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

Bi-exactness of relatively hyperbolic groups

Koichi Oyakawa
Abstract

We prove that finitely generated relatively hyperbolic groups are bi-exact if and only if all peripheral subgroups are bi-exact. This is a generalization of Ozawa’s result which claims that finitely generated relatively hyperbolic groups are bi-exact if all peripheral subgroups are amenable.

1 Introduction

MSC Primary: 20F65. Secondary: 20F67, 20F99, 46L10.Key words and phrases: bi-exact groups, relatively hyperbolic groups, proper arrays.

Bi-exactness is an analytic property of groups defined by Ozawa in [16] (by the name of class 𝒮\mathcal{S}). Recall that a group GG is exact if there exists a compact Hausdorff space on which GG acts topologically amenably and bi-exact if it is exact and there exists a map μ:GProb(G)\mu\colon G\to{\rm Prob}(G) such that for every s,tGs,t\in G, we have

limxμ(sxt)s.μ(x)1=0.\lim_{x\to\infty}\|\mu(sxt)-s.\mu(x)\|_{1}=0.

It is known that exactness is equivalent to Yu’s property A (cf. [13]).

The notion of bi-exactness is of fundamental importance to the study of operator algebras. Notably, Ozawa proved in [14] that the group von Neumann algebra L(G)L(G) of any non-amenable bi-exact icc group GG is prime (i.e. it cannot be decomposed as a tensor product of two II1\rm{II}_{1} factors) by showing that L(G)L(G) is solid. In [18], Ozawa and Popa proved a unique prime factorization theorem which states that if G1,,GnG_{1},\cdots,G_{n} are non-amenable bi-exact icc groups and L(G1)¯¯L(Gn)L(G_{1})\overline{\otimes}\cdots\overline{\otimes}L(G_{n}) is decomposed as a tensor product of mm II1\rm{II}_{1} factors with mnm\geq n i.e. L(G1)¯¯L(Gn)=𝒩1¯¯𝒩mL(G_{1})\overline{\otimes}\cdots\overline{\otimes}L(G_{n})=\mathcal{N}_{1}\overline{\otimes}\cdots\overline{\otimes}\mathcal{N}_{m}, then we have m=nm=n and after permutation of indices and unitary conjugation, each 𝒩i\mathcal{N}_{i} is isomorphic to an amplification of L(Gi)L(G_{i}). Subsequently, various rigidity results were proved for bi-exact groups. In particular, Sako showed measure equivalence rigidity for direct product (cf. [20]) and Chifan and Ioana showed rigidity result in von Neumann algebra sense for amalgamated free product (cf. [3]).

It is known that the class of bi-exact groups contains amenable groups, hyperbolic groups, discrete subgroups of connected simple Lie groups of rank one, and 2SL2()\mathbb{Z}^{2}\rtimes SL_{2}(\mathbb{Z}) (cf. [2][16][17]). On the other hand, unlike exactness, bi-exactness is not preserved under some basic group theoretic constructions. For example, direct products and increasing unions of bi-exact groups are not necessarily bi-exact (e.g. F2×F2F_{2}\times F_{2} and G×F2G\times F_{2}, where F2F_{2} is a free group of rank 2 and GG is an infinite countable locally finite group, are exact but not bi-exact).

The standard way of proving bi-exactness is using topologically amenable action on a special compact space called ‘boundary small at infinity’. Using this method, Ozawa proved in [15] that any finitely generated group hyperbolic relative to amenable peripheral subgroups is bi-exact. In this paper, we use a different approach and generalize Ozawa’s result by proving the following.

Theorem 1.1.

Suppose that GG is a finitely generated group hyperbolic relative to a collection of subgroups {Hμ}μΛ\{H_{\mu}\}_{\mu\in\Lambda} of GG. Then, GG is bi-exact if and only if all subgroups HμH_{\mu} are bi-exact.

The ‘Only if’ direction is obvious and the whole paper is devoted to proving the ‘if’ direction. In relation to this result, it is worth mentioning the following folklore conjecture.

Conjecture 1.2.

Suppose that a finitely generated group GG is hyperbolic relative to a collection of subgroups {Hμ}μΛ\{H_{\mu}\}_{\mu\in\Lambda} of GG. If all subgroups HμH_{\mu} are exact, then GG is bi-exact relative to {Hμ}μΛ\{H_{\mu}\}_{\mu\in\Lambda}.

For the definition of relative bi-exactness, the reader is referred to Definition 15.1.2 of [2]. Note that the ‘if’ direction of Theorem 1.1 is not a weak version of Conjecture 1.2. Indeed, while the assumption of Theorem 1.1 is stronger, the conclusion is also stronger.

Our proof of Theorem 1.1 is based on a characterization of bi-exactness using exactness and the existence of a proper array (cf. Proposition 2.2) and uses two technologies related to relatively hyperbolic groups. The first one is a bicombing of fine hyperbolic graphs, which was constructed by Mineyev and Yaman in [10] using the same idea as in [9]. The second one is based on the notion of separating cosets of hyperbolically embedded subgroups, which was introduced by Hull and Osin in [8] and developed further by Osin in [12]. We will construct two arrays using each of these techniques and combine them to make a proper array. It is worth noting that the Mineyev-Yaman’s bicombing of relatively hyperbolic groups alone is not sufficient to derive Conjecture 1.2 for the reason explained in Remark 3.7. Therefore, Conjecture 1.2 is still considered open.

The paper is organized as follows. In Section 2, we discuss the necessary definitions and known results about bi-exact groups and relatively hyperbolic groups. In Section 3.1, we give an outline of our proof. Section 3.2 and 3.3 discuss the constructions of arrays based on ideas of [10] and [8], respectively, and Section 3.4 provides the proof of Theorem 1.1.

Acknowledgment. I thank Denis Osin for introducing this topic to me, for sharing his ideas, and for many helpful discussions.

2 Preliminary

2.1 Bi-exact groups

In this section, we introduce some equivalent conditions of bi-exact groups. The definition of an array given below was suggested in [4].

Definition 2.1.

Suppose that GG is a group, 𝒦\mathcal{K} a Hilbert space, and π:G𝒰(𝒦)\pi\colon G\to\mathcal{U}(\mathcal{K}) a unitary representation. A map r:G𝒦r\colon G\to\mathcal{K} is called an array on GG into (𝒦,π)(\mathcal{K},\pi), if rr satisfies (1) and (2) below. When there exists such rr, we say that GG admits an array into (𝒦,π)(\mathcal{K},\pi).

  • (1)

    πg(r(g1))=r(g)\pi_{g}(r(g^{-1}))=-r(g) for all gGg\in G.

  • (2)

    For every gGg\in G, we have suphGr(gh)πg(r(h))<\sup_{h\in G}\|r(gh)-\pi_{g}(r(h))\|<\infty.

If, in addition, rr satisfies (3) below, rr is called proper.

  • (3)

    For any NN\in\mathbb{N}, {gGr(g)N}\{g\in G\mid\|r(g)\|\leq N\} is finite.

Conditions (2) and (3) in Proposition 2.2 are simplified versions of Proposition 2.3 of [4] and Proposition 2.7 of [19] respectively.

Proposition 2.2.

For any countable group GG, the following three conditions are equivalent.

  • (1)

    GG is bi-exact.

  • (2)

    GG is exact and admits a proper array into the left regular representation (2(G),λG)(\ell^{2}(G),\lambda_{G}).

  • (3)

    GG is exact, and there exist an orthogonal representation η:G𝒪(K)\eta\colon G\to\mathcal{O}(K_{\mathbb{R}}) to a real Hilbert space KK_{\mathbb{R}} that is weakly contained in the regular representation of GG and a map c:GKc\colon G\to K_{\mathbb{R}} that is proper and satisfies

    supkGc(gkh)ηgc(k)<\sup_{k\in G}\|c(gkh)-\eta_{g}c(k)\|<\infty

    for all g,hGg,h\in G.

2.2 Relatively hyperbolic groups

There are several equivalent definitions of relatively hyperbolic groups due to Gromov, Bowditch, Farb, and Osin. We adopt Farb’s definition with the Bounded Coset Penetration property replaced by fineness of coned-off Cayley graphs. [7], [5, Appendix], and [11, Appendix] contain proofs of the equivalence of these definitions. To state our definition, we first prepare some terminologies. When we consider a graph YY without loops or multiple edges, the edge set E(Y)E(Y) is defined as a irreflexive symmetric subset of V(Y)×V(Y)V(Y)\times V(Y), where V(Y)V(Y) is the vertex set (see Section 2.3 for details).

Definition 2.3 (Coned-off Cayley graph).

Suppose that GG is a finitely generated group with a finite generating set XX and {Hλ}λΛ\{H_{\lambda}\}_{\lambda\in\Lambda} is a collection of subgroups of GG. Let Γ(G,X)\Gamma(G,X) be the Cayley graph of GG with respect to XX, in which we don’t allow loops nor multiple edges. More precisely, its vertex set and edge set are defined by

V(Γ(G,X))=G,E(Γ(G,X))={(g,gs),(gs,g)gG,sX{1}}.V(\Gamma(G,X))=G,\;\;\;E(\Gamma(G,X))=\{(g,gs),(gs,g)\mid g\in G,s\in X\setminus\{1\}\}. (1)

We define a new graph Γ^\widehat{\Gamma} whose vertex set V(Γ^)V(\widehat{\Gamma}) and edge set E(Γ^)E(\widehat{\Gamma}) are given by

V(Γ^)=GλΛG/Hλ,V(\widehat{\Gamma})=G\sqcup\bigsqcup_{\lambda\in\Lambda}G/H_{\lambda},
E(Γ^)=E(Γ(G,X))λΛ{(g,xHλ),(xHλ,g)xHλG/Hλ,gxHλ}.E(\widehat{\Gamma})=E(\Gamma(G,X))\sqcup\bigsqcup_{\lambda\in\Lambda}\{(g,xH_{\lambda}),(xH_{\lambda},g)\mid xH_{\lambda}\in G/H_{\lambda}\ ,g\in xH_{\lambda}\}.

The graph Γ^\widehat{\Gamma} is called coned-off Cayley graph with respect to XX.

We set the length of any edge of Γ^\widehat{\Gamma} to be 1. GG acts on Γ^\widehat{\Gamma} naturally by graph automorphism. For a vertex vV(Γ^)v\in V(\widehat{\Gamma}), we denote its stabilizer by Gv={gGgv=v}G_{v}=\{g\in G\mid gv=v\}.

Definition 2.4 (Hyperbolic space).

A geodesic metric space (X,d)(X,d) is called hyperbolic if there exists δ[0,)\delta\in[0,\infty) such that for any x,y,zXx,y,z\in X, any geodesic path [x,y],[x,z],[z,y][x,y],[x,z],[z,y], and any point a[x,y]a\in[x,y], there exists b[x,z][z,y]b\in[x,z]\cup[z,y] satisfying d(a,b)δd(a,b)\leq\delta.

Definition 2.5 (Circuit).

A closed path pp in a graph is called circuit, if pp doesn’t have self-intersection except its initial and terminal vertices.

Definition 2.6 (Fine graph).

A graph YY is called fine, if for any nn\in\mathbb{N} and any edge ee of YY, the number of circuits of length nn containing ee is finite.

The following is the definition of relatively hyperbolic groups on which we work. The reader is referred to [5, Appendix] for the equivalence of Definition 2.7 and the other definitions.

Definition 2.7.

A finitely generated group GG is called hyperbolic relative to a collection of subgroups {Hλ}λΛ\{H_{\lambda}\}_{\lambda\in\Lambda}, if Λ\Lambda is finite and for some (equivalently, any) finite generating set XX of GG, the coned-off Cayley graph Γ^\widehat{\Gamma} is fine and hyperbolic. A member of the collection of subgroups {Hλ}λΛ\{H_{\lambda}\}_{\lambda\in\Lambda} is called a peripheral subgroup.

Some definitions (e.g. Osin’s definition in [11]) can be stated without requiring Λ\Lambda to be finite. In Osin’s definition, finiteness of Λ\Lambda follows if GG is finitely generated.

2.3 Mineyev-Yaman’s bicombing

In [10], Mineyev and Yaman constructed a bicombing of a fine hyperbolic graph using the same idea as in [9]. We first discuss 1-chains on a graph and a unitary representation of a group acting on a graph in general and then introduce Mineyev-Yaman’s bicombing.

In this section, we consider graphs without loops nor multiple edges, following [10] and [1]. More precisely, for a graph YY with a vertex set V(Y)V(Y), its edge set E(Y)E(Y) is a subset of V(Y)×V(Y){(v,v)vV(Y)}V(Y)\times V(Y)\setminus\{(v,v)\mid v\in V(Y)\} such that E(Y)¯=E(Y)\overline{E(Y)}=E(Y), where we define (u,v)¯=(v,u)\overline{(u,v)}=(v,u) for any (u,v)V(Y)2(u,v)\in V(Y)^{2}. We call a subset E+(V)E^{+}(V) of E(Y)E(Y) a set of positive edges if it satisfies E(V)=E+(V)E+(V)¯E(V)=E^{+}(V)\sqcup\overline{E^{+}(V)}. Choosing a set of positive edges E+(V)E^{+}(V) means choosing directions of edges. Note that when we consider a group action on a graph, we allow inversion of edges. We begin with auxiliary notations.

For a graph YY, we denote by C1(Y)C_{1}(Y) the set of 1-chains on YY over \mathbb{C}. More precisely, C1(Y)C_{1}(Y) is defined as follows. Consider a direct product E(Y)={eE(Y)cee|ce}\mathbb{C}^{E(Y)}=\Big{\{}\sum_{e\in E(Y)}c_{e}e\;\Big{|}\;c_{e}\in\mathbb{C}\Big{\}} of vector spaces over \mathbb{C}. Note that this summation notation is just a formal sum. We define a quotient space 0(Y)\ell^{0}(Y) by

0(Y)=E(Y)/{eE(Y)ceeE(Y)|ce=ce¯eE(Y)}.\ell^{0}(Y)=\mathbb{C}^{E(Y)}\;\Big{/}\;\Big{\{}\sum_{e\in E(Y)}c_{e}e\in\mathbb{C}^{E(Y)}\;\Big{|}\;c_{e}=c_{\bar{e}}\;\forall e\in E(Y)\Big{\}}.

Here, we denote by [e]=[u,v][e]=[u,v] the element in 0(Y)\ell^{0}(Y) corresponding to e=(u,v)E(Y)e=(u,v)\in\mathbb{C}^{E(Y)} where e=(u,v)E(Y)e=(u,v)\in E(Y). Note that for any e=(u,v)E(Y)e=(u,v)\in E(Y), we have

[e¯]=[v,u]=[u,v]=[e].[\bar{e}]=[v,u]=-[u,v]=-[e].

Fix a set of positive edges E+(Y)E^{+}(Y). The map

E+(Y)eE+(Y)ceeeE+(Y)ce[e]0(Y)\mathbb{C}^{E^{+}(Y)}\ni\sum_{e\in E^{+}(Y)}c_{e}e\mapsto\sum_{e\in E^{+}(Y)}c_{e}[e]\in\ell^{0}(Y)

is an isomorphism of vector spaces. The space of 1-chains is defined by

C1(Y)={eE+(Y)ce[e]0(Y)|FE+(Y),|F|<ce=0eF}.C_{1}(Y)=\Big{\{}\sum_{e\in E^{+}(Y)}c_{e}[e]\in\ell^{0}(Y)\;\Big{|}\;\exists F\subset E^{+}(Y),\;|F|<\infty\;\wedge\;c_{e}=0\;\forall e\notin F\Big{\}}.

Note that C1(Y)C_{1}(Y) is independent of the choice of E+(Y)E^{+}(Y). For a,bV(Y)a,b\in V(Y) and a path q=(v0,v1,,vn)q=(v_{0},v_{1},\cdots,v_{n}) from a=v0a=v_{0} to b=vnb=v_{n}, we define a 1-chain

q=[v0,v1]++[vn1,vn]C1(Y).q=[v_{0},v_{1}]+\cdots+[v_{n-1},v_{n}]\in C_{1}(Y). (2)

Here, we use the same notation qq to denote a path and the corresponding 1-chain by abuse of notation.

For each p[1,)p\in[1,\infty) (in fact, we use only cases p=1,2p=1,2), define a subspace p(Y)\ell^{p}(Y) of 0(Y)\ell^{0}(Y) by

p(Y)={eE+(Y)ce[e]0(Y)|eE+(Y)|ce|p<}\ell^{p}(Y)=\Big{\{}\sum_{e\in E^{+}(Y)}c_{e}[e]\in\ell^{0}(Y)\;\Big{|}\;\sum_{e\in E^{+}(Y)}|c_{e}|^{p}<\infty\Big{\}}

and for each ξ=eE+(Y)ce[e]p(Y)\xi=\sum_{e\in E^{+}(Y)}c_{e}[e]\in\ell^{p}(Y) define a norm

ξp=(eE+(Y)|ce|p)1/p.\|\xi\|_{p}=\left(\sum_{e\in E^{+}(Y)}|c_{e}|^{p}\right)^{1/p}.

(p(Y),p)(\ell^{p}(Y),\|\cdot\|_{p}) becomes a Banach space isomorphic to p(E+(Y))\ell^{p}(E^{+}(Y)). In particular, (2(Y),2)(\ell^{2}(Y),\|\cdot\|_{2}) is a Hilbert space. Note that (p(Y),p)(\ell^{p}(Y),\|\cdot\|_{p}) doesn’t depend on the choice of E+(Y)E^{+}(Y). We also have C1(Y)p(Y)C_{1}(Y)\subset\ell^{p}(Y) for any p[1,)p\in[1,\infty).

Definition 2.8.

Suppose that a group GG acts on YY by graph automorphisms, then GG acts on (p(Y),p)(\ell^{p}(Y),\|\cdot\|_{p}) isometrically by

G×p(Y)(g,(u,v)E+(Y)c(u,v)[u,v])(u,v)E+(Y)c(u,v)[gu,gv]p(Y).G\times\ell^{p}(Y)\ni\Big{(}g,\sum_{(u,v)\in E^{+}(Y)}c_{(u,v)}[u,v]\Big{)}\mapsto\sum_{(u,v)\in E^{+}(Y)}c_{(u,v)}[gu,gv]\in\ell^{p}(Y).

In particular, this action on (2(Y),2)(\ell^{2}(Y),\|\cdot\|_{2}) becomes a unitary representation of GG and we denote this unitary representation by (2(Y),π)(\ell^{2}(Y),\pi).

There exists a unique well-defined linear map

:C1(Y)C0(Y)={vFcvvFV(Y),|F|<,cv}\partial\colon C_{1}(Y)\to C_{0}(Y)=\Big{\{}\sum_{v\in F}c_{v}v\mid F\subset V(Y),\;|F|<\infty,\;c_{v}\in\mathbb{C}\Big{\}}

such that [u,v]=vu\partial[u,v]=v-u for any (u,v)E(Y)(u,v)\in E(Y).

Definition 2.9.

Suppose that YY is a graph. A map q:V(Y)2C1(Y)q\colon V(Y)^{2}\to C_{1}(Y) is called a homological bicombing, if for any (a,b)V(Y)2(a,b)\in V(Y)^{2}, we have q[a,b]=ba\partial q[a,b]=b-a. In particular, if all coefficients of q[a,b]q[a,b] are in \mathbb{Q} for any a,bV(Y)a,b\in V(Y), qq is called a \mathbb{Q}-bicombing.

Definition 2.10 (2-vertex-connectivity).

A graph YY is called 2-vertex-connected, if YY is connected and for any vV(Y)v\in V(Y), Y{v}Y\setminus\{v\} is connected, where Y{v}Y\setminus\{v\} is an induced subgraph of YY whose vertex set is V(Y){v}V(Y)\setminus\{v\}.

Theorem 2.11 is a simplified version of Theorem 47 and Proposition 46 (3) in [10].

Theorem 2.11.

Suppose that YY is a 2-vertex-connected fine hyperbolic graph and GG is a group acting on YY. If the number of GG-orbits of E(Y)E(Y) is finite and the edge stabilizer Ge=GuGvG_{e}=G_{u}\cap G_{v} is finite for any e=(u,v)E(Y)e=(u,v)\in E(Y), then there exists a \mathbb{Q}-bicombing qq of YY satisfying the following conditions.

  • (1)

    qq is GG-equivariant, i.e. q[ga,gb]=gq[a,b]q[ga,gb]=g\cdot q[a,b] for any a,bV(Y)a,b\in V(Y) and any gGg\in G.

  • (2)

    qq is anti-symmetric, i.e. q[b,a]=q[a,b]q[b,a]=-q[a,b] for any a,bV(Y)a,b\in V(Y).

  • (3)

    There exists a constant T0T\geq 0 such that for any a,b,cV(Y)a,b,c\in V(Y),

    q[a,b]+q[b,c]+q[c,a]1T.\|q[a,b]+q[b,c]+q[c,a]\|_{1}\leq T.
  • (4)

    There exist constants M0M^{\prime}\geq 0 and N0N^{\prime}\geq 0 such that for any a,bV(Y)a,b\in V(Y),

    q[a,b]1MdY(a,b)+N,\|q[a,b]\|_{1}\leq M^{\prime}d_{Y}(a,b)+N^{\prime},

    where dYd_{Y} is the graph metric of YY.

Remark 2.12.

By examining the explicit construction of qq in Section 6 of [10], we can see that for any a,bV(Y)a,b\in V(Y), q[a,b]q[a,b] is a convex combination of paths from aa to bb (see (2)), i.e. there exist paths p1,,pnp_{1},\cdots,p_{n} from aa to bb and α1,,αn0\alpha_{1},\cdots,\alpha_{n}\in\mathbb{Q}_{\geq 0} with j=1nαj=1\sum_{j=1}^{n}\alpha_{j}=1 such that

q[a,b]=j=1nαjpj.q[a,b]=\sum_{j=1}^{n}\alpha_{j}p_{j}.
Remark 2.13.

Theorem 2.11 was actually proved for what Mineyev and Yaman called a ‘hyperbolic tuple’ which was defined in their paper as a quadruple consisting of a graph, a group acting on it, a subset of the vertex set, and a set of vertex stabilizers (cf. [10, Definitions 20, 27, 29, 38]). However, the statement and the proof of Theorem 47 and Proposition 46 (3) in [10] does not require the whole quadruple, and we simplify their exposition.

Remark 2.14.

Theorem 2.11 (4) is necessary only for Remark 3.7.

2.4 Separating cosets of hyperbolically embedded subgroups

In this section, we define Hull-Osin’s separating cosets and explain their properties required for our discussion in Section 3.3. The notion of separating cosets of hyperbolically embedded subgroups was first introduced by Hull and Osin in [8] and further developed by Osin in [12]. There is one subtle difference in the definition of separationg cosets in [8] and in [12], which is explained in Remark 2.21, though other terminologies and related propositions are mostly the same between them. With regards to this difference, we follow definitions and notations of [12] in our discussion. In this section, suppose that GG is a group, XX is a subset of GG, and {Hλ}λΛ\{H_{\lambda}\}_{\lambda\in\Lambda} is a collection of subgroups of GG such that X(λΛHλ)X\cup(\bigcup_{\lambda\in\Lambda}H_{\lambda}) generates GG. Note that Λ\Lambda and XX are possibly infinite. We begin with defining some auxiliary concepts.

Let =λΛ(Hλ{1})\mathcal{H}=\bigsqcup_{\lambda\in\Lambda}(H_{\lambda}\setminus\{1\}). Note that this union is disjoint as sets of labels, not as subsets of GG. Let Γ(G,X)\Gamma(G,X\cup\mathcal{H}) be the Cayley graph of GG with respect to XX\sqcup\mathcal{H}, which allows loops and multiple edges, that is, its vertex set is GG and its positive edge set is G×(X)G\times(X\sqcup\mathcal{H}). We call Γ(G,X)\Gamma(G,X\cup\mathcal{H}) the relative Cayley graph.

For each λΛ\lambda\in\Lambda, we consider the Cayley graph Γ(Hλ,Hλ{1})\Gamma(H_{\lambda},H_{\lambda}\setminus\{1\}), which is a subgraph of Γ(G,X)\Gamma(G,X\cup\mathcal{H}), and define a metric dλ^:Hλ×Hλ[0,]\widehat{d_{\lambda}}\colon H_{\lambda}\times H_{\lambda}\to[0,\infty] as follows. We say that a path pp in Γ(G,X)\Gamma(G,X\cup\mathcal{H}) is λ\lambda-admissible, if pp doesn’t contain any edge of Γ(Hλ,Hλ{1})\Gamma(H_{\lambda},H_{\lambda}\setminus\{1\}). Note that pp can contain an edge whose label is an element of HλH_{\lambda} (e.g. the case when the initial vertex of the edge is not in HλH_{\lambda}) and also pp can pass vertices of Γ(Hλ,Hλ{1})\Gamma(H_{\lambda},H_{\lambda}\setminus\{1\}). For f,gHλf,g\in H_{\lambda}, we define dλ^\widehat{d_{\lambda}} to be the minimum of lengths of all λ\lambda-admissible paths from ff to gg. If there is no λ\lambda-admissible path from ff to gg, then we define dλ^(f,g)=\widehat{d_{\lambda}}(f,g)=\infty. For convenience, we extend dλ^\widehat{d_{\lambda}} to dλ^:G×G[0,]\widehat{d_{\lambda}}\colon G\times G\to[0,\infty] by defining dλ^(f,g)=dλ^(1,f1g)\widehat{d_{\lambda}}(f,g)=\widehat{d_{\lambda}}(1,f^{-1}g) if f1gHλf^{-1}g\in H_{\lambda} and dλ^(f,g)=\widehat{d_{\lambda}}(f,g)=\infty otherwise.

Definition 2.15.

Suppose that GG is a group and XX is a subset of GG. The collection of subgroups {Hλ}λΛ\{H_{\lambda}\}_{\lambda\in\Lambda} of GG is said to be hyperbolically embedded in (G,X)(G,X), if it satisfies the two conditions below.

  • (1)

    X(λΛHλ)X\cup(\bigcup_{\lambda\in\Lambda}H_{\lambda}) generates GG and Γ(G,X)\Gamma(G,X\cup\mathcal{H}) is hyperbolic.

  • (2)

    For any λΛ\lambda\in\Lambda, (Hλ,dλ^)(H_{\lambda},\widehat{d_{\lambda}}) is locally finite, i.e. for any nn\in\mathbb{N}, {gHλdλ^(1,g)n}\{g\in H_{\lambda}\mid\widehat{d_{\lambda}}(1,g)\leq n\} is finite.

The following is a simplified version of Proposition 4.28 of [6].

Proposition 2.16.

Suppose that GG is a finitely generated group hyperbolic relative to a collection of subgroups {Hμ}μΛ\{H_{\mu}\}_{\mu\in\Lambda} of GG. Then, for any finite generating set XX of GG, {Hλ}λΛ\{H_{\lambda}\}_{\lambda\in\Lambda} is hyperbolically embedded in (G,X)(G,X).

In the remainder of this section, suppose that {Hλ}λΛ\{H_{\lambda}\}_{\lambda\in\Lambda} is hyperbolically embedded in (G,X)(G,X).

Definition 2.17.

[12, Definition 4.1] Suppose that pp is a path in the relative Cayley graph Γ(G,X)\Gamma(G,X\cup\mathcal{H}). A subpath qq of pp is called an HλH_{\lambda}-subpath if the labels of all edges of qq are in HλH_{\lambda}. In the case pp is a closed path, qq can be a subpath of any cyclic shift of pp. An HλH_{\lambda}-subpath qq of a path pp is called HλH_{\lambda}-component if qq is not contained in any longer HλH_{\lambda}-subpath of pp. In the case pp is a closed path, we require that qq is not contained in any longer HλH_{\lambda}-subpath of any cyclic shift of pp. Further, by a component, we mean an HλH_{\lambda}-component for some HλH_{\lambda}. Two HλH_{\lambda}-components q1,q2q_{1},q_{2} of a path pp is called connected, if all vertices of q1q_{1} and q2q_{2} are in the same HλH_{\lambda}-coset. An HλH_{\lambda}-component qq of a path pp is called isolated, if qq is not connected to any other HλH_{\lambda}-component of pp.

Remark 2.18.

Note that all vertices of an HλH_{\lambda}-component lie in the same HλH_{\lambda}-coset.

The following proposition is important, which is a particular case of Proposition 4.13 of [6].

Proposition 2.19.

[8, Lemma 2.4] There exists a constant C>0C>0 such that for any geodesic nn-gon pp in Γ(G,X)\Gamma(G,X\cup\mathcal{H}) and any isolated HλH_{\lambda}-component aa of pp, we have

dλ^(a,a+)nC.\widehat{d_{\lambda}}(a_{-},a_{+})\leq nC.

In what follows, we fix any constant D>0D>0 with

D3C.D\geq 3C. (3)

We can now define separating cosets.

Definition 2.20.

[12, Definition 4.3] A path pp in Γ(G,X)\Gamma(G,X\cup\mathcal{H}) is said to penetrate a coset xHλxH_{\lambda} for some λΛ\lambda\in\Lambda, if pp decomposes as p1ap2p_{1}ap_{2}, where p1,p2p_{1},p_{2} are possibly trivial, aa is an HλH_{\lambda}-component, and axHλa_{-}\in xH_{\lambda}. Note that if pp is a geodesic, pp penetrates any coset of HλH_{\lambda} at most once. In this case, aa is called the component of pp corresponding to xHλxH_{\lambda} and also the vertices aa_{-} and a+a_{+} are called the entrance and exit points of pp and are denoted by pin(xHλ)p_{in}(xH_{\lambda}) and pout(xHλ)p_{out}(xH_{\lambda}) respectively. If in addition, we have dλ^(a,a+)>D\widehat{d_{\lambda}}(a_{-},a_{+})>D, then we say that pp essentially penetrates xHλxH_{\lambda}. For f,gGf,g\in G and λΛ\lambda\in\Lambda, if there exists a geodesic path from ff to gg in Γ(G,X)\Gamma(G,X\cup\mathcal{H}) which essentially penetrates an HλH_{\lambda}-coset xHλxH_{\lambda}, then xHλxH_{\lambda} is called an (f,g;D)(f,g;D)-separating coset. We denote the set of (f,g;D)(f,g;D)-separating HλH_{\lambda}-cosets by Sλ(f,g;D)S_{\lambda}(f,g;D).

Remark 2.21.

In Definition 3.1 of [8], whenever f,gGf,g\in G are in the same HλH_{\lambda}-coset xHλxH_{\lambda} for some λΛ\lambda\in\Lambda, xHλxH_{\lambda} is included in Sλ(f,g;D)S_{\lambda}(f,g;D), but in our Definition 2.20, Sλ(f,g;D)S_{\lambda}(f,g;D) can be empty even in this case.

The following lemma is immediate from the above definition.

Lemma 2.22.

[8, Lemma 3.2] For any f,g,hGf,g,h\in G and any λΛ\lambda\in\Lambda, the following holds.

  • (a)

    Sλ(f,g;D)=Sλ(g,f;D)S_{\lambda}(f,g;D)=S_{\lambda}(g,f;D).

  • (b)

    Sλ(hf,hg;D)={hxHλxHλSλ(f,g;D)}S_{\lambda}(hf,hg;D)=\{hxH_{\lambda}\mid xH_{\lambda}\in S_{\lambda}(f,g;D)\}.

We state some nice properties of separating cosets, all of which were proven in [8]. For f,gGf,g\in G, we denote by 𝒢(f,g)\mathcal{G}(f,g) the set of all geodesic paths in Γ(G,X)\Gamma(G,X\cup\mathcal{H}) from ff to gg.

Lemma 2.23.

[8, Lemma 3.3] For any λΛ\lambda\in\Lambda, any f,gGf,g\in G, and any (f,g;D)(f,g;D)-separating coset xHλxH_{\lambda}, the following holds.

  • (a)

    Every path in Γ(G,X)\Gamma(G,X\cup\mathcal{H}) connecting ff to gg and composed of at most 2 geodesics penetrates xHλxH_{\lambda}.

  • (b)

    For any p,q𝒢(f,g)p,q\in\mathcal{G}(f,g), we have

    dλ^(pin(xHλ),qin(xHλ))3Canddλ^(pout(xHλ),qout(xHλ))3C.\widehat{d_{\lambda}}(p_{in}(xH_{\lambda}),q_{in}(xH_{\lambda}))\leq 3C\;\;and\;\;\widehat{d_{\lambda}}(p_{out}(xH_{\lambda}),q_{out}(xH_{\lambda}))\leq 3C.

Proof of Lemma 2.23 is the same as Lemma 3.3 of [8], though their statements are slightly different. Actually, in Lemma 2.23, we don’t need to assume f1gHλf^{-1}g\notin H_{\lambda}. This difference comes from the difference of definitions of separating cosets, which was mentioned in Remark 2.21.

Corollary 2.24.

[8, Corollary 3.4] For any f,gGf,g\in G and any p𝒢(f,g)p\in\mathcal{G}(f,g), we have Sλ(f,g;D)Pλ(p)S_{\lambda}(f,g;D)\subset P_{\lambda}(p), where we define Pλ(p)P_{\lambda}(p) to be the set of all HλH_{\lambda}-cosets which pp penetrates. In particular, we have |Sλ(f,g;D)|dX(f,g)|S_{\lambda}(f,g;D)|\leq d_{X\cup\mathcal{H}}(f,g), hence Sλ(f,g;D)S_{\lambda}(f,g;D) is finite.

The following lemma makes Sλ(f,g;D)S_{\lambda}(f,g;D) into a totally ordered set.

Lemma 2.25.

[8, Lemma 3.5] Suppose f,gGf,g\in G and that p𝒢(f,g)p\in\mathcal{G}(f,g) penetrates an HλH_{\lambda}-coset xHλxH_{\lambda} and decomposes as p=p1ap2p=p_{1}ap_{2}, where p1,p2p_{1},p_{2} are possibly trivial and aa is an HλH_{\lambda}-component corresponding to xHλxH_{\lambda}. Then, we have dX(f,a)=dX(f,xHλ)d_{X\cup\mathcal{H}}(f,a_{-})=d_{X\cup\mathcal{H}}(f,xH_{\lambda}).

Definition 2.26.

[8, Definition 3.6] Given any f,gGf,g\in G, a relation \preceq on the set Sλ(f,g;D)S_{\lambda}(f,g;D) is defined as follows.

xHλyHλdX(f,xHλ)dX(f,yHλ).xH_{\lambda}\preceq yH_{\lambda}\iff d_{X\cup\mathcal{H}}(f,xH_{\lambda})\leq d_{X\cup\mathcal{H}}(f,yH_{\lambda}).

Next, we define a set of pairs of entrance and exit points of a separating coset. That is, for f,gGf,g\in G, λΛ\lambda\in\Lambda, and xHλSλ(f,g;D)xH_{\lambda}\in S_{\lambda}(f,g;D), we define

E(f,g;xHλ,D)={(pin(xHλ),pout(xHλ))p𝒢(f,g)}.E(f,g;xH_{\lambda},D)=\{(p_{in}(xH_{\lambda}),p_{out}(xH_{\lambda}))\mid p\in\mathcal{G}(f,g)\}.

Note that because xHλxH_{\lambda} is a (f,g;D)(f,g;D)-separating coset, any geodesic from ff to gg penetrates xHλxH_{\lambda} by Lemma 2.23 (a).

Lemma 2.27.

[8, Lemma 3.8] For any f,g,hGf,g,h\in G, λΛ\lambda\in\Lambda, and xHλSλ(f,g;D)xH_{\lambda}\in S_{\lambda}(f,g;D), the following holds.

  • (a)

    E(g,f;xHλ,D)={(v,u)(u,v)E(f,g;xHλ,D)}E(g,f;xH_{\lambda},D)=\{(v,u)\mid(u,v)\in E(f,g;xH_{\lambda},D)\}.

  • (b)

    E(hf,hg;hxHλ,D)={(hu,hv)(u,v)E(f,g;xHλ,D)}E(hf,hg;hxH_{\lambda},D)=\{(hu,hv)\mid(u,v)\in E(f,g;xH_{\lambda},D)\}.

  • (c)

    |E(f,g;xHλ,D)|<|E(f,g;xH_{\lambda},D)|<\infty.

Lemma 2.28.

For any f,g,hGf,g,h\in G and any xHλSλ(f,g;D)xH_{\lambda}\in S_{\lambda}(f,g;D), xHλxH_{\lambda} is either penetrated by all geodesics from ff to hh or penetrated by all geodesics from hh to gg.

Proof.

Suppose there exists a geodesic q𝒢(f,h)q\in\mathcal{G}(f,h) which doesn’t penetrate xHλxH_{\lambda}. For any geodesic r𝒢(h,g)r\in\mathcal{G}(h,g), we can apply Lemma 2.23 (a) to a path qrqr and conclude rr penetrates xHλxH_{\lambda}. ∎

The following lemma is crucial in constructing an array that satisfies the bounded area condition.

Lemma 2.29.

[8, Lemma 3.9] For any f,g,hGf,g,h\in G and any λΛ\lambda\in\Lambda, Sλ(f,g;D)S_{\lambda}(f,g;D) can be decomposed as Sλ(f,g;D)=SS′′FS_{\lambda}(f,g;D)=S^{\prime}\sqcup S^{\prime\prime}\sqcup F where

  • (a)

    SSλ(f,h;D)Sλ(h,g;D)S^{\prime}\subset S_{\lambda}(f,h;D)\setminus S_{\lambda}(h,g;D) and we have E(f,g;xHλ,D)=E(f,h;xHλ,D)E(f,g;xH_{\lambda},D)=E(f,h;xH_{\lambda},D) for any xHλSxH_{\lambda}\in S^{\prime},

  • (b)

    S′′Sλ(h,g;D)Sλ(f,h;D)S^{\prime\prime}\subset S_{\lambda}(h,g;D)\setminus S_{\lambda}(f,h;D) and we have E(f,g;xHλ,D)=E(h,g;xHλ,D)E(f,g;xH_{\lambda},D)=E(h,g;xH_{\lambda},D) for any xHλS′′xH_{\lambda}\in S^{\prime\prime},

  • (c)

    |F|2|F|\leq 2.

3 Main theorem

3.1 Overview of proof

In this section, we explain the idea of the construction of a proper array in the proof of Proposition 3.20 (see (9) (10)) by considering two particular cases.

First, suppose that GG is a hyperbolic group i.e., the collection of peripheral subgroups is empty. We take a symmetric finite generating set X0X_{0} of GG with 1X01\in X_{0} and define

X=X02={ghg,hX0}.X=X_{0}^{2}=\{gh\mid g,h\in X_{0}\}.

The Cayley graph Γ(G,X)\Gamma(G,X), as defined by (1), has no loops or multiple edges and is a 2-vertex-connected locally finite hyperbolic graph. Let YY be the barycentric subdivision of Γ(G,X)\Gamma(G,X). Note that YY is fine, GG acts on YY without inversion of edges, and all edge stabilizers are trivial. By Theorem 2.11, there exists a GG-equivariant anti-symmetric \mathbb{Q}-bicombing qq of YY and a constant T0T\geq 0 such that for any a,b,cV(Y)a,b,c\in V(Y), we have

q[a,b]+q[b,c]+q[c,a]1T.\|q[a,b]+q[b,c]+q[c,a]\|_{1}\leq T.

We define a map Q:G2(Y)Q\colon G\to\ell^{2}(Y) by (see Definition 3.4 for details of this definition)

Q(g)=q[1,g]~.Q(g)=\widetilde{q[1,g]}.

It is straightforward to check that QQ is an array into (2(Y),π)(\ell^{2}(Y),\pi) (cf. Definition 2.8) by using Lemma 3.5 (2). Also, we have

dY(1,g)q[1,g]1=Q(g)22d_{Y}(1,g)\leq\|q[1,g]\|_{1}=\|Q(g)\|_{2}^{2}

(see Remark 2.12, Lemma 3.3, and Lemma 3.5 (1)). Since YY is a barycentric subdivision of Γ(G,X)\Gamma(G,X), we have dY(1,g)=2dX(1,g)d_{Y}(1,g)=2d_{X}(1,g) for any gGg\in G, where dXd_{X} is a word metric on GG with respect to XX. Hence, for any NN\in\mathbb{N}, if gGg\in G satisfies Q(g)2N\|Q(g)\|_{2}\leq N, we have

dX(1,g)=12dY(1,g)12Q(g)2212N2.d_{X}(1,g)=\frac{1}{2}d_{Y}(1,g)\leq\frac{1}{2}\|Q(g)\|_{2}^{2}\leq\frac{1}{2}N^{2}.

This implies that QQ is proper. Combined with Proposition 2.2 (3), this gives another proof of the fact that hyperbolic groups are bi-exact. Here, note that (2(Y),π)(\ell^{2}(Y),\pi) is a direct sum of copies (2(G),λG)(\ell^{2}(G),\lambda_{G}), because GG acts on YY without inversion of edges and its action on E(Y)E(Y) is free. Hence, it is weakly contained by (2(G),λG)(\ell^{2}(G),\lambda_{G}). We can also use Mineyev’s bicombing in [9] instead of Theorem 2.11 in this case.

Second, suppose that GG is a free product of bi-exact groups H1H_{1} and H2H_{2}. In this case, {H1,H2}\{H_{1},H_{2}\} are peripheral subgroups and we can construct a proper array on GG using the normal forms of elements of the free product and proper arrays on H1H_{1} and H2H_{2} as follows. Let ri:Hi2(Hi)r_{i}\colon H_{i}\to\ell^{2}(H_{i}) be a proper array into the left regular representation (2(Hi),λHi)(\ell^{2}(H_{i}),\lambda_{H_{i}}) for each i=1,2i=1,2. We can assume ri(a)21\|r_{i}(a)\|_{2}\geq 1 for any aHi{1}a\in H_{i}\setminus\{1\} by changing values of rir_{i} on a finite subset of HiH_{i}, if necessary. Also, we regard each rir_{i} as a map from HiH_{i} to 2(G)\ell^{2}(G) by composing it with the embedding 2(Hi)2(G)\ell^{2}(H_{i})\hookrightarrow\ell^{2}(G). For gG{1}g\in G\setminus\{1\}, without loss of generality, let g=h1k1hnkng=h_{1}k_{1}\cdots h_{n}k_{n} be the normal form of gg, where h1,,hnH1{1}h_{1},\cdots,h_{n}\in H_{1}\setminus\{1\} and k1,,knH2{1}k_{1},\cdots,k_{n}\in H_{2}\setminus\{1\}. We define a map R1:G2(G)R_{1}\colon G\to\ell^{2}(G) by

R1(g)=λG(1)r1(h1)+λG(h1k1)r1(h2)++λG(h1k2hn1kn1)r1(hn).R_{1}(g)=\lambda_{G}(1)r_{1}(h_{1})+\lambda_{G}(h_{1}k_{1})r_{1}(h_{2})+\cdots+\lambda_{G}(h_{1}k_{2}\cdots h_{n-1}k_{n-1})r_{1}(h_{n}).

Here, we define R1(1)=0R_{1}(1)=0. In the same way, we define a map R2:G2(G)R_{2}\colon G\to\ell^{2}(G) from r2r_{2}. It is not difficult to show that the map R:G2(G)2(G)R\colon G\to\ell^{2}(G)\oplus\ell^{2}(G) defined by

R(g)=(R1(g),R2(g))R(g)=(R_{1}(g),R_{2}(g))

is a proper array into (2(G)2(G),λGλG)(\ell^{2}(G)\oplus\ell^{2}(G),\lambda_{G}\oplus\lambda_{G}).

The case of general relatively hyperbolic groups is a combination of the above two cases. To every finitely generated relatively hyperbolic group, we associate two arrays. The first one is constructed in Section 3.2 starting with a fine hyperbolic graph as in the case of a hyperbolic group above. The second array is a generalized version of the array RR for G=H1H2G=H_{1}*H_{2}. Hull-Osin’s separating cosets play the same role as syllables in the normal form in the free product case. One technical remark is that the condition ‘ri(a)21aHi{1}\|r_{i}(a)\|_{2}\geq 1\;\forall a\in H_{i}\setminus\{1\}’ in the free product case is used to prevent the set {gGR(g)=0}\{g\in G\mid\|R(g)\|=0\} from becoming infinite, but in the proof of Proposition 3.20, we don’t need this condition thanks to the first array.

3.2 First array

This section is basically a continuation of Section 2.3. Notation and terminology related to graphs follow those in Section 2.3. The goal of this section is to prove Proposition 3.1 by using Mineyev-Yaman’s bicombing.

Proposition 3.1.

Suppose that GG is a finitely generated group hyperbolic relative to a collection of subgroups {Hλ}λΛ\{H_{\lambda}\}_{\lambda\in\Lambda} of GG. Then, there exist a finite generating set XX of GG, a 2-vertex-connected fine hyperbolic graph YY on which GG acts without inversion of edges, and an array Q:G2(Y)Q\colon G\to\ell^{2}(Y) into (2(Y),π)(\ell^{2}(Y),\pi) that satisfy the following conditions.

  • (1)

    The edge stabilizer is trivial for any edge in E(Y)E(Y).

  • (2)

    For any gGg\in G, we have

    dΓ^(1,g)12Q(g)22,d_{\widehat{\Gamma}}(1,g)\leq\frac{1}{2}\|Q(g)\|_{2}^{2}, (4)

    where Γ^\widehat{\Gamma} is the coned-off Cayley graph of GG with respect to XX.

Remark 3.2.

The unitary representation (2(Y),π)(\ell^{2}(Y),\pi) in Proposition 3.1 is weakly contained by the left regular representation (2(G),λG)(\ell^{2}(G),\lambda_{G}). Indeed, since GG acts on YY without inversion of edges and all edge stabilizers are trivial, (2(Y),π)(\ell^{2}(Y),\pi) is a direct sum of copies of (2(G),λG)(\ell^{2}(G),\lambda_{G}).

We first prove a few general lemmas about graphs. Lemma 3.3 can be proven for graphs with loops and multiple edges as well exactly in the same way, but we stick to our current setting.

Lemma 3.3.

Suppose that YY is a connected graph without loops or multiple edges. If a,bV(Y)a,b\in V(Y) are two vertices, (pj)j=1NC1(Y)(p_{j})_{j=1}^{N}\subset C_{1}(Y) are paths from aa to bb as 1-chains, and (αj)j=1N(\alpha_{j})_{j=1}^{N}\subset\mathbb{C} are complex numbers, then we have

|j=1Nαj|dY(a,b)j=1Nαjpj1.\left|\sum_{j=1}^{N}\alpha_{j}\right|\cdot d_{Y}(a,b)\leq\left\|\sum_{j=1}^{N}\alpha_{j}p_{j}\right\|_{1}.
Proof.

We have |dY(o(e),a)dY(t(e),a)|dY(o(e),t(e))1|d_{Y}(o(e),a)-d_{Y}(t(e),a)|\leq d_{Y}(o(e),t(e))\leq 1 for any e=(u,v)E(Y)e=(u,v)\in E(Y), where we define o(e)=uo(e)=u and t(e)=vt(e)=v. Hence, by defining

C={eE(Y)dY(o(e),a)=dY(t(e),a)}C=\{e\in E(Y)\mid d_{Y}(o(e),a)=d_{Y}(t(e),a)\}

and

Dn={\displaystyle D_{n}=\{ eE(Y)dY(o(e),a)=n1dY(t(e),a)=n}\displaystyle e\in E(Y)\mid d_{Y}(o(e),a)=n-1\wedge d_{Y}(t(e),a)=n\}
{eE(Y)dY(t(e),a)=n1dY(o(e),a)=n}\displaystyle\cup\{e\in E(Y)\mid d_{Y}(t(e),a)=n-1\wedge d_{Y}(o(e),a)=n\}

for each nn\in\mathbb{N}, we get the decomposition of edges E(Y)=C(nDn)E(Y)=C\sqcup\left(\bigsqcup_{n\in\mathbb{N}}D_{n}\right). Also, we can choose directions of edges so that they point outward from aa, that is, there exists a set of positive edges E+(Y)E^{+}(Y) such that

DnE+(Y)={eE(Y)dY(o(e),a)=n1dY(t(e),a)=n}D_{n}\cap E^{+}(Y)=\{e\in E(Y)\mid d_{Y}(o(e),a)=n-1\wedge d_{Y}(t(e),a)=n\}

and

DnE(Y)={eE(Y)dY(t(e),a)=n1dY(o(e),a)=n}D_{n}\cap E^{-}(Y)=\{e\in E(Y)\mid d_{Y}(t(e),a)=n-1\wedge d_{Y}(o(e),a)=n\}

for all nn\in\mathbb{N}, where E(Y)=E+(Y)¯E^{-}(Y)=\overline{E^{+}(Y)}. For simplicity, we use the notation C+=CE+(Y)C^{+}=C\cap E^{+}(Y), Dn+=DnE+(Y)D_{n}^{+}=D_{n}\cap E^{+}(Y), and Dn=DnE(Y)D_{n}^{-}=D_{n}\cap E^{-}(Y). Given a 1-chain ξ=eE+(Y)ce[e]C1(Y)\xi=\sum_{e\in E^{+}(Y)}c_{e}[e]\in C_{1}(Y), we define linear maps PrC+,PrDn+:C1(Y)C1(Y)(n)Pr_{C^{+}},Pr_{D_{n}^{+}}\colon C_{1}(Y)\to C_{1}(Y)\;(n\in\mathbb{N}) by

PrC+(ξ)=eC+ce[e]andPrDn+(ξ)=eDn+ce[e].Pr_{C^{+}}(\xi)=\sum_{e\in C^{+}}c_{e}[e]\;\;\;{\rm and}\;\;\;Pr_{D_{n}^{+}}(\xi)=\sum_{e\in D_{n}^{+}}c_{e}[e].

By E+(Y)=C+(nDn+)E^{+}(Y)=C^{+}\sqcup\left(\bigsqcup_{n\in\mathbb{N}}D^{+}_{n}\right), we have for any ξC1(Y)\xi\in C_{1}(Y),

ξ=PrC+(ξ)+nPrDn+(ξ)andξ1=PrC+(ξ)1+nPrDn+(ξ)1.\xi=Pr_{C^{+}}(\xi)+\sum_{n\in\mathbb{N}}Pr_{D_{n}^{+}}(\xi)\;\;\;{\rm and}\;\;\;\|\xi\|_{1}=\|Pr_{C^{+}}(\xi)\|_{1}+\sum_{n\in\mathbb{N}}\|Pr_{D_{n}^{+}}(\xi)\|_{1}.

In particular, we have j=1Nαjpj1=PrC+(j=1Nαjpj)1+nPrDn+(j=1Nαjpj)1\left\|\sum_{j=1}^{N}\alpha_{j}p_{j}\right\|_{1}=\left\|Pr_{C^{+}}(\sum_{j=1}^{N}\alpha_{j}p_{j})\right\|_{1}+\sum_{n\in\mathbb{N}}\left\|Pr_{D_{n}^{+}}(\sum_{j=1}^{N}\alpha_{j}p_{j})\right\|_{1}.

We will show that, for any nn\in\mathbb{N} with 1ndY(a,b)1\leq n\leq d_{Y}(a,b), we have

|j=1Nαj|PrDn+(j=1Nαjpj)1.\Bigg{|}\sum_{j=1}^{N}\alpha_{j}\Bigg{|}\leq\Bigg{\|}Pr_{D_{n}^{+}}(\sum_{j=1}^{N}\alpha_{j}p_{j})\Bigg{\|}_{1}.

By identifying 0(Y)\ell^{0}(Y) with E+(Y)\mathbb{C}^{E^{+}(Y)} as explained in Section 2.3, we define a linear map ψ:C1(Y)\psi\colon C_{1}(Y)\to\mathbb{C} by ψ(eE+(Y)ce[e])=eE+(Y)ce\psi(\sum_{e\in E^{+}(Y)}c_{e}[e])=\sum_{e\in E^{+}(Y)}c_{e}. Note that the sums are finite and ψ\psi depends on the chosen set E+(Y)E^{+}(Y) of positive edges. For any ξ=eE+(Y)ce[e]C1(Y)\xi=\sum_{e\in E^{+}(Y)}c_{e}[e]\in C_{1}(Y), we have

|ψ(ξ)|=|eE+(Y)ce|eE+(Y)|ce|=ξ1.|\psi(\xi)|=\left|\sum_{e\in E^{+}(Y)}c_{e}\right|\leq\sum_{e\in E^{+}(Y)}|c_{e}|=\|\xi\|_{1}.

Here, for any path pjp_{j} from aa to bb and any nn\in\mathbb{N} with 1ndY(a,b)1\leq n\leq d_{Y}(a,b), we have

ψ(PrDn+(pj))=1.\psi(Pr_{D_{n}^{+}}(p_{j}))=1.

Indeed, let pj=(e1,,ek)p_{j}=(e_{1},\cdots,e_{k}) as a sequence of edges, where o(e1)=ao(e_{1})=a and t(ek)=bt(e_{k})=b, and pjDn=(ei1,,eim)p_{j}\cap D_{n}=(e_{i_{1}},\cdots,e_{i_{m}}) as its subsequence by abuse of notation, i.e. pj=l=1k[el]p_{j}=\sum_{l=1}^{k}[e_{l}] and PrDn+(pj)=l=1m[eil]Pr_{D_{n}^{+}}(p_{j})=\sum_{l=1}^{m}[e_{i_{l}}] as a 1-chain. We can see that mm is odd, i.e. m=2l1m=2l-1 with ll\in\mathbb{N} and also ei1Dn+,ei2Dn,ei3Dn+,ei4Dn,,ei2l1Dn+e_{i_{1}}\in D_{n}^{+},e_{i_{2}}\in D_{n}^{-},e_{i_{3}}\in D_{n}^{+},e_{i_{4}}\in D_{n}^{-},\cdots,e_{i_{2l-1}}\in D_{n}^{+}. This implies

ψ(PrDn+pj)\displaystyle\psi(Pr_{D_{n}^{+}}p_{j}) =ψ([ei1])+ψ([ei2])+ψ([ei3])+ψ([ei4])++ψ([ei2l1])\displaystyle=\psi([e_{i_{1}}])+\psi([e_{i_{2}}])+\psi([e_{i_{3}}])+\psi([e_{i_{4}}])+\cdots+\psi([e_{i_{2l-1}]})
=1+(1)+1+(1)++1\displaystyle=1+(-1)+1+(-1)+\cdots+1
=1.\displaystyle=1.

Hence,

PrDn+(j=1Nαjpj)1\displaystyle\Bigg{\|}Pr_{D_{n}^{+}}(\sum_{j=1}^{N}\alpha_{j}p_{j})\Bigg{\|}_{1} |ψ(PrDn+(j=1Nαjpj))|=|ψ(j=1NαjPrDn+(pj))|=|j=1Nαjψ(PrDn+(pj))|\displaystyle\geq\Bigg{|}\psi(Pr_{D_{n}^{+}}(\sum_{j=1}^{N}\alpha_{j}p_{j}))\Bigg{|}=\Bigg{|}\psi(\sum_{j=1}^{N}\alpha_{j}Pr_{D_{n}^{+}}(p_{j}))\Bigg{|}=\Bigg{|}\sum_{j=1}^{N}\alpha_{j}\psi(Pr_{D_{n}^{+}}(p_{j}))\Bigg{|}
=|j=1Nαj|.\displaystyle=\Bigg{|}\sum_{j=1}^{N}\alpha_{j}\Bigg{|}.

Thus, we finally get

j=1Nαjpj1\displaystyle\Bigg{\|}\sum_{j=1}^{N}\alpha_{j}p_{j}\Bigg{\|}_{1} =PrC+(j=1Nαjpj)1+nPrDn+(j=1Nαjpj)1\displaystyle=\Bigg{\|}Pr_{C^{+}}(\sum_{j=1}^{N}\alpha_{j}p_{j})\Bigg{\|}_{1}+\sum_{n\in\mathbb{N}}\Bigg{\|}Pr_{D_{n}^{+}}(\sum_{j=1}^{N}\alpha_{j}p_{j})\Bigg{\|}_{1}
0+n=1dY(a,b)|j=1Nαj|=|j=1Nαj|dY(a,b).\displaystyle\geq 0+\sum_{n=1}^{d_{Y}(a,b)}\Bigg{|}\sum_{j=1}^{N}\alpha_{j}\Bigg{|}=\Bigg{|}\sum_{j=1}^{N}\alpha_{j}\Bigg{|}\cdot d_{Y}(a,b).

Definition 3.4.

Suppose that YY is a graph without loops or multiple edges and E+(Y)E^{+}(Y) is a set of positive edges. For ξ=eE+(Y)ce[e]1(Y)\xi=\sum_{e\in E^{+}(Y)}c_{e}[e]\in\ell^{1}(Y) such that cec_{e}\in\mathbb{R} for any eE+(Y)e\in E^{+}(Y), we define an element ξ~=eE+(Y)be[e]2(Y)\widetilde{\xi}=\sum_{e\in E^{+}(Y)}b_{e}[e]\in\ell^{2}(Y) by

be={ceifce0|ce|ifce<0.b_{e}=\begin{cases}\sqrt{c_{e}}&\mathrm{if}\;\;c_{e}\geq 0\\ -\sqrt{|c_{e}|}&\mathrm{if}\;\;c_{e}<0.\end{cases}

Note that the above definition is independent of the choice of a set of positive edges.

Lemma 3.5.

For any ξ1,ξ21(Y)\xi_{1},\xi_{2}\in\ell^{1}(Y) whose coefficients are real, the following hold.

  • (1)

    ξ1~22=ξ11\|\widetilde{\xi_{1}}\|_{2}^{2}=\|\xi_{1}\|_{1}.

  • (2)

    ξ1~ξ2~222ξ1ξ21\|\widetilde{\xi_{1}}-\widetilde{\xi_{2}}\|_{2}^{2}\leq 2\|\xi_{1}-\xi_{2}\|_{1}.

Proof.

(1) follows trivially from Definition 3.4. We will prove (2). For any c1,c20c_{1},c_{2}\in\mathbb{R}_{\geq 0}, we have

|c1c2|2|c1c2|(c1+c2)=|c1c2|,|\sqrt{c_{1}}-\sqrt{c_{2}}|^{2}\leq|\sqrt{c_{1}}-\sqrt{c_{2}}|(\sqrt{c_{1}}+\sqrt{c_{2}})=|c_{1}-c_{2}|,
|c1+c2|2=2(c1+c2)(c1c2)22|c1+c2|.|\sqrt{c_{1}}+\sqrt{c_{2}}|^{2}=2(c_{1}+c_{2})-(\sqrt{c_{1}}-\sqrt{c_{2}})^{2}\leq 2|c_{1}+c_{2}|.

Therefore, given

ξ1=eE+(Y)c1e[e],ξ2=eE+(Y)c2e[e],ξ1~=eE+(Y)b1e[e],ξ2~=eE+(Y)b2e[e],\xi_{1}=\sum_{e\in E^{+}(Y)}c_{1e}[e],\;\xi_{2}=\sum_{e\in E^{+}(Y)}c_{2e}[e],\;\widetilde{\xi_{1}}=\sum_{e\in E^{+}(Y)}b_{1e}[e],\;\widetilde{\xi_{2}}=\sum_{e\in E^{+}(Y)}b_{2e}[e],

we have |b1eb2e|22|c1ec2e||b_{1e}-b_{2e}|^{2}\leq 2|c_{1e}-c_{2e}| for any eE+(Y)e\in E^{+}(Y). ∎

Lemma 3.6.

If GG is a group and X0X_{0} is a generating set of GG such that X0=X01X_{0}=X_{0}^{-1} and 1X01\in X_{0}, then the Cayley graph Γ(G,X02)\Gamma(G,X_{0}^{2}) is 2-vertex-connected, where

X02={ghGg,hX0}.X_{0}^{2}=\{gh\in G\mid g,h\in X_{0}\}.
Proof.

Note that X02X_{0}^{2} generates GG since X0X02X_{0}\subset X_{0}^{2}. Since GG acts transitively on V(Γ(G,X02))V(\Gamma(G,X_{0}^{2})) by graph automorphisms, it’s enough to prove that Γ(G,X02){1}\Gamma(G,X_{0}^{2})\setminus\{1\} is connected. Any two vertices g,hX0{1}g,h\in X_{0}\setminus\{1\} are connected in Γ(G,X02){1}\Gamma(G,X_{0}^{2})\setminus\{1\} by an edge having label g1hX02g^{-1}h\in X_{0}^{2}. Also, for any vertex ghX02{1}gh\in X_{0}^{2}\setminus\{1\}, if g1g\neq 1, then ghgh is connected to a vertex gg in Γ(G,X02){1}\Gamma(G,X_{0}^{2})\setminus\{1\} by an edge of label h1X0h^{-1}\in X_{0}. Hence, any two vertices adjacent to 11 are connected in Γ(G,X02){1}\Gamma(G,X_{0}^{2})\setminus\{1\} by a path of length at most 3. This implies that Γ(G,X02){1}\Gamma(G,X_{0}^{2})\setminus\{1\} is connected. ∎

In the following proof, recall that our coned-off Cayley graph is a graph without loops or multiple edges (see Definition 2.3).

Proof of Proposition 3.1.

We take a symmetric finite generating set X0X_{0} of GG with 1X01\in X_{0}, and define X=X02X=X_{0}^{2}. The Cayley graph Γ(G,X)\Gamma(G,X) is 2-vertex-connected by Lemma 3.6. Without loss of generality, we can assume that the subgroups HλH_{\lambda} are non-trivial. The coned-off Cayley graph Γ^\widehat{\Gamma} with respect to XX is 2-vertex-connected, because Γ(G,X)\Gamma(G,X) is 2-vertex-connected and {Hλ}λΛ\{H_{\lambda}\}_{\lambda\in\Lambda} doesn’t contain the trivial subgroup. We denote by YY the barycentric subdivision of Γ^\widehat{\Gamma}. GG acts on YY without inversion of edges. Since Γ^\widehat{\Gamma} is a 2-vertex-connected fine hyperbolic graph, so is YY. Here, we used the fact that Γ^\widehat{\Gamma} doesn’t have loops to ensure YY is 2-vertex-connected. Also, since the number of GG-orbits of E(Γ^)E(\widehat{\Gamma}) is finite and the edge stabilizer is trivial for any edge in E(Γ^)E(\widehat{\Gamma}), the action of GG on YY satisfies these conditions as well. Therefore, by Theorem 2.11, there exist a GG-equivariant anti-symmetric \mathbb{Q}-bicombing qq of YY and a constant T0T\geq 0 such that for any a,b,cV(Y)a,b,c\in V(Y), we have

q[a,b]+q[b,c]+q[c,a]1T.\|q[a,b]+q[b,c]+q[c,a]\|_{1}\leq T.

We define a map Q:G2(Y)Q\colon G\to\ell^{2}(Y) by

Q(g)=q[1,g]~.Q(g)=\widetilde{q[1,g]}.

For every gGg\in G, we have

πgQ(g1)=q[g,1]~=q[1,g]~=Q(g),\pi_{g}Q(g^{-1})=\widetilde{q[g,1]}=-\widetilde{q[1,g]}=-Q(g),

because qq is GG-equivariant and anti-symmetric. Given any g,hGg,h\in G, Lemma 3.5 (2) implies

Q(gh)πg(Q(h))22\displaystyle\|Q(gh)-\pi_{g}(Q(h))\|_{2}^{2} =q[1,gh]~q[g,gh]~222q[1,gh]q[g,gh]1\displaystyle=\|\widetilde{q[1,gh]}-\widetilde{q[g,gh]}\|_{2}^{2}\leq 2\|q[1,gh]-q[g,gh]\|_{1}
2(q[1,gh]+q[gh,g]+q[g,1]1+q[1,g]1)\displaystyle\leq 2(\|q[1,gh]+q[gh,g]+q[g,1]\|_{1}+\|q[1,g]\|_{1})
2(T+q[1,g]1).\displaystyle\leq 2(T+\|q[1,g]\|_{1}).

Hence, QQ is an array into (2(Y),π)(\ell^{2}(Y),\pi). By Remark 2.12, for any gGg\in G, the 1-chain q[1,g]q[1,g] is a convex combination of paths from 11 to gg i.e. there exist paths p1,,pNp_{1},\cdots,p_{N} from 11 to gg and α1,,αN0\alpha_{1},\cdots,\alpha_{N}\in\mathbb{Q}_{\geq 0} with j=1Nαj=1\sum_{j=1}^{N}\alpha_{j}=1 such that

q[1,g]=j=1Nαjpj.q[1,g]=\sum_{j=1}^{N}\alpha_{j}p_{j}.

Hence, by Lemma 3.3 and Lemma 3.5 (1), we have

dY(1,g)=|j=1Nαj|dY(1,g)j=1Nαjpj1=q[1,g]1=Q(g)22.d_{Y}(1,g)=\left|\sum_{j=1}^{N}\alpha_{j}\right|\cdot d_{Y}(1,g)\leq\left\|\sum_{j=1}^{N}\alpha_{j}p_{j}\right\|_{1}=\|q[1,g]\|_{1}=\|Q(g)\|_{2}^{2}.

Since YY is a barycentric subdivision of Γ^\widehat{\Gamma}, we have for any gGg\in G,

2dΓ^(1,g)=dY(1,g),2d_{\widehat{\Gamma}}(1,g)=d_{Y}(1,g),

hence

dΓ^(1,g)=12dY(1,g)12Q(g)22.d_{\widehat{\Gamma}}(1,g)=\frac{1}{2}d_{Y}(1,g)\leq\frac{1}{2}\|Q(g)\|_{2}^{2}.

Remark 3.7.

Note that QQ is not even proper relative to {Hλ}λΛ\{H_{\lambda}\}_{\lambda\in\Lambda}. For example, if GG is a free product of infinite groups H1H_{1} and H2H_{2}, then by Theorem 2.11 (4), we can show that there exists some N′′N^{\prime\prime}\in\mathbb{N} such that {h1h2h1H1,h2H2}{gGQ(g)2N′′}\{h_{1}h_{2}\mid h_{1}\in H_{1},\;h_{2}\in H_{2}\}\subset\{g\in G\mid\|Q(g)\|_{2}\leq N^{\prime\prime}\}.

3.3 Second array

This section is a continuation of Section 2.4. The goal of this section is to prove Proposition 3.8. We will construct an array on GG from an array on a subgroup HμH_{\mu} which is a member of a hyperbolicaly embedded collection of subgroups {Hλ}λΛ\{H_{\lambda}\}_{\lambda\in\Lambda}. The construction follows Section 4 of [8] and uses the notion of separating cosets explained in Section 2.4.

Proposition 3.8.

Suppose that GG is a group, XX is a subset of GG, and {Hλ}λΛ\{H_{\lambda}\}_{\lambda\in\Lambda} is a collection of subgroups hyperbolically embedded in (G,X)(G,X). Then, for any μΛ\mu\in\Lambda and any array rr on HμH_{\mu} into (2(Hμ),λHμ)(\ell^{2}(H_{\mu}),\lambda_{H_{\mu}}), there exists an array RR on GG into (2(G),λG)(\ell^{2}(G),\lambda_{G}) and a constant Kμ0K_{\mu}\geq 0 satisfying the following:\colon for any gGg\in G, any separating coset xHμSμ(1,g;D)xH_{\mu}\in S_{\mu}(1,g;D), and any geodesic path pp in the relative Cayley graph Γ(X)\Gamma(X\cup\mathcal{H}) from 11 to gg, we have

r(pin(xHμ)1pout(xHμ))2R(g)2+Kμ.\|r(p_{in}(xH_{\mu})^{-1}p_{out}(xH_{\mu}))\|_{2}\leq\|R(g)\|_{2}+K_{\mu}. (5)

The proof of Proposition 3.8 is essentially the same as the proof of Theorem 4.2 of [8], but because we deal with arrays instead of quasi-cocycles, we give full details with all necessary changes to make the proof self-contained.

Suppose that r:Hμ2(Hμ)r\colon H_{\mu}\to\ell^{2}(H_{\mu}) is an array on HμH_{\mu} into the left regular representation (2(Hμ),λHμ)(\ell^{2}(H_{\mu}),\lambda_{H_{\mu}}). By the embedding 2(Hμ)2(G)\ell^{2}(H_{\mu})\hookrightarrow\ell^{2}(G), we can think of rr as a map r:Hμ2(G)r\colon H_{\mu}\to\ell^{2}(G). We define a map r~:G×G2(G)\widetilde{r}\colon G\times G\to\ell^{2}(G) by

r~(f,g)={λG(f)r(f1g)iff1gHμ0iff1gHμ,\widetilde{r}(f,g)=\begin{cases}\lambda_{G}(f)r(f^{-1}g)&\mathrm{if}\;\;f^{-1}g\in H_{\mu}\\ 0&\mathrm{if}\;\;f^{-1}g\notin H_{\mu},\end{cases}

where (2(G),λG)(\ell^{2}(G),\lambda_{G}) is the left regular representation of GG.

Remark 3.9.

If f,gGf,g\in G are in the same coset of HμH_{\mu}, i.e. there exists a HμH_{\mu}-coset xHμxH_{\mu} for some xGx\in G such that f,gxHμf,g\in xH_{\mu}, then the support of r~(f,g)\widetilde{r}(f,g) is in xHμxH_{\mu}.

Lemma 3.10.

For any f,g,hGf,g,h\in G, the following hold.

  • (1)

    r~(g,f)=r~(f,g)\widetilde{r}(g,f)=-\widetilde{r}(f,g).

  • (2)

    r~(hf,hg)=λG(h)r~(f,g)\widetilde{r}(hf,hg)=\lambda_{G}(h)\widetilde{r}(f,g).

Proof.

(1) For any f,gGf,g\in G, f1gHμf^{-1}g\in H_{\mu} if and only if g1fHμg^{-1}f\in H_{\mu}. If f1gHμf^{-1}g\in H_{\mu}, we have

r~(g,f)\displaystyle\widetilde{r}(g,f) =λG(g)r(g1f)\displaystyle=\lambda_{G}(g)r(g^{-1}f)
=λG(g)(λHμ(g1f)r((g1f)1))=λG(f)r(f1g)=r~(f,g).\displaystyle=\lambda_{G}(g)\left(-\lambda_{H_{\mu}}(g^{-1}f)r((g^{-1}f)^{-1})\right)=-\lambda_{G}(f)r(f^{-1}g)=-\widetilde{r}(f,g).

(2) If f1gHμf^{-1}g\in H_{\mu}, we have

r~(hf,hg)=λG(hf)r((hf)1hg)=λG(h)λG(f)r(f1g)=λG(h)r~(f,g).\widetilde{r}(hf,hg)=\lambda_{G}(hf)r((hf)^{-1}hg)=\lambda_{G}(h)\lambda_{G}(f)r(f^{-1}g)=\lambda_{G}(h)\widetilde{r}(f,g).

Lemma 3.11.

For any gHμg\in H_{\mu}, we have

suphHμr~(g,h)r~(1,h)2=suphHμr~(1,h)r~(1,hg)2<.\sup_{h\in H_{\mu}}\|\widetilde{r}(g,h)-\widetilde{r}(1,h)\|_{2}=\sup_{h\in H_{\mu}}\|\widetilde{r}(1,h)-\widetilde{r}(1,hg)\|_{2}<\infty.
Proof.

For any gHμg\in H_{\mu}, we have

suphHμr~(g,h)r~(1,h)2\displaystyle\sup_{h\in H_{\mu}}\|\widetilde{r}(g,h)-\widetilde{r}(1,h)\|_{2} =suphHμλHμ(g)r(g1h)r(h)2\displaystyle=\sup_{h\in H_{\mu}}\|\lambda_{H_{\mu}}(g)r(g^{-1}h)-r(h)\|_{2}
=suphHμr(g1h)λHμ(g1)r(h)2<,\displaystyle=\sup_{h\in H_{\mu}}\|r(g^{-1}h)-\lambda_{H_{\mu}}(g^{-1})r(h)\|_{2}<\infty,

and by Lemma 3.10,

suphHμr~(1,h)r~(1,hg)2\displaystyle\sup_{h\in H_{\mu}}\|\widetilde{r}(1,h)-\widetilde{r}(1,hg)\|_{2} =suphHμλHμ(h1)(r~(1,h)r~(1,hg))2\displaystyle=\sup_{h\in H_{\mu}}\|\lambda_{H_{\mu}}(h^{-1})\left(\widetilde{r}(1,h)-\widetilde{r}(1,hg)\right)\|_{2}
=suphHμr~(h1,1)r~(h1,g)2\displaystyle=\sup_{h\in H_{\mu}}\|\widetilde{r}(h^{-1},1)-\widetilde{r}(h^{-1},g)\|_{2}
=suphHμr~(1,h1)+r~(g,h1)2\displaystyle=\sup_{h\in H_{\mu}}\|-\widetilde{r}(1,h^{-1})+\widetilde{r}(g,h^{-1})\|_{2}
=suphHμr~(g,h)r~(1,h)2.\displaystyle=\sup_{h\in H_{\mu}}\|\widetilde{r}(g,h)-\widetilde{r}(1,h)\|_{2}.

In the following, we denote

Kg=suphHμr~(g,h)r~(1,h)2=suphHμr~(1,h)r~(1,hg)2.K_{g}=\sup_{h\in H_{\mu}}\|\widetilde{r}(g,h)-\widetilde{r}(1,h)\|_{2}=\sup_{h\in H_{\mu}}\|\widetilde{r}(1,h)-\widetilde{r}(1,hg)\|_{2}. (6)
Remark 3.12.

For any gHμg\in H_{\mu}, we have

Kg\displaystyle K_{g} =suphHμr~(1,h)r~(1,hg)2=suphHμr~(1,hg1)r~(1,(hg1)g)2\displaystyle=\sup_{h\in H_{\mu}}\|\widetilde{r}(1,h)-\widetilde{r}(1,hg)\|_{2}=\sup_{h\in H_{\mu}}\|\widetilde{r}(1,hg^{-1})-\widetilde{r}(1,(hg^{-1})g)\|_{2}
=suphHμr~(1,h)r~(1,hg1)2=Kg1.\displaystyle=\sup_{h\in H_{\mu}}\|\widetilde{r}(1,h)-\widetilde{r}(1,hg^{-1})\|_{2}=K_{g^{-1}}.
Remark 3.13.

By Lemma 3.10 (1), r~(1,1)=0\widetilde{r}(1,1)=0, hence for any gHμg\in H_{\mu}, we have

r~(1,g)2=r~(g,1)r~(1,1)2suphHμr~(g,h)r~(1,h)2=Kg.\|\widetilde{r}(1,g)\|_{2}=\|\widetilde{r}(g,1)-\widetilde{r}(1,1)\|_{2}\leq\sup_{h\in H_{\mu}}\|\widetilde{r}(g,h)-\widetilde{r}(1,h)\|_{2}=K_{g}.
Lemma 3.14.

For any elements f1,f2,g1,g2Gf_{1},f_{2},g_{1},g_{2}\in G that are in the same coset of HμH_{\mu}, we have

r~(f1,g1)r~(f2,g2)2Kf11f2+Kg11g2.\|\widetilde{r}(f_{1},g_{1})-\widetilde{r}(f_{2},g_{2})\|_{2}\leq K_{f_{1}^{-1}f_{2}}+K_{g_{1}^{-1}g_{2}}.
Proof.

By Lemma 3.11, we have

r~(f1,\displaystyle\|\widetilde{r}(f_{1}, g1)r~(f2,g2)2r~(f1,g1)r~(f2,g1)2+r~(f2,g1)r~(f2,g2)2\displaystyle g_{1})-\widetilde{r}(f_{2},g_{2})\|_{2}\leq\|\widetilde{r}(f_{1},g_{1})-\widetilde{r}(f_{2},g_{1})\|_{2}+\|\widetilde{r}(f_{2},g_{1})-\widetilde{r}(f_{2},g_{2})\|_{2}
r~(1,f11g1)r~(f11f2,f11g1)2+r~(1,f21g1)r~(1,f21g1g11g2)2\displaystyle\leq\|\widetilde{r}(1,f_{1}^{-1}g_{1})-\widetilde{r}(f_{1}^{-1}f_{2},f_{1}^{-1}g_{1})\|_{2}+\|\widetilde{r}(1,f_{2}^{-1}g_{1})-\widetilde{r}(1,f_{2}^{-1}g_{1}\cdot g_{1}^{-1}g_{2})\|_{2}
Kf11f2+Kg11g2.\displaystyle\leq K_{f_{1}^{-1}f_{2}}+K_{g_{1}^{-1}g_{2}}.

For f,gGf,g\in G and xHμSμ(f,g;D)xH_{\mu}\in S_{\mu}(f,g;D), we define R~(f,g;xHμ)2(G)\widetilde{R}(f,g;xH_{\mu})\in\ell^{2}(G) by

R~(f,g;xHμ)=1|E(f,g;xHμ,D)|(u,v)E(f,g;xHμ,D)r~(u,v).\widetilde{R}(f,g;xH_{\mu})=\frac{1}{|E(f,g;xH_{\mu},D)|}\sum_{(u,v)\in E(f,g;xH_{\mu},D)}\widetilde{r}(u,v).
Remark 3.15.

This is well-defined because E(f,g;xHμ,D)E(f,g;xH_{\mu},D) is finite by Lemma 2.27 (c). Also, the support of R~(f,g;xHμ)\widetilde{R}(f,g;xH_{\mu}) is in xHμxH_{\mu} by Remark 3.9.

Lemma 3.16.

For any f,g,hGf,g,h\in G and xHμSμ(f,g;D)xH_{\mu}\in S_{\mu}(f,g;D), the following holds.

  • (a)

    R~(g,f;xHμ)=R~(f,g;xHμ)\widetilde{R}(g,f;xH_{\mu})=-\widetilde{R}(f,g;xH_{\mu}).

  • (b)

    R~(hf,hg;hxHμ)=λG(h)R~(f,g;xHμ)\widetilde{R}(hf,hg;hxH_{\mu})=\lambda_{G}(h)\widetilde{R}(f,g;xH_{\mu}).

Proof.

It follows from Lemma 2.22, Lemma 2.27, and Lemma 3.10. ∎

For n0n\in\mathbb{R}_{\geq}0, we define a constant by

Kn=max{KggHμd^μ(1,g)n},K_{n}=\max\{K_{g}\mid g\in H_{\mu}\wedge\widehat{d}_{\mu}(1,g)\leq n\}, (7)

where KgK_{g} is defined by (6). Because (Hμ,d^μ)(H_{\mu},\widehat{d}_{\mu}) is a locally finite metric space (cf. Definition 2.15 (2)), we have Kn<K_{n}<\infty.

Lemma 3.17.

For any f,gGf,g\in G, any xHμSμ(f,g;D)xH_{\mu}\in S_{\mu}(f,g;D), and any (u,v)E(f,g;xHμ,D)(u,v)\in E(f,g;xH_{\mu},D), we have

R~(f,g;xHμ)r~(u,v)22KD.\|\widetilde{R}(f,g;xH_{\mu})-\widetilde{r}(u,v)\|_{2}\leq 2K_{D}.
Proof.

For any (u,v),(u,v)E(f,g;xHμ,D)(u,v),(u^{\prime},v^{\prime})\in E(f,g;xH_{\mu},D), we have d^μ(u,u)3CD\widehat{d}_{\mu}(u,u^{\prime})\leq 3C\leq D and d^μ(v,v)3CD\widehat{d}_{\mu}(v,v^{\prime})\leq 3C\leq D by Lemma 2.23 (b). This implies, by Lemma 3.14,

r~(u,v)r~(u,v)2Ku1u+Kv1vKD+KD=2KD.\|\widetilde{r}(u,v)-\widetilde{r}(u^{\prime},v^{\prime})\|_{2}\leq K_{u^{-1}u^{\prime}}+K_{v^{-1}v^{\prime}}\leq K_{D}+K_{D}=2K_{D}.

Thus, for any (u,v)E(f,g;xHμ,D)(u,v)\in E(f,g;xH_{\mu},D), we have

R~(f,g;xHμ)r~(u,v)2\displaystyle\|\widetilde{R}(f,g;xH_{\mu})-\widetilde{r}(u,v)\|_{2} =1|E(f,g;xHμ,D)|(u,v)E(f,g;xHμ,D)(r~(u,v)r~(u,v))2\displaystyle=\left\|\frac{1}{|E(f,g;xH_{\mu},D)|}\sum_{(u^{\prime},v^{\prime})\in E(f,g;xH_{\mu},D)}\left(\widetilde{r}(u^{\prime},v^{\prime})-\widetilde{r}(u,v)\right)\right\|_{2}
1|E(f,g;xHμ,D)|(u,v)E(f,g;xHμ,D)r~(u,v)r~(u,v)2\displaystyle\leq\frac{1}{|E(f,g;xH_{\mu},D)|}\sum_{(u^{\prime},v^{\prime})\in E(f,g;xH_{\mu},D)}\left\|\widetilde{r}(u^{\prime},v^{\prime})-\widetilde{r}(u,v)\right\|_{2}
1|E(f,g;xHμ,D)|(u,v)E(f,g;xHμ,D)2KD=2KD.\displaystyle\leq\frac{1}{|E(f,g;xH_{\mu},D)|}\sum_{(u^{\prime},v^{\prime})\in E(f,g;xH_{\mu},D)}2K_{D}=2K_{D}.

Finally, we define a map R~:G×G2(G)\widetilde{R}\colon G\times G\to\ell^{2}(G) by

R~(f,g)=xHμSμ(f,g;D)R~(f,g;xHμ).\widetilde{R}(f,g)=\sum_{xH_{\mu}\in S_{\mu}(f,g;D)}\widetilde{R}(f,g;xH_{\mu}).

This implicitly means that if Sμ(f,g;D)S_{\mu}(f,g;D) is empty, then R~(f,g)=0\widetilde{R}(f,g)=0.

Lemma 3.18.

For any f,g,hGf,g,h\in G, the following hold.

  • (a)

    R~(g,f)=R~(f,g)\widetilde{R}(g,f)=-\widetilde{R}(f,g).

  • (b)

    R~(hf,hg)=λG(h)R~(f,g)\widetilde{R}(hf,hg)=\lambda_{G}(h)\widetilde{R}(f,g).

  • (c)

    R~(f,g)22=xHμSμ(f,g;D)R~(f,g;xHμ)22\|\widetilde{R}(f,g)\|_{2}^{2}=\sum_{xH_{\mu}\in S_{\mu}(f,g;D)}\|\widetilde{R}(f,g;xH_{\mu})\|_{2}^{2}.

Proof.

(a) and (b) follow from Lemma 2.22 and Lemma 3.16. (c) follows from the fact that the support of R~(f,g;xHμ)\widetilde{R}(f,g;xH_{\mu}) is in xHμxH_{\mu} as stated in Remark 3.15. ∎

For gGg\in G, we define

Lg=max{d^μ(u,v)(u,v)E(1,g;xHμ,D),xHμSμ(1,g;D)}.L_{g}=\max\{\widehat{d}_{\mu}(u,v)\mid(u,v)\in E(1,g;xH_{\mu},D),\;xH_{\mu}\in S_{\mu}(1,g;D)\}. (8)

Lg{0}L_{g}\in\mathbb{N}\cup\{0\} is well-defined by Corollary 2.24 and Lemma 2.27 (c).

The proof of the following lemma is similar to Lemma 4.7 of [8].

Lemma 3.19.

For any gGg\in G, we have

suphGR~(1,h)+R~(h,g)+R~(g,1)2<.\sup_{h\in G}\|\widetilde{R}(1,h)+\widetilde{R}(h,g)+\widetilde{R}(g,1)\|_{2}<\infty.
Proof.

For g,hGg,h\in G, suppose that R~(1,h)+R~(h,g)+R~(g,1)=vGαvv2(G)\widetilde{R}(1,h)+\widetilde{R}(h,g)+\widetilde{R}(g,1)=\sum_{v\in G}\alpha_{v}v\in\ell^{2}(G). We define ξxHμ=vxHμαvv\xi_{xH_{\mu}}=\sum_{v\in xH_{\mu}}\alpha_{v}v for each HμH_{\mu}-coset xHμxH_{\mu}. Note that we have

R~(1,h)+R~(h,g)+R~(g,1)=xHμG/HμξxHμ.\widetilde{R}(1,h)+\widetilde{R}(h,g)+\widetilde{R}(g,1)=\sum_{xH_{\mu}\in G/H_{\mu}}\xi_{xH_{\mu}}.

Let Sμ(1,h;D)=S1,hS1,h′′F1,hS_{\mu}(1,h;D)=S_{1,h}^{\prime}\sqcup S_{1,h}^{\prime\prime}\sqcup F_{1,h}, Sμ(h,g;D)=Sh,gSh,g′′Fh,gS_{\mu}(h,g;D)=S_{h,g}^{\prime}\sqcup S_{h,g}^{\prime\prime}\sqcup F_{h,g}, and Sμ(g,1;D)=Sg,1Sg,1′′Fg,1S_{\mu}(g,1;D)=S_{g,1}^{\prime}\sqcup S_{g,1}^{\prime\prime}\sqcup F_{g,1} be the decomposition in Lemma 2.29.

If xHμSμ(1,h;D)Sμ(h,g;D)Sμ(g,1;D)xH_{\mu}\notin S_{\mu}(1,h;D)\cup S_{\mu}(h,g;D)\cup S_{\mu}(g,1;D), then we have ξxHμ=0\xi_{xH_{\mu}}=0 by Remark 3.15.

If xHμS1,hxH_{\mu}\in S_{1,h}^{\prime}, we have

ξxHμ=R~(1,h;xHμ)+R~(g,1;xHμ)=R~(1,h;xHμ)R~(1,g;xHμ)=0\xi_{xH_{\mu}}=\widetilde{R}(1,h;xH_{\mu})+\widetilde{R}(g,1;xH_{\mu})=\widetilde{R}(1,h;xH_{\mu})-\widetilde{R}(1,g;xH_{\mu})=0

since S1,hSμ(g,1;D)Sμ(h,g;D)S_{1,h}^{\prime}\subset S_{\mu}(g,1;D)\setminus S_{\mu}(h,g;D) and E(1,h;xHμ,D)=E(1,g;xHμ,D)E(1,h;xH_{\mu},D)=E(1,g;xH_{\mu},D). We can argue similarly for S1,h′′S_{1,h}^{\prime\prime}, Sh,gS_{h,g}^{\prime}, Sh,g′′S_{h,g}^{\prime\prime}, Sg,1S_{g,1}^{\prime}, Sg,1′′S_{g,1}^{\prime\prime}. Hence, we have

R~(1,h)+R~(h,g)+R~(g,1)=xHμF1,hFh,gFg,1ξxHμ.\widetilde{R}(1,h)+\widetilde{R}(h,g)+\widetilde{R}(g,1)=\sum_{xH_{\mu}\in F_{1,h}\cup F_{h,g}\cup F_{g,1}}\xi_{xH_{\mu}}.

If xHμF1,hxH_{\mu}\in F_{1,h}, there are three cases to consider (see Figures 2, 2). We fix geodesic paths p𝒢(1,h)p\in\mathcal{G}(1,h), q𝒢(h,g)q\in\mathcal{G}(h,g), and r𝒢(g,1)r\in\mathcal{G}(g,1).

Refer to caption
11
hh
gg
aa_{-}
a+a_{+}
pp
rr
qq
bb_{-}
b+b_{+}
c+c_{+}
cc_{-}
p1p_{1}
p2p_{2}
q1q_{1}
q2q_{2}
r1r_{1}
r2r_{2}
aa
bb
cc
e3e_{3}
e2e_{2}
e1e_{1}
Refer to caption
Figure 1: Case 1, Case 2 a), Case 3 a)
Refer to caption
11
pp
rr
qq
Refer to caption
e2e_{2}
Refer to caption
cc
cc_{-}
r2r_{2}
r1r_{1}
p2p_{2}
a+a_{+}
aa
aa_{-}
p1p_{1}
gg
hh
c+c_{+}
e3e_{3}
Refer to caption
Figure 2: Case 2 b), Case 3 b)

Case 1:\colon xHμSμ(h,g;D)Sμ(g,1;D)xH_{\mu}\in S_{\mu}(h,g;D)\cap S_{\mu}(g,1;D). Let a,b,ca,b,c be the HμH_{\mu}-components of p,q,rp,q,r respectively, corresponding to xHμxH_{\mu}, i.e. we have p=p1ap2,q=q1bq2,r=r1cr2p=p_{1}ap_{2},~{}q=q_{1}bq_{2},~{}r=r_{1}cr_{2}. Let e1,e2,e3e_{1},e_{2},e_{3} be paths of length at most 1 connecting a+a_{+} to bb_{-}, b+b_{+} to cc_{-}, c+c_{+} to aa_{-} respectively, whose labels are in HμH_{\mu}. We claim that e1e_{1} is isolated in the geodesic triangle e1q11p21e_{1}q_{1}^{-1}p_{2}^{-1}. Indeed, if e1e_{1} is connected to an HμH_{\mu}-component ff of p2p_{2} where we have p2=s1fs2p_{2}=s_{1}fs_{2}, then there exists a path tt of length at most 1 connecting aa_{-} to f+f_{+} since aa_{-} and f+f_{+} are in the same HμH_{\mu}-coset. Hence, the path p1ts2p_{1}ts_{2} conneting 11 to hh is shorter than pp. This contradicts that pp is a geodesic. Similarly, e1e_{1} is not connected to any HμH_{\mu}-component of q1q_{1}. Hence, we have d^μ(a+,b)3C\widehat{d}_{\mu}(a_{+},b_{-})\leq 3C by Proposition 2.19. Similarly, we also have d^μ(b+,c)3C,d^μ(c+,a)3C\widehat{d}_{\mu}(b_{+},c_{-})\leq 3C,~{}\widehat{d}_{\mu}(c_{+},a_{-})\leq 3C, hence

d^μ(b+,a)d^μ(b+,c)+d^μ(c,c+)+d^μ(c+,a)3C+Lg+3C,\widehat{d}_{\mu}(b_{+},a_{-})\leq\widehat{d}_{\mu}(b_{+},c_{-})+\widehat{d}_{\mu}(c_{-},c_{+})+\widehat{d}_{\mu}(c_{+},a_{-})\leq 3C+L_{g}+3C,

where LgL_{g} is defined by (8). Note that the inclusion (c,c+)E(g,1;xHμ,D)(c_{-},c_{+})\in E(g,1;xH_{\mu},D) follows from the inclusions r𝒢(g,1)r\in\mathcal{G}(g,1) and xHμSμ(g,1;D)xH_{\mu}\in S_{\mu}(g,1;D). Hence, by Lemma 3.14 (see also (7)), we have

r~(a,a+)r~(b+,b)2Ka1b++Ka+1bK6C+Lg+K3C.\|\widetilde{r}(a_{-},a_{+})-\widetilde{r}(b_{+},b_{-})\|_{2}\leq K_{a_{-}^{-1}b_{+}}+K_{a_{+}^{-1}b_{-}}\leq K_{6C+L_{g}}+K_{3C}.

Combining with Lemma 3.17 and Lemma 3.18 (c) (also note (3)), we obtain

ξxHμ2\displaystyle\|\xi_{xH_{\mu}}\|_{2} =R~(1,h;xHμ)+R~(h,g;xHμ)+R~(g,1;xHμ)2\displaystyle=\|\widetilde{R}(1,h;xH_{\mu})+\widetilde{R}(h,g;xH_{\mu})+\widetilde{R}(g,1;xH_{\mu})\|_{2}
(R~(1,h;xHμ)+R~(h,g;xHμ))(r~(a,a+)+r~(b,b+))2\displaystyle\leq\|\big{(}\widetilde{R}(1,h;xH_{\mu})+\widetilde{R}(h,g;xH_{\mu})\big{)}-\left(\widetilde{r}(a_{-},a_{+})+\widetilde{r}(b_{-},b_{+})\right)\|_{2}
+r~(a,a+)+r~(b,b+)2+R~(g,1;xHμ)2\displaystyle~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}+\|\widetilde{r}(a_{-},a_{+})+\widetilde{r}(b_{-},b_{+})\|_{2}+\|\widetilde{R}(g,1;xH_{\mu})\|_{2}
2KD+2KD+K6C+Lg+K3C+R~(g,1)2\displaystyle\leq 2K_{D}+2K_{D}+K_{6C+L_{g}}+K_{3C}+\|\widetilde{R}(g,1)\|_{2}
10K10D+K6C+Lg+R~(g,1)2.\displaystyle\leq 10K_{10D}+K_{6C+L_{g}}+\|\widetilde{R}(g,1)\|_{2}.

Case 2:\colon xHμSμ(g,1;D)Sμ(h,g;D)xH_{\mu}\in S_{\mu}(g,1;D)\setminus S_{\mu}(h,g;D) or xHμSμ(h,g;D)Sμ(g,1;D)xH_{\mu}\in S_{\mu}(h,g;D)\setminus S_{\mu}(g,1;D). We can assume xHμSμ(g,1;D)Sμ(h,g;D)xH_{\mu}\in S_{\mu}(g,1;D)\setminus S_{\mu}(h,g;D) without loss of generality.

2 a) If qq penetrates xHμxH_{\mu}, then as in Case 1, let a,b,ca,b,c be the HμH_{\mu}-components of p,q,rp,q,r respectively, corresponding to xHμxH_{\mu} and e1,e2,e3e_{1},e_{2},e_{3} be paths of length at most 1 connecting a+a_{+} to bb_{-}, b+b_{+} to cc_{-}, c+c_{+} to aa_{-} respectively, whose labels are in HμH_{\mu}. In the same way as Case 1, we have d^μ(a+,b)3C,d^μ(b+,c)3C,d^μ(c+,a)3C\widehat{d}_{\mu}(a_{+},b_{-})\leq 3C,~{}\widehat{d}_{\mu}(b_{+},c_{-})\leq 3C,~{}\widehat{d}_{\mu}(c_{+},a_{-})\leq 3C. Also, by xHμSμ(h,g;D)xH_{\mu}\notin S_{\mu}(h,g;D), we have d^μ(b,b+)D\widehat{d}_{\mu}(b_{-},b_{+})\leq D, hence

d^μ(a+,c)d^μ(a+,b)+d^μ(b,b+)+d^μ(b+,c)3C+D+3C3D.\widehat{d}_{\mu}(a_{+},c_{-})\leq\widehat{d}_{\mu}(a_{+},b_{-})+\widehat{d}_{\mu}(b_{-},b_{+})+\widehat{d}_{\mu}(b_{+},c_{-})\leq 3C+D+3C\leq 3D.

2 b) If qq doesn’t penetrate xHμxH_{\mu}, let a,ca,c be the HμH_{\mu}-components of p,rp,r respectively, corresponding to xHμxH_{\mu}, i.e. p=p1ap2,r=r1cr2p=p_{1}ap_{2},~{}r=r_{1}cr_{2}, and let e2,e3e_{2},e_{3} be paths of length at most 1 connecting a+a_{+} to cc_{-}, c+c_{+} to aa_{-} respectively, whose labels are in HμH_{\mu}. Because e3e_{3} is isolated in the geodesic triangle e3p11r21e_{3}p_{1}^{-1}r_{2}^{-1}, we have d^μ(c+,a)3C\widehat{d}_{\mu}(c_{+},a_{-})\leq 3C. Also, because qq doesn’t penetrate xHμxH_{\mu}, e2e_{2} is isolated in the geodesic 4-gon e2r11q1p21e_{2}r_{1}^{-1}q^{-1}p_{2}^{-1}, hence we have d^μ(a+,c)4C3D\widehat{d}_{\mu}(a_{+},c_{-})\leq 4C\leq 3D by Proposition 2.19.

In both 2 a) and 2 b), we have d^μ(c+,a)3C\widehat{d}_{\mu}(c_{+},a_{-})\leq 3C and d^μ(a+,c)3D\widehat{d}_{\mu}(a_{+},c_{-})\leq 3D. Thus, by Lemma 3.14 and 3.17, we have

ξxHμ2\displaystyle\|\xi_{xH_{\mu}}\|_{2} =R~(1,h;xHμ)+R~(g,1;xHμ)2\displaystyle=\|\widetilde{R}(1,h;xH_{\mu})+\widetilde{R}(g,1;xH_{\mu})\|_{2}
(R~(1,h;xHμ)+R~(g,1;xHμ))(r~(a,a+)+r~(c,c+))2\displaystyle\leq\|\left(\widetilde{R}(1,h;xH_{\mu})+\widetilde{R}(g,1;xH_{\mu})\right)-\left(\widetilde{r}(a_{-},a_{+})+\widetilde{r}(c_{-},c_{+})\right)\|_{2}
+r~(a,a+)r~(c+,c)2\displaystyle~{}~{}~{}~{}+\|\widetilde{r}(a_{-},a_{+})-\widetilde{r}(c_{+},c_{-})\|_{2}
2KD+2KD+K3C+K3D10K10D.\displaystyle\leq 2K_{D}+2K_{D}+K_{3C}+K_{3D}\leq 10K_{10D}.

When xHμSμ(h,g;D)Sμ(g,1;D)xH_{\mu}\in S_{\mu}(h,g;D)\setminus S_{\mu}(g,1;D), we can show ξxHμ210K10D\|\xi_{xH_{\mu}}\|_{2}\leq 10K_{10D} in the same way.

Case 3:\colon xHμSμ(h,g;D)Sμ(g,1;D)xH_{\mu}\notin S_{\mu}(h,g;D)\cup S_{\mu}(g,1;D). Note that by xHμF1,hSμ(1,h;D)xH_{\mu}\in F_{1,h}\subset S_{\mu}(1,h;D) and Lemma 2.23 (a), at least one of qq and rr penetrates xHμxH_{\mu}.

3 a) If both qq and rr penetrate xHμxH_{\mu}, let a,b,c,e1,e2,e3a,b,c,e_{1},e_{2},e_{3} be as in Case 1. Then, we have d^μ(a+,b)3C,d^μ(b+,c)3C,d^μ(c+,a)3C\widehat{d}_{\mu}(a_{+},b_{-})\leq 3C,~{}\widehat{d}_{\mu}(b_{+},c_{-})\leq 3C,~{}\widehat{d}_{\mu}(c_{+},a_{-})\leq 3C in the same way as Case 1. Also, by xHμSμ(h,g;D)Sμ(g,1;D)xH_{\mu}\notin S_{\mu}(h,g;D)\cup S_{\mu}(g,1;D), we have d^μ(b,b+)D,d^μ(c,c+)D\widehat{d}_{\mu}(b_{-},b_{+})\leq D,~{}\widehat{d}_{\mu}(c_{-},c_{+})\leq D. Hence,

d^μ(a+,a)\displaystyle\widehat{d}_{\mu}(a_{+},a_{-}) d^μ(a+,b)+d^μ(b,b+)+d^μ(b+,c)+d^μ(c,c+)+d^μ(c+,a)\displaystyle\leq\widehat{d}_{\mu}(a_{+},b_{-})+\widehat{d}_{\mu}(b_{-},b_{+})+\widehat{d}_{\mu}(b_{+},c_{-})+\widehat{d}_{\mu}(c_{-},c_{+})+\widehat{d}_{\mu}(c_{+},a_{-})
3C+D+3C+D+3C5D.\displaystyle\leq 3C+D+3C+D+3C\leq 5D.

3 b) If only one of qq and rr penetrates xHμxH_{\mu}, assume rr penetrates xHμxH_{\mu} without loss of generality, and let a,c,e2,e3a,c,e_{2},e_{3} be as in 2 b) of Case 2. Then, we have d^μ(c+,a)3C\widehat{d}_{\mu}(c_{+},a_{-})\leq 3C and d^μ(a+,c)4C\widehat{d}_{\mu}(a_{+},c_{-})\leq 4C. Also by xHμSμ(g,1;D)xH_{\mu}\notin S_{\mu}(g,1;D), we have d^μ(c,c+)D\widehat{d}_{\mu}(c_{-},c_{+})\leq D. hence,

d^μ(a+,a)d^μ(a+,c)+d^μ(c,c+)+d^μ(c+,a)4C+D+3C5D.\displaystyle\widehat{d}_{\mu}(a_{+},a_{-})\leq\widehat{d}_{\mu}(a_{+},c_{-})+\widehat{d}_{\mu}(c_{-},c_{+})+\widehat{d}_{\mu}(c_{+},a_{-})\leq 4C+D+3C\leq 5D.

In both 3 a) and 3 b), we have d^μ(a+,a)5D\widehat{d}_{\mu}(a_{+},a_{-})\leq 5D, hence by Lemma 3.14 and Remark 3.13, we have

ξxHμ2\displaystyle\|\xi_{xH_{\mu}}\|_{2} =R~(1,h;xHμ)2R~(1,h;xHμ)r~(a,a+)2+r~(a,a+)2\displaystyle=\|\widetilde{R}(1,h;xH_{\mu})\|_{2}\leq\|\widetilde{R}(1,h;xH_{\mu})-\widetilde{r}(a_{-},a_{+})\|_{2}+\|\widetilde{r}(a_{-},a_{+})\|_{2}
2KD+K5D10K10D.\displaystyle\leq 2K_{D}+K_{5D}\leq 10K_{10D}.

Here, we used r~(a,a+)2=r~(1,a1a+)2Ka1a+K5D\|\widetilde{r}(a_{-},a_{+})\|_{2}=\|\widetilde{r}(1,a_{-}^{-1}a_{+})\|_{2}\leq K_{a_{-}^{-1}a_{+}}\leq K_{5D}. When only qq penetrates xHμxH_{\mu}, we can show ξxHμ210K10D\|\xi_{xH_{\mu}}\|_{2}\leq 10K_{10D} in the same way.

Summarizing Case 1,2,3, if xHμF1,hxH_{\mu}\in F_{1,h}, we have

ξxHμ210K10D+K6C+Lg+R~(g,1)2.\|\xi_{xH_{\mu}}\|_{2}\leq 10K_{10D}+K_{6C+L_{g}}+\|\widetilde{R}(g,1)\|_{2}.

We can argue similarly for Fh,gF_{h,g} and Fg,1F_{g,1} as well. Also, by Lemma 2.29 (c), we have |F1,h|,|Fh,g|,|Fg,1|2|F_{1,h}|,|F_{h,g}|,|F_{g,1}|\leq 2. Thus, for any g,hGg,h\in G we have

R~(1,h)+R~(h,g)+R~(g,1)2\displaystyle\|\widetilde{R}(1,h)+\widetilde{R}(h,g)+\widetilde{R}(g,1)\|_{2} xHμF1,hFh,gFg,1ξxHμ2\displaystyle\leq\sum_{xH_{\mu}\in F_{1,h}\cup F_{h,g}\cup F_{g,1}}\|\xi_{xH_{\mu}}\|_{2}
xHμF1,hFh,gFg,1(10K10D+K6C+Lg+R~(g,1)2)\displaystyle\leq\sum_{xH_{\mu}\in F_{1,h}\cup F_{h,g}\cup F_{g,1}}\left(10K_{10D}+K_{6C+L_{g}}+\|\widetilde{R}(g,1)\|_{2}\right)
6(10K10D+K6C+Lg+R~(g,1)2).\displaystyle\leq 6\left(10K_{10D}+K_{6C+L_{g}}+\|\widetilde{R}(g,1)\|_{2}\right).

Proof of Proposition 3.8.

We define a map R:G2(G)R\colon G\to\ell^{2}(G) by

R(g)=R~(1,g).R(g)=\widetilde{R}(1,g).

By Lemma 3.18 (a) and (b), we have for any gGg\in G,

λG(g)R(g1)=R~(g,1)=R~(1,g)=R(g).\lambda_{G}(g)R(g^{-1})=\widetilde{R}(g,1)=-\widetilde{R}(1,g)=-R(g).

Also, by Lemma 3.19, we have for any gGg\in G,

suphGR(gh)λG(g)R(h)2\displaystyle\sup_{h\in G}\|R(gh)-\lambda_{G}(g)R(h)\|_{2} =suphGR(gg1h)λG(g)R(g1h)2\displaystyle=\sup_{h\in G}\|R(g\cdot g^{-1}h)-\lambda_{G}(g)R(g^{-1}h)\|_{2}
=suphGR~(1,h)R~(g,h)2\displaystyle=\sup_{h\in G}\|\widetilde{R}(1,h)-\widetilde{R}(g,h)\|_{2}
suphG(R~(1,h)+R~(h,g)+R~(g,1)2+R~(1,g)2)\displaystyle\leq\sup_{h\in G}\left(\|\widetilde{R}(1,h)+\widetilde{R}(h,g)+\widetilde{R}(g,1)\|_{2}+\|\widetilde{R}(1,g)\|_{2}\right)
=R~(1,g)2+suphGR~(1,h)+R~(h,g)+R~(g,1)2\displaystyle=\|\widetilde{R}(1,g)\|_{2}+\sup_{h\in G}\|\widetilde{R}(1,h)+\widetilde{R}(h,g)+\widetilde{R}(g,1)\|_{2}
<.\displaystyle<\infty.

Thus, RR is an array. For any gGg\in G, any xHμSμ(1,g;D)xH_{\mu}\in S_{\mu}(1,g;D), and any geodesic path pp in Γ(X)\Gamma(X\cup\mathcal{H}) from 11 to gg, let aa be the component of pp corresponding to xHμxH_{\mu}. By Lemma 3.17 and Lemma 3.18 (c), we have

r(a1a+)2\displaystyle\|r(a_{-}^{-1}a_{+})\|_{2} =r~(a,a+)2\displaystyle=\|\widetilde{r}(a_{-},a_{+})\|_{2}
R~(1,g;xHμ)2+2KD\displaystyle\leq\|\widetilde{R}(1,g;xH_{\mu})\|_{2}+2K_{D}
R~(1,g)2+2KD=R(g)2+2KD,\displaystyle\leq\|\widetilde{R}(1,g)\|_{2}+2K_{D}=\|R(g)\|_{2}+2K_{D},

hence RR satisfies (5) with a constant Kμ=2KDK_{\mu}=2K_{D}. ∎

3.4 Proof of main theorem

Proposition 3.20.

Suppose that GG is a finitely generated group hyperbolic relative to a collection of subgroups {Hμ}μΛ\{H_{\mu}\}_{\mu\in\Lambda} of GG. If all subgroups HμH_{\mu} are bi-exact, then GG is also bi-exact.

Proof.

Note that Λ\Lambda is finite by definition. Because HμH_{\mu}’s are exact, GG is also exact by Corollary 3 of [15]. In the following, we will verify the condition of Proposition 2.2 (3). We take a finite generating set XX of GG, a unitary representation (2(Y),π)(\ell^{2}(Y),\pi), and an array QQ as in Proposition 3.1. Since every HμH_{\mu} is bi-exact, there exists a proper array rμr_{\mu} on HμH_{\mu} into (2(Hμ),λHμ)(\ell^{2}(H_{\mu}),\lambda_{H_{\mu}}) for each μΛ\mu\in\Lambda by Proposition 2.2 (2). By Proposition 3.8, for each rμr_{\mu}, there exist an array RμR_{\mu} on GG into (2(G),λG)(\ell^{2}(G),\lambda_{G}) and a constant Kμ0K_{\mu}\geq 0 satisfying (5). Here, we used the fact that {Hμ}μΛ\{H_{\mu}\}_{\mu\in\Lambda} is hyperbolically embedded in (G,X)(G,X) by Proposition 2.16. We define a Hilbert space 𝒦\mathcal{K} and a unitary representation ρ\rho of GG by

𝒦=2(Y)(μΛ2(G)),ρ=π(μΛλG).\mathcal{K}=\ell^{2}(Y)\oplus\Big{(}\bigoplus_{\mu\in\Lambda}\ell^{2}(G)\Big{)},~{}~{}~{}~{}~{}\rho=\pi\oplus\Big{(}\bigoplus_{\mu\in\Lambda}\lambda_{G}\Big{)}. (9)

Since (𝒦,ρ)(\mathcal{K},\rho) is a direct sum of copies of (2(G),λG)(\ell^{2}(G),\lambda_{G}) by Remark 3.2, (𝒦,ρ)(\mathcal{K},\rho) is weakly contained by (2(G),λG)(\ell^{2}(G),\lambda_{G}). Now, we define a map

P:Gg(Q(g),(Rμ(g))μΛ)2(Y)(μΛ2(G))=𝒦.P\colon G\ni g\mapsto\left(Q(g),(R_{\mu}(g))_{\mu\in\Lambda}\right)\in\ell^{2}(Y)\oplus\Big{(}\bigoplus_{\mu\in\Lambda}\ell^{2}(G)\Big{)}=\mathcal{K}. (10)

Because QQ and RμR_{\mu}’s are arrays, PP is an array on GG into (𝒦,ρ)(\mathcal{K},\rho). Hence, for any g,h,kGg,h,k\in G, we have

P(kh)ρgP(g1k)\displaystyle\|P(kh)-\rho_{g}P(g^{-1}k)\| P(kh)P(k)+P(k)ρgP(g1k)\displaystyle\leq\|P(kh)-P(k)\|+\|P(k)-\rho_{g}P(g^{-1}k)\|
=ρkhP((kh)1)+ρkP(k1)+ρg1P(k)P(g1k)\displaystyle=\|-\rho_{kh}P((kh)^{-1})+\rho_{k}P(k^{-1})\|+\|\rho_{g^{-1}}P(k)-P(g^{-1}k)\|
=P(h1k1)+ρh1P(k1)+ρg1P(k)P(g1k)\displaystyle=\|-P(h^{-1}k^{-1})+\rho_{h^{-1}}P(k^{-1})\|+\|\rho_{g^{-1}}P(k)-P(g^{-1}k)\|
C(h1)+C(g1),\displaystyle\leq C(h^{-1})+C(g^{-1}),

where we denote for each sGs\in G,

C(s)=suptGP(st)ρs(P(t))<.C(s)=\sup_{t\in G}\|P(st)-\rho_{s}(P(t))\|<\infty.

Hence, for any g,hGg,h\in G, we have

supkGP(gkh)ρgP(k)=supkGP(g(g1k)h)ρgP(g1k)C(h1)+C(g1).\sup_{k\in G}\|P(gkh)-\rho_{g}P(k)\|=\sup_{k\in G}\|P(g(g^{-1}k)h)-\rho_{g}P(g^{-1}k)\|\leq C(h^{-1})+C(g^{-1}).

Finally, we will show that PP is proper. Let NN\in\mathbb{N} and gGg\in G satisfy P(g)N\|P(g)\|\leq N. Since Q(g)22+μΛRμ(g)22=P(g)2\|Q(g)\|_{2}^{2}+\sum_{\mu\in\Lambda}\|R_{\mu}(g)\|_{2}^{2}=\|P(g)\|^{2}, we get Q(g)N\|Q(g)\|\leq N and Rμ(g)N\|R_{\mu}(g)\|\leq N for any μΛ\mu\in\Lambda. Because the identity map id:(G,dΓ^)(G,dX)id\colon(G,d_{\widehat{\Gamma}})\to(G,d_{X\cup\mathcal{H}}) is bi-Lipschitz, there exists α\alpha\in\mathbb{N} such that dX(1,x)αdΓ^(1,x)d_{X\cup\mathcal{H}}(1,x)\leq\alpha d_{\widehat{\Gamma}}(1,x) for any xGx\in G. By (4), this implies

dX(1,g)αdΓ^(1,g)12αQ(g)2212αN2.d_{X\cup\mathcal{H}}(1,g)\leq\alpha d_{\widehat{\Gamma}}(1,g)\leq\frac{1}{2}\alpha\|Q(g)\|_{2}^{2}\leq\frac{1}{2}\alpha N^{2}.

We denote αN=12αN2\alpha_{N}=\frac{1}{2}\alpha N^{2} for simplicity.

Let g=w0h1w1hnwng=w_{0}h_{1}w_{1}\cdots h_{n}w_{n} be the label of a geodesic path from 11 to gg in Γ(G,X)\Gamma(G,X\cup\mathcal{H}), where wiw_{i} is a word in the alphabet XX1X\cup X^{-1} and hiμΛ(Hμ{1})h_{i}\in\bigsqcup_{\mu\in\Lambda}(H_{\mu}\setminus\{1\}). We have

|w0|+1+|w1|++1+|wn|=dX(1,g)αN,|w_{0}|+1+|w_{1}|+\cdots+1+|w_{n}|=d_{X\cup\mathcal{H}}(1,g)\leq\alpha_{N},

where |wi||w_{i}| denotes the number of letters in wiw_{i}. In particular, we have nαNn\leq\alpha_{N} and |wi|αN|w_{i}|\leq\alpha_{N} for any wiw_{i}.

On the other hand, for each hih_{i}, there exists μΛ\mu\in\Lambda such that hiHμh_{i}\in H_{\mu}. For simplicity, we denote

xi=w0h1wi1.x_{i}=w_{0}h_{1}\cdots w_{i-1}.

If xiHμSμ(1,g;D)x_{i}H_{\mu}\notin S_{\mu}(1,g;D), then we have d^μ(1,hi)=d^μ(xi,xihi)D\widehat{d}_{\mu}(1,h_{i})=\widehat{d}_{\mu}(x_{i},x_{i}h_{i})\leq D by definition of separating cosets. If xiHμSμ(1,g;D)x_{i}H_{\mu}\in S_{\mu}(1,g;D), then by (5), we have

rμ(hi)2Rμ(g)2+KμN+Kμ.\displaystyle\|r_{\mu}(h_{i})\|_{2}\leq\|R_{\mu}(g)\|_{2}+K_{\mu}\leq N+K_{\mu}.

In either case, we have hiAμh_{i}\in A_{\mu}, where

Aμ={hHμd^μ(1,h)D}{hHμrμ(h)2N+Kμ}.A_{\mu}=\{h\in H_{\mu}\mid\widehat{d}_{\mu}(1,h)\leq D\}\cup\{h\in H_{\mu}\mid\|r_{\mu}(h)\|_{2}\leq N+K_{\mu}\}.

Note that AμA_{\mu} is finite, because d^μ\widehat{d}_{\mu} is a locally finite metric on HμH_{\mu} and rμr_{\mu} is proper. Therefore, we have

{gGP(g)N}{w0h1w1hnwn|nαN,|wi|αN,hiμΛAμ}.\{g\in G\mid\|P(g)\|\leq N\}\subset\Big{\{}w_{0}h_{1}w_{1}\cdots h_{n}w_{n}\;\Big{|}\;n\leq\alpha_{N},\;|w_{i}|\leq\alpha_{N},\;h_{i}\in\bigcup_{\mu\in\Lambda}A_{\mu}\Big{\}}.

Since XX, AμA_{\mu}’s, and Λ\Lambda are finite, the set on the right-hand side above is finite, hence PP is proper. ∎

Lemma 3.21.

Subgroups of countable bi-exact groups are also bi-exact.

Proof.

Let GG be a countable bi-exact group and HH be a subgroup of GG. HH is exact, because GG is exact and subgroups of exact groups are exact (cf. [2]). By Proposition 2.2 (2), there exists a proper array r:G2(G)r\colon G\to\ell^{2}(G) on GG into (2(G),λG)(\ell^{2}(G),\lambda_{G}). Note that the restriction of λG\lambda_{G} to HH, that is, (2(G),λG|H)(\ell^{2}(G),\lambda_{G}|_{H}) is unitarily isomorphic to (HxH\G2(H),λH)(\bigoplus_{Hx\in H\backslash G}\ell^{2}(H),\lambda_{H}), hence we have (2(G),λG|H)(2(H),λH)(\ell^{2}(G),\lambda_{G}|_{H})\prec(\ell^{2}(H),\lambda_{H}). Also, it is straightforward to show that r|H:H2(G)r|_{H}\colon H\to\ell^{2}(G) is a proper array on HH into (2(G),λG|H)(\ell^{2}(G),\lambda_{G}|_{H}). Thus, by Proposition 2.2 (3), HH is bi-exact. ∎

Proof of Theorem 1.1.

It follows from Proposition 3.20 and Lemma 3.21. ∎

References

  • [1] B. H. Bowditch, Relatively hyperbolic groups, Internat. J. Algebra Comput. 22 (2012), no. 3, 1250016, 66. MR 2922380
  • [2] N. P. Brown and N. Ozawa, CC^{*}-algebras and finite-dimensional approximations, Graduate Studies in Mathematics, vol. 88, American Mathematical Society, Providence, RI, 2008. MR 2391387
  • [3] I. Chifan and A. Ioana, Amalgamated free product rigidity for group von Neumann algebras, Adv. Math. 329 (2018), 819–850. MR 3783429
  • [4] I. Chifan, T. Sinclair, and B. Udrea, On the structural theory of II1\rm II_{1} factors of negatively curved groups, II: Actions by product groups, Adv. Math. 245 (2013), 208–236. MR 3084428
  • [5] F. Dahmani, Les groupes relativement hyperboliques et leurs bords thèse de doctorat, université louis pasteur, strasbourg I. (2003), https://www-fourier.ujf-grenoble.fr/~dahmani/Files/These.pdf.
  • [6] F. Dahmani, V. Guirardel, and D. Osin, Hyperbolically embedded subgroups and rotating families in groups acting on hyperbolic spaces, Mem. Amer. Math. Soc. 245 (2017), no. 1156, v+152. MR 3589159
  • [7] G. C. Hruska, Relative hyperbolicity and relative quasiconvexity for countable groups, Algebr. Geom. Topol. 10 (2010), no. 3, 1807–1856. MR 2684983
  • [8] M. Hull and D. Osin, Induced quasicocycles on groups with hyperbolically embedded subgroups, Algebr. Geom. Topol. 13 (2013), no. 5, 2635–2665. MR 3116299
  • [9] I. Mineyev, Straightening and bounded cohomology of hyperbolic groups, Geom. Funct. Anal. 11 (2001), no. 4, 807–839. MR 1866802
  • [10] I. Mineyev and A. Yaman, Relative hyperbolicity and bounded cohomology, https://faculty.math.illinois.edu/~mineyev/math/art/rel-hyp.pdf.
  • [11] D. V. Osin, Relatively hyperbolic groups: intrinsic geometry, algebraic properties, and algorithmic problems, Mem. Amer. Math. Soc. 179 (2006), no. 843, vi+100. MR 2182268
  • [12]  , Acylindrically hyperbolic groups, Trans. Amer. Math. Soc. 368 (2016), no. 2, 851–888. MR 3430352
  • [13] N. Ozawa, Amenable actions and exactness for discrete groups, C. R. Acad. Sci. Paris Sér. I Math. 330 (2000), no. 8, 691–695. MR 1763912
  • [14]  , Solid von Neumann algebras, Acta Math. 192 (2004), no. 1, 111–117. MR 2079600
  • [15]  , Boundary amenability of relatively hyperbolic groups, Topology Appl. 153 (2006), no. 14, 2624–2630. MR 2243738
  • [16]  , A Kurosh-type theorem for type II1\rm II_{1} factors, Int. Math. Res. Not. (2006), Art. ID 97560, 21. MR 2211141
  • [17]  , An example of a solid von Neumann algebra, Hokkaido Math. J. 38 (2009), no. 3, 557–561. MR 2548235
  • [18] N. Ozawa and S. Popa, Some prime factorization results for type II1{\rm II}_{1} factors, Invent. Math. 156 (2004), no. 2, 223–234. MR 2052608
  • [19] S. Popa and S. Vaes, Unique Cartan decomposition for II1\rm II_{1} factors arising from arbitrary actions of hyperbolic groups, J. Reine Angew. Math. 694 (2014), 215–239. MR 3259044
  • [20] H. Sako, Measure equivalence rigidity and bi-exactness of groups, J. Funct. Anal. 257 (2009), no. 10, 3167–3202. MR 2568688

Department of Mathematics, Vanderbilt University, Nashville 37240, U.S.A.

E-mail: koichi.oyakawa@vanderbilt.edu