g,η)\lambda_{m}\in\text{CGR}(g,\eta) such that d(γm(t),λm(t))2νd(\gamma_{m}(t),\lambda_{m}(t))\leq 2\nu for every tmt\leq m. Fixing any λCGR(g,η)\lambda\in\text{CGR}(g,\eta), we obtain that d(γm(t),λ(t))4νd(\gamma_{m}(t),\lambda(t))\leq 4\nu for every tmt\leq m and every mm, since λm,λCGR(g,η)\lambda_{m},\lambda\in\text{CGR}(g,\eta) for all mm and hence are 2ν2\nu close for each mm. We claim that every subsequence of (γn)n(\gamma_{n})_{n} has a convergent subsequence. First, let us argue as in the proof of Theorem 3.2 to show that for each ii, |{γn(i):n}||\{\gamma_{n}(i):n\in\mathbb{N}\}| is finite.

Given ii\in\mathbb{N}, set k=i+5ν+1k=i+5\nu+1. For each nkn\geq k, we have d(γn(k),λ(k))4νd(\gamma_{n}(k),\lambda(k))\leq 4\nu. Let uu denote a geodesic between γn(k)\gamma_{n}(k) and λ(k)\lambda(k) (see Figure 2).

Refer to caption
Figure 2: The geometry of the geodesics γn\gamma_{n}, λ\lambda.

Then arguing as in the proof of Theorem 3.2, there exists a vertex vv on λ\lambda with dX(γn(i),v)νd_{X}(\gamma_{n}(i),v)\leq\nu. It follows that |{γn(i):nk}||\{\gamma_{n}(i):n\geq k\}| is finite, and therefore that |{γn(i):n}||\{\gamma_{n}(i):n\in\mathbb{N}\}| is finite. Therefore, nγnλ\bigcup_{n}\gamma_{n}\cup\lambda is locally finite. Since nγnλ\bigcup_{n}\gamma_{n}\cup\lambda is locally finite, by Kőnig’s lemma it follows that every subsequence of (γn)n(\gamma_{n})_{n} has a convergent subsequence. The limit CGR γ\gamma of this subsequence is in CGR(g,η)\text{CGR}(g,\eta) because for each tt, d(γk(t),λ(t))4νd(\gamma_{k}(t),\lambda(t))\leq 4\nu for all but finitely many kk, so d(γ(t),λ(t))4νd(\gamma(t),\lambda(t))\leq 4\nu for all tt.

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 CC below is proved to be compact, while here it is only closed.

Claim 4.6.

The set C={γG:γ is a CGR}C=\{\gamma\in G^{\mathbb{N}}:\gamma\text{ is a CGR}\} is closed. Furthermore, for any gGg\in G and any ηΓ^\eta\in\partial\hat{\Gamma}, the set CGR(g,η)GCGR(g,\eta)\subseteq G^{\mathbb{N}} is compact.

Proof.

Let (γn)n(\gamma_{n})_{n} be a sequence of elements of CC converging to some γG\gamma\in G^{\mathbb{N}}. We claim that γ\gamma is a geodesic. Indeed, since γnγ\gamma_{n}\to\gamma, for each mm\in\mathbb{N}, there exists NN\in\mathbb{N} such that for all nNn\geq N, we have γn|m=γ|m\gamma_{n}|_{m}=\gamma|_{m}. In particular, it follows that γ|m\gamma|_{m} is a geodesic, since γn|m\gamma_{n}|_{m} is a geodesic for each nn. Thus, γ\gamma is a geodesic ray based at limnγn(0)\lim_{n}\gamma_{n}(0) and is hence a CGR, so γC\gamma\in C. Therefore, CC is closed.

The "furthermore" statement follows immediately from Kőnig’s lemma, since Geo(g,η)\text{Geo}(g,\eta) 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 R={(η,g,γ)Γ^×G:γCGR(g,η)}R=\{(\eta,g,\gamma)\in\partial\hat{\Gamma}\times G^{\mathbb{N}}:\gamma\in CGR(g,\eta)\} is closed in Γ^×G\partial\hat{\Gamma}\times G^{\mathbb{N}}.

Proof.

Suppose that (ηn,gn,γn)R(\eta_{n},g_{n},\gamma_{n})\in R for all nn and that (ηn,gn,γn)(η,g,γ)(\eta_{n},g_{n},\gamma_{n})\to(\eta,g,\gamma). Then ηnηΓ^\eta_{n}\to\eta\in\partial\hat{\Gamma}, gngg_{n}\to g in GG (so that (gn)(g_{n}) is eventually equal to gg, by discreteness of GG) and γnγ\gamma_{n}\to\gamma in GG^{\mathbb{N}}, so that γCGR(g,η)\gamma\in\text{CGR}(g,\eta^{\prime}) for some ηΓ^\eta^{\prime}\in\partial\hat{\Gamma} (by Claim 4.6). We will show that η=η\eta=\eta^{\prime}.

As ηnη\eta_{n}\to\eta, by Proposition 4.5, there exists a sequence (γn)n(\gamma^{\prime}_{n})_{n} with γnCGR(g,ηn)\gamma^{\prime}_{n}\in\text{CGR}(g,\eta_{n}) which has a subsequence (γnk)k(\gamma_{n_{k}}^{\prime})_{k} that converges to some γCGR(g,η)\gamma^{\prime}\in CGR(g,\eta). Choose kk large enough such that all gnkg_{n_{k}} equal gg, so that γnk,γnkCGR(g,ηnk)\gamma_{n_{k}},\gamma^{\prime}_{n_{k}}\in CGR(g,\eta_{n_{k}}). We then have that d(γnk(m),γnk(m))2νd(\gamma_{n_{k}}(m),\gamma^{\prime}_{n_{k}}(m))\leq 2\nu for each mm. Taking kk\to\infty, we obtain that d(γ(m),γ(m))2νd(\gamma(m),\gamma^{\prime}(m))\leq 2\nu for all mm, and therefore that η=η\eta=\eta^{\prime}. Thus, (ηn,gn,γn)(η,g,γ)(\eta_{n},g_{n},\gamma_{n})\to(\eta,g,\gamma) with γCGR(g,η)\gamma\in\text{CGR}(g,\eta), so (η,g,γ)R(\eta,g,\gamma)\in R and so RR is closed.

Claim 4.8.

The set F={(η,g,(γ(0),γ(1),γ(n)))Γ^×G×G<:γCGR(g,η)}F=\{(\eta,g,(\gamma(0),\gamma(1)...,\gamma(n)))\in\partial\hat{\Gamma}\times G\times G^{<\mathbb{N}}:\gamma\in CGR(g,\eta)\} is Borel in Γ^×G×G<\partial\hat{\Gamma}\times G\times G^{<\mathbb{N}}.

Proof.

Let F={(η,g,(γ(0),γ(1),,γ(n)),γ)Γ^×G×G<×G:(η,g,γ)R and γi=γi for each 0in}F^{\prime}=\{(\eta,g,(\gamma(0),\gamma(1),...,\gamma(n)),\gamma^{\prime})\in\partial\hat{\Gamma}\times G\times G^{<\mathbb{N}}\times G^{\mathbb{N}}:(\eta,g,\gamma^{\prime})\in R\text{ and }\gamma^{\prime}_{i}=\gamma_{i}\text{ for each }0\leq i\leq n\}. By Claim 4.7, FF^{\prime} is closed in Γ^×G×G<×G\partial\hat{\Gamma}\times G\times G^{<\mathbb{N}}\times G^{\mathbb{N}}. Note that FF is the projection of FF^{\prime} to the first 3 components Γ^×G×G<\partial\hat{\Gamma}\times G\times G^{<\mathbb{N}}. Note also that the section F(η,g,(γ(0),γ(1),,γ(n)))F^{\prime}_{(\eta,g,(\gamma(0),\gamma(1),...,\gamma(n)))} is compact for every (η,g,(γ(0),γ(1),,γ(n)))Γ^×G<(\eta,g,(\gamma(0),\gamma(1),...,\gamma(n)))\in\partial\hat{\Gamma}\times G^{<\mathbb{N}}. Indeed, F(η,g,(γ(0),γ(1),,γ(n)))={γCGR(g,η):γ(i)=γ(i) for all 0in}F^{\prime}_{(\eta,g,(\gamma(0),\gamma(1),...,\gamma(n)))}=\{\gamma^{\prime}\in CGR(g,\eta):\gamma^{\prime}(i)=\gamma(i)\text{ for all }0\leq i\leq n\}, which is a closed subset of the compact set CGR(g,η)\text{CGR}(g,\eta), hence it is compact. By [13, Theorem 18.18], it follows that FF is Borel in Γ^×G×G<\partial\hat{\Gamma}\times G\times G^{<\mathbb{N}}. ∎

Claim 4.9.

The set M={(η,ξ)Γ^×Chb(Γ^):ξΞ(η)}M=\{(\eta,\xi)\in\partial\hat{\Gamma}\times C_{hb}(\hat{\Gamma}):\xi\in\Xi(\eta)\} is Borel in Γ^×Chb(Γ^)\partial\hat{\Gamma}\times C_{hb}(\hat{\Gamma}).

Proof.

We follow a similar proof to the proof of [15, Claim 4]. We will show that MM is both analytic and coanalytic, hence Borel by [13, Theorem 14.11]. By definition of Ξ(η)\Xi(\eta), we have that (η,ξ)M(\eta,\xi)\in M if and only if

γG:(η,γ(0),γ)R and ξγ=ξ\exists\gamma\in G^{\mathbb{N}}:(\eta,\gamma(0),\gamma)\in R\text{ and }\xi_{\gamma}=\xi

We also have that

ξγ=ξgG n mn fγ(m)(g)=ξ(g)\xi_{\gamma}=\xi\iff\forall g\in G\text{ }\exists n\in\mathbb{N}\text{ }\forall m\geq n\text{ }f_{\gamma(m)}(g)=\xi(g)

which gives a Borel definition of the set {(ξ,γ)Chb(Γ^)×C:ξγ=ξ}\{(\xi,\gamma)\in C_{hb}(\hat{\Gamma})\times C:\xi_{\gamma}=\xi\}. Thus, from Claim 4.7 we have that MM is analytic. To show that MM is coanalytic, we will show the following, denoting Nk(A)N_{k}(A) the kk-neighbourhood of a subset AA of GG:

(η,ξ)M if and only if λG if (η,e,λ)R, then k,γkGk+1 a geodesic path with\displaystyle(\eta,\xi)\in M\text{ if and only if }\forall\lambda\in G^{\mathbb{N}}\text{ if }(\eta,e,\lambda)\in R,\text{ then }\forall k\in\mathbb{N},\exists\gamma^{k}\in G^{k+1}\text{ a geodesic path with }
γk(0)=e such that γkN2ν(λ) and such that gG,ng such that i,j>ng,fγj(i)(g)=ξ(g)\displaystyle\gamma^{k}(0)=e\text{ such that }\gamma^{k}\subseteq N_{2\nu}(\lambda)\text{ and such that }\forall g\in G,\exists n_{g}\in\mathbb{N}\text{ such that }\forall i,j>n_{g},f_{\gamma^{j}(i)}(g)=\xi(g)

This formula defines a coanalytic set since there is a single universal quantifier \forall ranging over an uncountable standard Borel space GG^{\mathbb{N}}.

For the forward direction, if (η,ξ)M(\eta,\xi)\in M, then there exists γCGR(e,η)\gamma\in\text{CGR}(e,\eta) converging to ξ\xi. We simply take γk=γ|k\gamma^{k}=\gamma|_{k} (the restriction of γ\gamma from 0 to kk) for each kk\in\mathbb{N}. Then for each λCGR(e,η)\lambda\in\text{CGR}(e,\eta), we have d(γ(n),λ(n))2νd(\gamma(n),\lambda(n))\leq 2\nu for each nn\in\mathbb{N}, so γkN2ν(λ)\gamma^{k}\subseteq N_{2\nu}(\lambda) for each kk. Furthermore, since γ\gamma converges to ξ\xi, we have that for all gG\forall g\in G, there exists ngn_{g} such that for all i,j>ngi,j>n_{g}, we have fγj(i)(g)=ξ(g)f_{\gamma^{j}(i)}(g)=\xi(g).

For the reverse direction, let λCGR(e,η)\lambda\in\text{CGR}(e,\eta). Then there exists a sequence γkGk+1\gamma^{k}\in G^{k+1} of geodesic paths starting at ee, each contained in N2ν(λ)N_{2\nu}(\lambda) and such that fγi(j)(g)ξ(g)f_{\gamma^{i}(j)}(g)\to\xi(g). For each ii, fix k=i+3ν+1k=i+3\nu+1 and using ν\nu-hyperbolicity, choose an NN sufficiently large such that for all nNn\geq N, we have

d(γn(t),λ(t))2νd(\gamma_{n}(t),\lambda(t))\leq 2\nu

for all tkt\leq k. Arguing as in the proof of Theorem 3.2, we have that {γj(i):jN}\{\gamma^{j}(i):j\geq N\} is finite, so that {γj(i):j}\{\gamma^{j}(i):j\in\mathbb{N}\} is finite for each ii. Therefore, by Kőnig’s lemma, (γk)k(\gamma^{k})_{k} has a subsequence converging to some CGR γ\gamma based at ee, and γN2ν(λ)\gamma\subseteq N_{2\nu}(\lambda), so γCGR(e,η)\gamma\in\text{CGR}(e,\eta). From fγi(j)(g)ξ(g)f_{\gamma^{i}(j)}(g)\to\xi(g) as i,ji,j\to\infty, we have that ξγ=ξ\xi_{\gamma}=\xi. Since γCGR(e,η)\gamma\in\text{CGR}(e,\eta), we conclude that (η,ξ)M(\eta,\xi)\in M.

By [15, Proposition 5.2], for each ηΓ^\eta\in\partial\hat{\Gamma}, we have that the section Mη=Ξ(η)M_{\eta}=\Xi(\eta) is finite, having cardinality bounded above by the constant BB from Theorem 3.2. Since MM is Borel and has finite sections of size at most BB, by the Lusin-Novikov theorem we have Borel functions ξ1,,ξB:Γ^Chb(Γ^)\xi_{1},...,\xi_{B}:\partial\hat{\Gamma}\to C_{hb}(\hat{\Gamma}) such that MM is the union of the graphs Gξi={(η,ξi(η)):ηΓ^}G_{\xi_{i}}=\{(\eta,\xi_{i}(\eta)):\eta\in\partial\hat{\Gamma}\} of the ξi\xi_{i}.

Claim 4.10.

For each i=1,,Bi=1,...,B, Qi={(η,g,h)Γ^×G2:hQ(g,ξi(η))}Q_{i}=\{(\eta,g,h)\in\partial\hat{\Gamma}\times G^{2}:h\in Q(g,\xi_{i}(\eta))\} is Borel in Γ^×G2\partial\hat{\Gamma}\times G^{2}.

Proof.

By [15, Lemma 4.2], for x,yGx,y\in G, denoting γ(x,y)\gamma(x,y) the union of all geodesic paths in Γ^\hat{\Gamma} from xx to yy, we have that Q(g,ξi(η))=nγ(g,xn)Q(g,\xi_{i}(\eta))=\bigcup_{n\in\mathbb{N}}\gamma(g,x_{n}) for some, equivalently any, (xn)nCGR(g,η)(x_{n})_{n}\in\text{CGR}(g,\eta) converging to ξi(η)\xi_{i}(\eta). From this, we obtain that:

hQ(g,ξi(η))λC(resp. λC) : λ(0)=g and ξλ=ξi(η) and n : hγ(g,λ(n))h\in Q(g,\xi_{i}(\eta))\iff\exists\lambda\in C\text{(resp. $\forall\lambda\in C$) : $\lambda(0)=g$ and $\xi_{\lambda}=\xi_{i}(\eta)$ and $\exists n\in\mathbb{N}$ : $h\in\gamma(g,\lambda(n))$}

This yields the analyticity (from the \exists above) and coanalyticity (from the \forall above) of QiQ_{i}, hence Borelness of QiQ_{i}.

Claim 4.11.

The set P={(η,h)Γ^×G:hΓ^s,η}P=\{(\eta,h)\in\partial\hat{\Gamma}\times G:h\in\hat{\Gamma}_{s,\eta}\} is Borel in Γ^×G\partial\hat{\Gamma}\times G.

Proof.

We have that hΓ^s,ηh\in\hat{\Gamma}_{s,\eta} if and only if:

n,γnGn+1:(η,h,γn)F and iB k<n,(η,h,γn(k))Qi\forall n\in\mathbb{N},\exists\gamma^{n}\in G^{n+1}:(\eta,h,\gamma^{n})\in F\text{ and }\forall i\leq B\text{ }\forall k<n,(\eta,h,\gamma^{n}(k))\in Q_{i}

Indeed, if hΓ^s,ηh\in\hat{\Gamma}_{s,\eta}, then ξΞ(η)Q(h,ξ)\bigcap_{\xi\in\Xi(\eta)}Q(h,\xi) contains a CGR γCGR(h,η)\gamma\in\text{CGR}(h,\eta), so we can take γn=γ|n\gamma^{n}=\gamma|_{n} (the restriction from 0 to nn) for all nn\in\mathbb{N} to satisfy the above condition.

Conversely, if the above condition holds, then by local finiteness of Geo(h,η)\text{Geo}(h,\eta), the sequence (γn)n(\gamma^{n})_{n\in\mathbb{N}} with (η,h,γn)F(\eta,h,\gamma^{n})\in F will have a subsequence converging to some γCGR(h,η)\gamma\in\text{CGR}(h,\eta) and the above condition yields that γξΞ(η)Q(h,ξ)\gamma\subseteq\bigcap_{\xi\in\Xi(\eta)}Q(h,\xi), so that hΓ^s,ηh\in\hat{\Gamma}_{s,\eta}.

Since FF and QiQ_{i} are Borel, we conclude that PP is Borel.

Claim 4.12.

The set P1={(ξ,η,h)Chb(Γ^)×Γ^×G:hΓ^s,η and ξ=ξh,η}P_{1}=\{(\xi,\eta,h)\in C_{hb}(\hat{\Gamma})\times\partial\hat{\Gamma}\times G:h\in\hat{\Gamma}_{s,\eta}\text{ and }\xi=\xi_{h,\eta}\} is Borel in Chb(Γ^)×Γ^×GC_{hb}(\hat{\Gamma})\times\partial\hat{\Gamma}\times G.

Proof.

We have (ξ,η,h)P1(\xi,\eta,h)\in P_{1} if and only if:

(η,h)P and iB : (η,ξ)Gξi and jBQ(h,ξi(η))Q(h,ξj(η))(\eta,h)\in P\text{ and $\exists i\leq B$ : $(\eta,\xi)\in G_{\xi_{i}}$ and $\forall j\leq B$, $Q(h,\xi_{i}(\eta))\subseteq Q(h,\xi_{j}(\eta))$}

Since PP is Borel (Claim 4.11), GξiG_{\xi_{i}} is Borel (as ξi\xi_{i} is Borel), and QiQ_{i} is Borel (Claim 4.10), the above yields that P1P_{1} is Borel.

Claim 4.13.

The set L={(h,ξ,η)G×Chb(Γ^)×Γ^:hY(e,ξ),ξΞ(η)}L=\{(h,\xi,\eta)\in G\times C_{hb}(\hat{\Gamma})\times\partial\hat{\Gamma}:h\in Y(e,\xi),\xi\in\Xi(\eta)\} is Borel in G×Chb(Γ^)×Γ^G\times C_{hb}(\hat{\Gamma})\times\partial\hat{\Gamma}.

Proof.

We have that (h,ξ,η)L(h,\xi,\eta)\in L if and only if (η,ξ)M(\eta,\xi)\in M and hh is the closest element to ee (in the metric dd) such that hGeo(e,η)h\in\text{Geo}(e,\eta) and (ξ,η,h)P1(\xi,\eta,h)\in P_{1}. Thus, by Claims 4.9, 4.10, 4.12, LL is Borel (note that hGeo(e,η)(η,e,h)Qih\in\text{Geo}(e,\eta)\iff(\eta,e,h)\in Q_{i} for some iBi\leq B, so {(h,η)G×Γ^:hGeo(e,η)}\{(h,\eta)\in G\times\partial\hat{\Gamma}:h\in\text{Geo}(e,\eta)\} is Borel in G×Γ^G\times\partial\hat{\Gamma} by Claim 4.10).

Claim 4.14.

The set B={(g,h,ξ,η)G2×Chb(Γ^)×Γ^:gQ(h,ξ),hY(e,ξ),ξΞ(η)}B=\{(g,h,\xi,\eta)\in G^{2}\times C_{hb}(\hat{\Gamma})\times\partial\hat{\Gamma}:g\in Q(h,\xi),h\in Y(e,\xi),\xi\in\Xi(\eta)\} is Borel in G2×Chb(Γ^)×Γ^G^{2}\times C_{hb}(\hat{\Gamma})\times\partial\hat{\Gamma}.

Proof.

We have that (g,h,ξ,η)B(g,h,\xi,\eta)\in B if and only if ir\exists i\leq r such that ξ=ξi(η)\xi=\xi_{i}(\eta) and (η,h,g)Qi(\eta,h,g)\in Q_{i} and (h,ξ,η)L(h,\xi,\eta)\in L. Since L,QiL,Q_{i} and ξi\xi_{i} are Borel, it follows that BB is Borel.

Claim 4.15.

The set A={(η,g)Γ^×G:gGeo1(e,η)}A=\{(\eta,g)\in\partial\hat{\Gamma}\times G:g\in\text{Geo}_{1}(e,\eta)\} is Borel in Γ^×G\partial\hat{\Gamma}\times G.

Proof.

Since Geo1(e,η)=ξΞ(η)hY(e,ξ)Q(h,ξ)\text{Geo}_{1}(e,\eta)=\bigcup_{\xi\in\Xi(\eta)}\bigcup_{h\in Y(e,\xi)}Q(h,\xi), we have:

(η,g)AξΞ(η),hY(e,ξ):gQ(h,ξ)ξΞ(η),hY(e,ξ):(g,h,ξ,η)B(\eta,g)\in A\iff\exists\xi\in\Xi(\eta),\exists h\in Y(e,\xi):g\in Q(h,\xi)\iff\exists\xi\in\Xi(\eta),\exists h\in Y(e,\xi):(g,h,\xi,\eta)\in B

Therefore, AA is the projection (g,h,ξ,η)(η,g)(g,h,\xi,\eta)\mapsto(\eta,g) of BB onto Γ^×G\partial\hat{\Gamma}\times G. By Claim 4.14, BB is Borel. Also, the sections {(h,ξ)G×Chb(Γ^):hY(e,ξ),ξΞ(η)}\{(h,\xi)\in G\times C_{hb}(\hat{\Gamma}):h\in Y(e,\xi),\xi\in\Xi(\eta)\} of BB are finite by Theorem 3.2 and [15, Proposition 5.2]. Therefore, by the Lusin-Novikov theorem, AA is Borel.

Claim 4.16.

The set D={(η,(γ(0),γ(1),,γ(n)))Γ^×G<:γ(0)Geo1(e,η) and γCGR(γ(0),η)}D=\{(\eta,(\gamma(0),\gamma(1),...,\gamma(n)))\in\partial\hat{\Gamma}\times G^{<\mathbb{N}}:\gamma(0)\in Geo_{1}(e,\eta)\text{ and }\gamma\in CGR(\gamma(0),\eta)\} is Borel in Γ^×G<\partial\hat{\Gamma}\times G^{<\mathbb{N}}.

Proof.

We have that (η,(γ(0),γ(1),,γ(n)))D(\eta,(\gamma(0),\gamma(1),...,\gamma(n)))\in D if and only if (η,γ(0),(γ(0),γ(1),,γ(n)))F(\eta,\gamma(0),(\gamma(0),\gamma(1),...,\gamma(n)))\in F and (η,γ(0))A(\eta,\gamma(0))\in A. By Claim 4.15, AA is Borel in Γ^×G\partial\hat{\Gamma}\times G. Also, FF is Borel by Claim 4.8. Therefore, DD is Borel.

Claim 4.17.

For each nn, the set Sn:={(η,sn)Γ^×(2n)n:sn=snη}S_{n}:=\{(\eta,s^{n})\in\partial\hat{\Gamma}\times(2^{n})^{n}:s^{n}=s_{n}^{\eta}\} is Borel in Γ^×(2n)n\partial\hat{\Gamma}\times(2^{n})^{n}.

Proof.

We have that (η,sn)Sn(\eta,s^{n})\in S_{n} if and only if sns^{n} is the <n<_{n}-minimal element in (2n)n(2^{n})^{n} for which the following holds:

m,(γ(0),γ(1),,γ(n))Gn+1:d(γ(0),e)m,(η,(γ(0),γ(1),,γ(n)))D and lab(γ)|n=sn\forall m\in\mathbb{N},\exists(\gamma(0),\gamma(1),...,\gamma(n))\in G^{n+1}:d(\gamma(0),e)\geq m,(\eta,(\gamma(0),\gamma(1),...,\gamma(n)))\in D\text{ and $\text{lab}(\gamma)|_{n}=s^{n}$}

Note that the the "only if" holds by local finiteness of Geo(e,η)\text{Geo}(e,\eta). Thus, SnS_{n} is Borel by Claim 4.16.

Now let EE denote the orbit equivalence relation of the action of GG on Γ^\partial\hat{\Gamma}.

Definition 4.18.

Let Z={ηΓ^:knη}Z=\{\eta\in\partial\hat{\Gamma}:k_{n}^{\eta}\nrightarrow\infty\}.

Since Geo1(e,η)\text{Geo}_{1}(e,\eta) is locally finite (as Geo(e,η)\text{Geo}(e,\eta) is locally finite and Geo1(e,η)Geo(e,η)\text{Geo}_{1}(e,\eta)\subseteq\text{Geo}(e,\eta)), we have that ZZ is the set of all η\eta such that there exists gηg_{\eta} belonging to TnηT_{n}^{\eta} for all nn, i.e. for which there exists γηCGR(gη,η)\gamma^{\eta}\in CGR(g_{\eta},\eta) with label sη(2<)s^{\eta}\in(2^{<\mathbb{N}})^{\mathbb{N}}.

Lemma 4.19.

The map α:(Z,E|Z)(Γ^,=)\alpha:(Z,E|_{Z})\to(\partial\hat{\Gamma},=) given by ηgη1η\eta\mapsto g_{\eta}^{-1}\eta is a Borel reduction.

Proof.

We argue as in [15]. First, let us show that snη=sngηs_{n}^{\eta}=s_{n}^{g\eta} for each gGg\in G, each ηΓ^\eta\in\partial\hat{\Gamma} and each nn\in\mathbb{N}. If there are infinitely many pairs (h,snη)Cη(h,s_{n}^{\eta})\in C^{\eta}, then since the left action of GG on Γ^\hat{\Gamma} preserves labels of geodesics, there are infinitely many pairs (gh,snη)(gh,s_{n}^{\eta}), where snη=lab(γ)|ns_{n}^{\eta}=\text{lab}(\gamma)|_{n} for some γCGR(γ(0),gη)\gamma\in\text{CGR}(\gamma(0),g\eta) and where γ(0)gGeo1(e,η)=Geo1(g,gη)\gamma(0)\in g\text{Geo}_{1}(e,\eta)=\text{Geo}_{1}(g,g\eta) (using [15, Lemma 5.10] in the last line).

By [15, Theorem 5.9], the symmetric difference between Geo1(g,gη)\text{Geo}_{1}(g,g\eta) and Geo1(e,gη)\text{Geo}_{1}(e,g\eta) is finite and so there are infinitely many pairs (gh,snη)Geo1(e,gη)(gh,s_{n}^{\eta})\in\text{Geo}_{1}(e,g\eta). Hence, there are infinitely many pairs (gh,snη)Cgη(gh,s_{n}^{\eta})\in C^{g\eta}. Thus, as snηs_{n}^{\eta} is least in the order <n<_{n} that appears infinitely often in CηC^{\eta}, we have that snη=sngηs_{n}^{\eta}=s_{n}^{g\eta}. As snη=sngηs_{n}^{\eta}=s_{n}^{g\eta} for each nn, we have sη=sgηs^{\eta}=s^{g\eta}.

This implies that α\alpha is constant on GG-orbits. Indeed, suppose θ=gη\theta=g\eta for some gGg\in G, η,θZ\eta,\theta\in Z. We have that α\alpha maps the boundary point [γθ][\gamma^{\theta}] to the boundary point [gθ1γθ][g_{\theta}^{-1}\gamma^{\theta}]. Note that gθ1γθCGR(e,gθ1θ)g_{\theta}^{-1}\gamma^{\theta}\in\text{CGR}(e,g_{\theta}^{-1}\theta) and lab(gθ1γθ)=sθ\text{lab}(g_{\theta}^{-1}\gamma^{\theta})=s^{\theta}, because γθ\gamma^{\theta} has label sθs^{\theta} and left multiplication preserves labels of geodesics. On the other hand, α\alpha maps η=[γη]\eta=[\gamma^{\eta}] to gη1η=[gη1γη]g_{\eta}^{-1}\eta=[g_{\eta}^{-1}\gamma^{\eta}]. We have that gη1γηCGR(e,gη1η)g_{\eta}^{-1}\gamma^{\eta}\in\text{CGR}(e,g_{\eta}^{-1}\eta) and lab(gη1γη)=sη\text{lab}(g_{\eta}^{-1}\gamma^{\eta})=s^{\eta}. But by above, sη=sgη=sθs^{\eta}=s^{g\eta}=s^{\theta}. Therefore, gη1γηg_{\eta}^{-1}\gamma^{\eta} and gθ1γθg_{\theta}^{-1}\gamma^{\theta} both start at ee and have the same label. Therefore, they are the same geodesic. Hence, gθ1θ=gη1ηg_{\theta}^{-1}\theta=g_{\eta}^{-1}\eta i.e. α(θ)=α(η)\alpha(\theta)=\alpha(\eta).

It follows that α\alpha is reduction to == on Γ^\partial\hat{\Gamma}. Indeed, the above shows that θEηα(θ)=α(η)\theta E\eta\implies\alpha(\theta)=\alpha(\eta). Conversely, if α(θ)=α(η)\alpha(\theta)=\alpha(\eta), then gη1η=gθ1θg_{\eta}^{-1}\eta=g_{\theta}^{-1}\theta, so θ=gθgη1η\theta=g_{\theta}g_{\eta}^{-1}\eta, and therefore θEη\theta E\eta.

It remains to show that α\alpha is Borel. To show this, let us first show that the set U:={(η,s)Z×(2):s=sη}U:=\{(\eta,s)\in Z\times(2^{\mathbb{N}})^{\mathbb{N}}:s=s^{\eta}\} is Borel. We have s=sηs=s^{\eta} if and only if (η,s|n)Sn(\eta,s|_{n})\in S_{n} for each nn\in\mathbb{N}, so {(η,s)Z×(2):s=sη}\{(\eta,s)\in Z\times(2^{\mathbb{N}})^{\mathbb{N}}:s=s^{\eta}\} is Borel by Claim 4.17 (note that the map (η,s)(η,s|n)(\eta,s)\mapsto(\eta,s|_{n}) is continuous, hence Borel, for each nn\in\mathbb{N}).

Now the Borelness of UU implies the Borelness of the graph of α\alpha. Indeed, note that for ηZ\eta\in Z and θΓ^\theta\in\partial\hat{\Gamma}, denoting lab:Z×CZ×(2)\text{lab}:Z\times C\to Z\times(2^{\mathbb{N}})^{\mathbb{N}} the continuous map (η,γ)(η,lab(γ))(\eta,\gamma)\mapsto(\eta,\text{lab}(\gamma)), we have:

θ=gη1η\displaystyle\theta=g_{\eta}^{-1}\eta γC:γCGR(e,θ) and lab(γ)=sη\displaystyle\iff\exists\gamma\in C:\gamma\in\text{CGR}(e,\theta)\text{ and }\text{lab}(\gamma)=s^{\eta}
γC:(θ,e,Γ)R and (η,γ)lab1(U)\displaystyle\iff\exists\gamma\in C:(\theta,e,\Gamma)\in R\text{ and }(\eta,\gamma)\in\text{lab}^{-1}(U)

Putting T={(η,θ,γ)Z×Γ^×C:(θ,e,γ)R and (η,γ)lab1(U)}T=\{(\eta,\theta,\gamma)\in Z\times\partial\hat{\Gamma}\times C:(\theta,e,\gamma)\in R\text{ and }(\eta,\gamma)\in\text{lab}^{-1}(U)\}, we have that TT is Borel because RR and UU are Borel (see Claim 4.7 for the Borelness of RR). By above, the graph of α\alpha is the projection projZ×Γ^(T)\text{proj}_{Z\times\partial\hat{\Gamma}}(T) of TT onto the first two coordinates (η,θ)(\eta,\theta). For each (η,θ)Z×Γ^(\eta,\theta)\in Z\times\partial\hat{\Gamma}, the section T(η,θ)={γC:(η,θ,γ)T}={γC:γCGR(e,θ) and lab(γ)=sη}T_{(\eta,\theta)}=\{\gamma\in C:(\eta,\theta,\gamma)\in T\}=\{\gamma\in C:\gamma\in\text{CGR}(e,\theta)\text{ and }\text{lab}(\gamma)=s^{\eta}\} 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 projZ×Γ^(T)\text{proj}_{Z\times\partial\hat{\Gamma}}(T) is Borel. Thus, the graph of α\alpha is Borel, so α\alpha is Borel.

Lemma 4.20.

EE is smooth on the saturation [Z]E={ηΓ^:θZ such that θEη}[Z]_{E}=\{\eta\in\partial\hat{\Gamma}:\exists\theta\in Z\text{ such that }\theta E\eta\}.

Proof.

By Lemma 4.19, EE is smooth on ZZ, yielding the claim. ∎

Definition 4.21.

Let Y=Γ^[Z]EY=\partial\hat{\Gamma}\setminus[Z]_{E}. For each nn\in\mathbb{N}, define Hn:Γ^2GH_{n}:\partial\hat{\Gamma}\to 2^{G} by Hn(η)=(gnη)1TnηH_{n}(\eta)=(g_{n}^{\eta})^{-1}T_{n}^{\eta}. Let FnF_{n} be the equivalence relation on imHn\mathrm{im}H_{n} which is the restriction of the shift action of GG on 2G2^{G} to imHn\mathrm{im}H_{n}.

The following lemma is a generalization of [15, Lemma 6.7].

Lemma 4.22.

There exists a constant KK such that for each nn\in\mathbb{N}, each equivalence class of FnF_{n} has size at most KK.

Proof.

Note that by Thereom 3.2, we have that each closed ball of radius rr in Geo(x,η)\text{Geo}(x,\eta) has cardinality at most (2(r+2ν)+1)B(2(r+2\nu)+1)B, where BB is the constant from Theorem 3.2. We will show that we can take K=(20ν+1)BK=(20\nu+1)B.

Let η,θΓ^\eta,\theta\in\partial\hat{\Gamma} and suppose that Hn(η)=gHn(θ)H_{n}(\eta)=gH_{n}(\theta). 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 Γ^\hat{\Gamma}), we have d(e,g)8νd(e,g)\leq 8\nu. For completeness, let us reproduce this proof.

By defiinition, TnηT_{n}^{\eta} (resp. TnθT_{n}^{\theta}) is an infinite subset of Geo(e,η)\text{Geo}(e,\eta) (resp. Geo(e,θ)\text{Geo}(e,\theta)). Since Geo(e,η)\text{Geo}(e,\eta) is locally finite, this means that TnηT_{n}^{\eta} (resp. TnθT_{n}^{\theta}) uniquely determines η\eta (resp. θ\theta). From Hn(η)=gHn(θ)H_{n}(\eta)=gH_{n}(\theta), we have (gnη)1Tnη=g(gnθ)1Tnθ(g_{n}^{\eta})^{-1}T_{n}^{\eta}=g(g_{n}^{\theta})^{-1}T_{n}^{\theta} and since TnηT_{n}^{\eta} and TnθT_{n}^{\theta} determine their boundary points, this implies that (gnη)1η=g(gnθ)1θ(g_{n}^{\eta})^{-1}\eta=g(g_{n}^{\theta})^{-1}\theta. Let us denote by σ\sigma the common boundary point (gnη)1η=g(gnθ)1θ(g_{n}^{\eta})^{-1}\eta=g(g_{n}^{\theta})^{-1}\theta.

Refer to caption
Figure 3: The geometry of the proof of Lemma 4.22.

We have that g,e(gnη)1Tnη=g(gnθ)1TnθGeo(g(gnθ)1,σ)g,e\in(g_{n}^{\eta})^{-1}T_{n}^{\eta}=g(g_{n}^{\theta})^{-1}T_{n}^{\theta}\subseteq\text{Geo}(g(g_{n}^{\theta})^{-1},\sigma), so there exists λCGR(g(gnθ)1,σ)\lambda\in\text{CGR}(g(g_{n}^{\theta})^{-1},\sigma) passing through gg and λCGR(g(gnθ)1,σ)\lambda^{\prime}\in\text{CGR}(g(g_{n}^{\theta})^{-1},\sigma) passing through ee. Write g=λ(m1)g=\lambda(m_{1}) and e=λ(m2)e=\lambda^{\prime}(m_{2}) for some m1,m2m_{1},m_{2}\in\mathbb{N}. Note that by ν\nu-hyperbolicity, we have d(e,λ(m2))2νd(e,\lambda(m_{2}))\leq 2\nu. Also, we have m2m1m_{2}\geq m_{1}. Indeed, since gnθg1gnθg1(gnη)1Tnη=Tnθg_{n}^{\theta}g^{-1}\in g_{n}^{\theta}g^{-1}(g_{n}^{\eta})^{-1}T_{n}^{\eta}=T_{n}^{\theta}, we have:

m2=d(e,g(gnθ)1)=d(e,gnθg1)d(e,gnθ)=d(e,(gnθ)1)=d(g,g(gnθ)1)=m1m_{2}=d(e,g(g_{n}^{\theta})^{-1})=d(e,g_{n}^{\theta}g^{-1})\geq d(e,g_{n}^{\theta})=d(e,(g_{n}^{\theta})^{-1})=d(g,g(g_{n}^{\theta})^{-1})=m_{1}

where d(e,gnθg1)d(e,gnθ)d(e,g_{n}^{\theta}g^{-1})\geq d(e,g_{n}^{\theta}) holds by \leq-minimality of gnθg_{n}^{\theta} in TnθT_{n}^{\theta}.

Similarly, from g,e(gnη)1Tnηg,e\in(g_{n}^{\eta})^{-1}T_{n}^{\eta}, we have g,eGeo((gnη)1,σ)g,e\in\text{Geo}((g_{n}^{\eta})^{-1},\sigma), and so there exists γCGR((gnθ)1,σ)\gamma\in\text{CGR}((g_{n}^{\theta})^{-1},\sigma) passing through gg and γCGR((gnθ)1,σ)\gamma^{\prime}\in\text{CGR}((g_{n}^{\theta})^{-1},\sigma) passing through ee. Write g=γ(m3)g=\gamma(m_{3}) and e=γ(m4)e=\gamma^{\prime}(m_{4}) for some m3,m4m_{3},m_{4}\in\mathbb{N} (see Figure 3). By ν\nu-hyperbolicity, we have d(e,γ(m4))2νd(e,\gamma(m_{4}))\leq 2\nu and m4m3m_{4}\leq m_{3} because gnηggnηg(gnθ)1Tnθ=Tnηg_{n}^{\eta}g\in g_{n}^{\eta}g(g_{n}^{\theta})^{-1}T_{n}^{\theta}=T_{n}^{\eta} and so by the \leq-minimality of gnηg_{n}^{\eta} in TnηT_{n}^{\eta}, we have that:

m3=d((gnη)1,g)=d(e,gnηg)d(e,gnη)=d(e,(gnη)1)=m4m_{3}=d((g_{n}^{\eta})^{-1},g)=d(e,g_{n}^{\eta}g)\geq d(e,g_{n}^{\eta})=d(e,(g_{n}^{\eta})^{-1})=m_{4}

Let us now consider the sub-CGRs of λ\lambda and γ\gamma starting at gg. Using ν\nu-hyperbolicity, since m2m1m_{2}\geq m_{1}, there exists m5m3m_{5}\geq m_{3} such that d(λ(m2),γ(m5))2νd(\lambda(m_{2}),\gamma(m_{5}))\leq 2\nu. Then by the triangle inequality and our above estimates, we have:

d(γ(m4),γ(m5))d(γ(m4),e)+d(e,λ(m2))+d(λ(m2),γ(m5))6νd(\gamma(m_{4}),\gamma(m_{5}))\leq d(\gamma(m_{4}),e)+d(e,\lambda(m_{2}))+d(\lambda(m_{2}),\gamma(m_{5}))\leq 6\nu

Therefore,

d(e,g)=d(e,γ(m3))d(e,γ(m4))+d(γ(m4),γ(m3))2ν+d(γ(m4),γ(m5))8νd(e,g)=d(e,\gamma(m_{3}))\leq d(e,\gamma(m_{4}))+d(\gamma(m_{4}),\gamma(m_{3}))\leq 2\nu+d(\gamma(m_{4}),\gamma(m_{5}))\leq 8\nu

where we have d(γ(m4),γ(m3))d(γ(m4),γ(m5))d(\gamma(m_{4}),\gamma(m_{3}))\leq d(\gamma(m_{4}),\gamma(m_{5})) because m5m3m4m_{5}\geq m_{3}\geq m_{4}.

Thus, Hn(η)=gHn(θ)H_{n}(\eta)=gH_{n}(\theta) implies that gg is in the ball of radius 8ν8\nu about ee in Geo((gηn)1,σ)\text{Geo}((g_{\eta}^{n})^{-1},\sigma), which has cardinality at most (2(8ν+2ν)+1)B=(20ν+1)B=K(2(8\nu+2\nu)+1)B=(20\nu+1)B=K. Thus, FnF_{n}-classes have cardinality at most KK.

The following remaining results have the same proof as in [15].

Lemma 4.23.

Let nn\in\mathbb{N}. Then the map HnH_{n} is Borel and so imHn\mathrm{im}H_{n} is analytic.

Proof.

The sets {(η,gnη)Γ^×G},{(η,Tnη)Γ^×2G}\{(\eta,g_{n}^{\eta})\in\partial\hat{\Gamma}\times G\},\{(\eta,T_{n}^{\eta})\in\partial\hat{\Gamma}\times 2^{G}\} and GHn={(η,Hn(η)):Γ^×2G}G_{H_{n}}=\{(\eta,H_{n}(\eta)):\partial\hat{\Gamma}\times 2^{G}\} are all definable using formulas with countable quantifiers and references to the Borel sets DD and SnS_{n} (see Claims 4.16 and 4.17), so these sets are all Borel. As GHnG_{H_{n}} is the graph of HnH_{n}, it is Borel, so HnH_{n} is Borel and hence imHn\text{im}H_{n} is analytic. ∎

Using [15, Lemma 2.3], there exists a finite Borel equivalence relation FnF_{n}^{\prime} on 2G2^{G} with FnFnF_{n}\subseteq F_{n}^{\prime}. Since FnF_{n}^{\prime} is finite, Borel, there exists a Borel reduction fn:2G2f_{n}:2^{G}\to 2^{\mathbb{N}} from FnF_{n}^{\prime} to E0E_{0} for each nn\in\mathbb{N}, using which we define f:Γ^(2)f:\partial\hat{\Gamma}\to(2^{\mathbb{N}})^{\mathbb{N}} by f(η)=(fn(Hn(η)))nf(\eta)=(f_{n}(H_{n}(\eta)))_{n}. Put E=f1(E1)E^{\prime}=f^{-1}(E_{1}), i.e. θEηf(θ)E1f(η)\theta E^{\prime}\eta\iff f(\theta)E_{1}f(\eta).

Lemma 4.24.

The equivalence relation EE^{\prime} is a hyperfinite countable Borel equivalence relation.

Proof.

Since HnH_{n} is Borel, we have that EE^{\prime} is Borel. We also have that EE^{\prime} 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 EE^{\prime} is countable.

For each nn\in\mathbb{N}, define the relation EnE_{n}^{\prime} on Γ^\partial\hat{\Gamma} by ηEnθ\eta E_{n}^{\prime}\theta if fm(Hm(η))=fm(Hm(θ))f_{m}(H_{m}(\eta))=f_{m}(H_{m}(\theta)) for all mnm\geq n. Each EnE_{n}^{\prime} is countable because if ηEnθ\eta E_{n}^{\prime}\theta, then fn(Hn(η))=fn(Hn(θ))f_{n}(H_{n}(\eta))=f_{n}(H_{n}(\theta)), and fnHnf_{n}\circ H_{n} is countable-to-one since HnH_{n} is countable-to-one because if Hn(η)=Hn(θ)H_{n}(\eta)=H_{n}(\theta), then ηEθ\eta E\theta and EE is countable, and fnf_{n} is finite-to-one since FnF_{n}^{\prime} is finite. Therefore, there are only countably many choices for η\eta such that ηEnθ\eta E_{n}^{\prime}\theta once θ\theta is fixed. Thus, EnE_{n}^{\prime} is countable. Noting that E=nEnE^{\prime}=\bigcup_{n\in\mathbb{N}}E_{n}^{\prime}, we obtain that EE^{\prime} is countable.

Lemma 4.25.

ff is a homomorphism from E|YE|_{Y} to E1E_{1}.

Proof.

Suppose η,θY\eta,\theta\in Y are EE-related, as witnessed by gGg\in G (so gη=θg\eta=\theta). By [15, Theorem 5.9] and [15, Lemma 3.10], we have that gGeo1(e,η)g\text{Geo}_{1}(e,\eta) and Geo1(e,θ)\text{Geo}_{1}(e,\theta) differ by a finite set. By local finiteness of Geo(e,η)\text{Geo}(e,\eta) and since η,θY\eta,\theta\in Y, we have that there exists NN\in\mathbb{N} such that gTnηGeo1(e,θ)gT_{n}^{\eta}\subseteq\text{Geo}_{1}(e,\theta) for all nNn\geq N. By the proof of Lemma 4.19, we have snη=snθs_{n}^{\eta}=s_{n}^{\theta}, which gives, together with gTnηGeo1(e,θ)gT_{n}^{\eta}\subseteq\text{Geo}_{1}(e,\theta), that gTnη=TnθgT_{n}^{\eta}=T_{n}^{\theta}. This then yields (gnθ)1ggnηHnη=Hnθ(g_{n}^{\theta})^{-1}gg_{n}^{\eta}H_{n}^{\eta}=H_{n}^{\theta} for all nNn\geq N. Thus, we have Hn(η)FnHn(θ)H_{n}(\eta)F_{n}H_{n}(\theta) and so Hn(η)FnHn(θ)H_{n}(\eta)F_{n}^{\prime}H_{n}(\theta) for all nNn\geq N since FnFnF_{n}\subseteq F_{n}^{\prime}. Therefore, fn(Hn(η))=fn(Hn(θ))f_{n}(H_{n}(\eta))=f_{n}(H_{n}(\theta)) for all nNn\geq N and so f(η)E1f(θ)f(\eta)E_{1}f(\theta).

Let us now establish Theorem A on the hyperfiniteness of EE, following the proof of [15, Theorem A].

Proof of Theorem A.

Note that E|YE|_{Y} is a sub-relation of EE^{\prime}. Indeed, if θ,ηY\theta,\eta\in Y and θEη\theta E\eta, then by Lemma 4.25, we have that f(θ)E1f(η)f(\theta)E_{1}f(\eta), which implies θEη\theta E^{\prime}\eta. By Lemma 4.24, we have that EE^{\prime} is hyperfinite, so E|YE|_{Y} is hyperfinite, since a sub-relation of a hyperfinite equivalence relation is hyperfinite. On Γ^Y=[Z]E\partial\hat{\Gamma}\setminus Y=[Z]_{E}, EE is smooth by Lemma 4.20, and hence hyperfinite. Therefore, EE is hyperfinite on Γ^\partial\hat{\Gamma}.

Recall that we worked with a fixed a finite generating set XX in this section. If we use a different finite generating set XX^{\prime} for GG, then the relative Cayley graph Γ^\hat{\Gamma}^{\prime} corresponding to XX^{\prime} is GG-equivariantly quasi-isometric to Γ^\hat{\Gamma} (via the identity map on GG), so Γ^\partial\hat{\Gamma}^{\prime} is GG-equivariantly homeomorphic to Γ^\partial\hat{\Gamma}. It follows that the orbit equivalence relation of GG on Γ^\partial\hat{\Gamma}^{\prime} is also hyperfinite. \square

As a corollary, we obtain Corollary B on the hyperfiniteness of the action of GG on (G,𝒫)\partial(G,\mathcal{P}), where 𝒫\mathcal{P} is the collection of parabolic subgroups.

Proof of Corollary B.

By Theorem 2.1, Γ^\partial\hat{\Gamma} embeds GG-equivariantly and topologically into (G,𝒫)\partial(G,\mathcal{P}) with countable complement. Therefore, the orbit equivalence relation of GG on Γ^\partial\hat{\Gamma} is a subrelation of the orbit equivalence relation of GG on (G,𝒫)\partial(G,\mathcal{P}). Since the orbit equivalence relation of GG on Γ^\partial\hat{\Gamma} is hyperfinite (by Theorem A) and since (G,𝒫)Γ^\partial(G,\mathcal{P})\setminus\partial\hat{\Gamma} is countable, it follows that the orbit equivalence relation of GG on (G,𝒫)\partial(G,\mathcal{P}) is also hyperfinite. \square

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 PSL(2,)PSL(2,\mathbb{Z}) in 1\mathbb{R}^{1} 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.