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

Confined subgroups in groups with contracting elements

Inhyeok Choi June E Huh Center for Mathematical Challenges
Korea Institute for Advanced Study
Seoul, South Korea
[email protected]
Ilya Gekhtman Wenyuan Yang Beijing International Center for Mathematical Research
Peking University
Beijing 100871, China P.R.
[email protected]
 and  Tianyi Zheng Department of Mathematics, University of California San Diego, 9500 Gilman Dr. La Jolla, CA 92093, USA. [email protected]
(Date: May 4th 2024)
Abstract.

In this paper, we study the growth of confined subgroups through boundary actions of groups with contracting elements. We establish that the growth rate of a confined subgroup is strictly greater than half of the ambient growth rate in groups with purely exponential growth. Along the way, several results are obtained on the Hopf decomposition for boundary actions of subgroups with respect to conformal measures. In particular, we prove that confined subgroups are conservative, and examples of subgroups with nontrivial Hopf decomposition are constructed. We show a connection between Hopf decomposition and quotient growth and provide a dichotomy on quotient growth of Schierer graphs for subgroups in hyperbolic groups.

Key words and phrases:
confined subgroups, Patterson-Sullivan measures, contracting elements, Hopf decomposition, growth rate
2000 Mathematics Subject Classification:
Primary 20F65, 20F67, 37D40
(I.C.) Supported by Samsung Science & Technology Foundation (SSTF-BA1301-51), Mid-Career Researcher Program (RS-2023-00278510) through the NRF funded by the government of Korea, KIAS Individual Grant (SG091901) via the June E Huh Center for Mathematical Challenges at KIAS, and by the Fields Institute for Research in Mathematical Sciences.
(W.Y.) Partially supported by National Key R & D Program of China (SQ2020YFA070059) and National Natural Science Foundation of China (No. 12131009 and No.12326601)

1. Introduction

Let GG be a locally compact, second countable topological group. The space Sub(G)\mathrm{Sub}(G) of all closed subgroups in GG, equipped with the Chabauty topology, is a compact metrizable space on which GG acts continuously by conjugation. Measurable and topological dynamics of the action GSub(G)G\curvearrowright\mathrm{Sub}(G) have been instrumental in recent advances in the study of lattices in semi-simple Lie groups, see [1, 29]. In particular, on the measurable dynamics side, invariant random subgroups (IRS), which are invariant measures on Sub(G)\mathrm{Sub}(G), have attracted a lot of interest since the terminology was coined in [2]. Their topological counterpart, namely GG–minimal systems in Sub(G)\mathrm{Sub}(G), were introduced as uniformly recurrent subgroups (URS) in [34]. URS play an important role in the connections between reduced CC^{\ast}–algebras and topological dynamics developed in [43, 15].

The notion of confined subgroups was introduced in [37] in the study of the ideal structure of group rings of simple locally finite groups. More recently, it has been further investigated in [13, 14] for a variety of countable groups of dynamical origin. Let Γ,H\Gamma,H be two subgroups of GG, where HH is not necessarily contained in Γ\Gamma. We say that HH is confined by Γ\Gamma in GG if its Γ\Gamma–orbit (i.e. all conjugates of HH under Γ\Gamma) does not accumulate in Sub(G)\mathrm{Sub}(G) to the trivial subgroup. Equivalently, there exists a compact confining subset PGP\subseteq G such that g1HgP{1}g^{-1}Hg\cap P\setminus\{1\}\neq\emptyset for all gΓg\in\Gamma. When Γ=G\Gamma=G, we simply say that HH is a confined subgroup of GG. Every subgroup HH in a non-trivial URS of Γ\Gamma is confined by Γ\Gamma. Conversely, if HH is confined by Γ\Gamma, then the Γ\Gamma–orbit closure of HH in Sub(G)\mathrm{Sub}(G) contains a non-trivial Γ\Gamma–minimal system.

In a semi-simple Lie group, discrete confined torsion-free subgroups can be described by the geometric condition that the action on the associated symmetric space has bounded injectivity radius from above. Fraczyk-Gelander [29] showed that in a higher-rank simple Lie group, discrete confined subgroups are exactly lattices, confirming a conjecture of Margulis. This is obviously false in the rank-1 case: indeed, normal subgroups of uniform lattices are confined in both the lattice and the ambient Lie group. Instead, Gelander posed a conjecture on the growth rates of discrete confined subgroups. This conjecture was recently proved by Gekhtman-Levit [31]. The methods of [29, 31] have probabilistic ingredients, using stationary random subgroups to derive geometric results.

The main goal of the present article is to investigate confined subgroups in a general class of discrete groups Γ\Gamma, which acts on a geodesic metric space (X,d)(\mathrm{X},d) with contracting elements defined as follows.

Definition 1.1.
111There are several notions of contracting elements in the literature. Our notion here is usually called strongly contracting by other authors. As there are no other ones used in this paper, we’d keep use of this terms to be consistent with [71, 72]

An isometry gIsom(X)g\in\mathrm{Isom}(\mathrm{X}) is called contracting if the orbital map

ngnoXn\in\mathbb{Z}\longmapsto g^{n}o\in\mathrm{X}

is a quasi-isometric embedding, so that the image A:={gno:n}A:=\{g^{n}o:n\in\mathbb{Z}\} has contracting property: any ball BB disjoint with AA has CC-bounded projection to AA for a constant C0C\geq 0 independent of BB.

Contracting elements are usually thought of as hyperbolic directions, exhibiting negatively curved feature of the ambient space. The main examples we keep in mind with application include the following:

Examples 1.2.
  • Fundamental groups of rank-1 Riemannian manifolds with a finite geodesic flow invariant measure of maximal entropy (the BMS measure), where contracting elements are exactly loxodromic elements.

  • Hyperbolic groups; more generally, relative hyperbolic groups, where loxodromic elements are contracting elements.

  • CAT(0)-groups with rank-1 elements, including right-angled Coxeter/Artin groups, where contracting elements coincide with rank-1 elements.

  • Mapping class groups on Teichmüller space endowed with Teichmüller metric, where contracting elements are exactly pseudo-Anosov elements.

On a heuristic level, one might ask to what extent confined subgroups are large, or resemble normal subgroups. To be more precise, we consider two closely related aspects: the dynamics of a discrete subgroup H<Isom(X,d)H<\mathrm{Isom}(\mathrm{X},d) acting on a suitable boundary of the space X\mathrm{X}; and the growth rates of HH and its quotient X/H\mathrm{X}/H. Our investigation focuses mainly on the questions related to these aspects for a discrete confined subgroup HH.

Fix a basepoint oXo\in\mathrm{X}. We define the growth rate (or critical exponent) of the orbit HoHo as

(1) ωH=lim supnlogNH(o,n)n,\omega_{H}=\limsup_{n\to\infty}\frac{\log\sharp N_{H}(o,n)}{n},

where NH(o,n)={go:gH,d(o,go)n}N_{H}(o,n)=\{go:g\in H,d(o,go)\leq n\}. When Γ\Gamma is discrete, we also consider the growth rate ωΓ/H\omega_{\Gamma/H}, defined below in (3) analogously to (1), of the image Γo\Gamma o under the projection XX/H\mathrm{X}\to\mathrm{X}/H equipped with the quotient metric. 222In the literature, if HH is a subgroup of Γ\Gamma, ωH\omega_{H} is usually called relative growth; the relative growth of a normal subgroup HH is referred as cogrowth relative to the growth of Γ/H\Gamma/H (particularly on Γ=𝐅d\Gamma=\mathbf{F}_{d}, see [33]). Occasionally, the terminology is reversed: ωΓ/H\omega_{\Gamma/H} is called cogrowth relative to the growth ωΓ\omega_{\Gamma} of Γ\Gamma (see [53, 8]). We find that the former is more convenient in this paper.

These growth rates have been intertwined with other quantities in the study of the geometric and dynamic aspects of discrete groups for a long history. In rank-1 symmetric spaces, the famous Elstrodt-Patterson-Sullivan-Corlette formula relates growth rate to the bottom λ0\lambda_{0} of Laplace-Beltrami spectrum on X/H\mathrm{X}/H as follows:

λ0(X/H)={ωH(h(X)ωH),ωHh(X)/2h2(X)4,ωHh(X)/2\displaystyle\lambda_{0}(\mathrm{X}/H)=\begin{cases}\omega_{H}(h(\mathrm{X})-\omega_{H}),&\omega_{H}\geq h(\mathrm{X})/2\\ \frac{h^{2}(\mathrm{X})}{4},&\omega_{H}\leq h(\mathrm{X})/2\end{cases}

where h(X)h(\mathrm{X}) is the Hausdorff dimension of visual boundary. If Γ\Gamma is a lattice, then h(X)=ωΓh(\mathrm{X})=\omega_{\Gamma}.

In the discrete setting, by the cogrowth formula due to Grigorchuk, for H<𝐅dH<\mathbf{F}_{d} a subgroup in the free group on dd generators, the spectral radius ρ(X/H)\rho(\mathrm{X}/H) of simple random walks on X/H\mathrm{X}/H is given by

ρ(X/H)={2d12d(2d1eωH+eωH2d1),eωH2d12d1d,eωH2d1\displaystyle\rho(\mathrm{X}/H)=\begin{cases}\frac{\sqrt{2d-1}}{2d}(\frac{\sqrt{2d-1}}{\mathrm{e}^{\omega_{H}}}+\frac{\mathrm{e}^{\omega_{H}}}{\sqrt{2d-1}}),&\mathrm{e}^{\omega_{H}}\geq\sqrt{2d-1}\\ \frac{\sqrt{2d-1}}{d},&\mathrm{e}^{\omega_{H}}\leq\sqrt{2d-1}\end{cases}

where X\mathrm{X} is the 2d2d–regular tree (the standard Cayley graph of 𝐅d\mathbf{F}_{d}), and log(2d1)\log(2d-1) is the Hausdorff dimension of the space of ends of X\mathrm{X}.

The seminal works of Kesten [44] and Brooks [16] characterize amenability in terms of spectral radius of random walks and Laplace spectrum on Riemannian manifolds, respectively. Using the above formulae, these results can be interpreted in a common geometric setting: for Γ\Gamma of some natural classes of groups, a normal subgroup H<ΓH<\Gamma attains the maximal relative growth rate ωH=ωΓ\omega_{H}=\omega_{\Gamma} if and only if the quotient Γ/H\Gamma/H is amenable. This perspective is fruitful, with the latest results in [20, 21] for strongly positively recurrent actions (SPR) or statistically convex-cocompact actions (SCC) on hyperbolic spaces. In a different direction, Kesten’s theorem is generalized beyond normal subgroups: for IRS in [3] and URS in [28]. We refer the reader to these papers and reference therein for a more comprehensive introduction. To put into perspective, we draw in Fig. 1 the relation between various actions including SPR and SCC actions considered in this paper.

The question of whether the inequality ωΓ/H<ωΓ\omega_{\Gamma/H}<\omega_{\Gamma} holds for normal subgroups HΓH\neq\Gamma has been actively investigated [6, 62, 26] since the introduction of the notion of growth tightness in [33]. The most general results currently available for normal subgroups of divergence type actions are given by [5, 71], while the situation for general subgroups HH remains largely unexplored in the literature. Furthermore, the relations between ωH,ωΓ\omega_{H},\omega_{\Gamma} and ωΓ/H\omega_{\Gamma/H} still remain mysterious, despite recent works [39, 19].

In this paper, for a confined subgroup HH of Γ\Gamma, we establish the conservativity of the boundary action of HH, a strict lower bound ωH>ωΓ/2\omega_{H}>\omega_{\Gamma}/2 (such an inequality is sometimes referred to as cogrowth tightness), and an inequality relating ωH\omega_{H} to ωΓ\omega_{\Gamma} and ωΓ/H\omega_{\Gamma/H}. Our approach studies boundary actions equipped with conformal measures and relies on geometric arguments with contracting elements. Notably, we do not rely on any input from random walks or considerations of probability measures on the Chabauty spaces. We first present some applications before stating our general results.

1.1. Main applications

Before presenting them in full generality, we state some of our main results in several well-studied geometric settings.

First of all, let us consider a proper geodesic Gromov hyperbolic space or a CAT(0)(0) space X\mathrm{X}, equipped with the Gromov or visual boundary X\partial{\mathrm{X}} in the first and second cases respectively. Let Γ<Isom(X)\Gamma<\mathrm{Isom}(\mathrm{X}) be a non-elementary discrete subgroup containing a loxodromic or rank-1 element accordingly. If the action ΓX\Gamma\curvearrowright\mathrm{X} has purely exponential growth (or more generally, of divergence type), there exists a unique ωΓ\omega_{\Gamma}–dimensional Patterson-Sullivan measure class μPS\mu_{\mathrm{PS}} on the Gromov or visual boundary X\partial{\mathrm{X}} constructed from the action ΓX\Gamma\curvearrowright\mathrm{X}. Denote by E(Γ)E(\Gamma) the set of isometries in Isom(X)\mathrm{Isom}(\mathrm{X}) which fix pointwise the limit set Λ(Γo)\Lambda(\Gamma o) (Definition 5.5).

Theorem 1.3.

Assume the action ΓX\Gamma\curvearrowright\partial{\mathrm{X}} has purely exponential growth. Let H<Isom(X)H<\mathrm{Isom}(\mathrm{X}) be a nontrivial torsion-free discrete subgroup confined by Γ\Gamma with a compact confining subset. Then ωHωΓ/2\omega_{H}\geq\omega_{\Gamma}/2. Furthermore,

  1. (1)

    If HH preserves the measure class of μPS\mu_{\mathrm{PS}}, then the action H(X,μPS)H\curvearrowright(\partial X,\mu_{\mathrm{PS}}) is conservative.

  2. (2)

    If HH admits a finite confining subset intersecting trivially E(Γ)E(\Gamma), then ωH>ωΓ/2\omega_{H}>\omega_{\Gamma}/2 and ωH+ωΓ/H/2ωΓ\omega_{H}+\omega_{\Gamma/H}/2\geq\omega_{\Gamma}.

In many circumstances, the group of isometries fixing boundary pointwise X\partial{\mathrm{X}} is trivial (or finite), e.g. if X\mathrm{X} is a CAT(-1) space or a geodesically complete CAT(0) space with a rank-1 geodesic (i.e., bounding no half flat). In case of Λ(Γo)=X\Lambda(\Gamma o)=\partial{\mathrm{X}}, E(Γ)E(\Gamma) is finite and the assumption in the item (2) is always fulfilled.

An important subclass of (uniquely) geodesically complete CAT(0)(0) spaces are provided by Hadamard manifolds (i.e. a complete, simply connected nn–dimensional Riemannian manifold with non-positive sectional curvature). Its visual boundary X\partial\mathrm{X}, which is defined as the set of equivalence classes of geodesic rays, is homeomorphic to 𝕊n1\mathbb{S}^{n-1}. Let Γ<Isom(X)\Gamma<\mathrm{Isom}(\mathrm{X}) be a torsion-free discrete group. The quotient manifold M=X/ΓM=\mathrm{X}/\Gamma is called rank-1 if it admits a closed geodesic without a perpendicular parallel Jacobi field. In other words, MM is rank-1 if and only if Γ\Gamma contains a contracting element [12, Theorem 5.4].

Theorem 1.4.

Let M=X/ΓM=\mathrm{X}/\Gamma be a rank-1 manifold, and assume that the geodesic flow on the unit tangent bundle T1MT^{1}M has finite measure of maximal entropy. Let μPS\mu_{\mathrm{PS}} be the unique ωΓ\omega_{\Gamma}–dimensional Patterson-Sullivan measure class on X\partial\mathrm{X} constructed from the action ΓX\Gamma\curvearrowright\mathrm{X}. Let H<Isom(X)H<\mathrm{Isom}(\mathrm{X}) be a nontrivial torsion-free discrete subgroup confined by Γ\Gamma with a compact confining subset. Then ωHωΓ/2\omega_{H}\geq\omega_{\Gamma}/2. Furthermore,

  1. (1)

    If HH preserves the measure class of μPS\mu_{\mathrm{PS}}, then the action H(X,μPS)H\curvearrowright(\partial\mathrm{X},\mu_{\mathrm{PS}}) is conservative.

  2. (2)

    If HH admits a finite confining subset intersecting trivially E(Γ)E(\Gamma), then ωH>ωΓ/2\omega_{H}>\omega_{\Gamma}/2 and ωH+ωΓ/H/2ωΓ\omega_{H}+\omega_{\Gamma/H}/2\geq\omega_{\Gamma}.

As noted above, if ΓX\Gamma\curvearrowright\mathrm{X} has full limit set Λ(Γo)=X\Lambda(\Gamma o)=\partial{\mathrm{X}}, then E(Γ)E(\Gamma) is trivial.

A few remarks are in order on the statements and their background. We refer the reader to the subsequent subsections for more detailed discussions.

Remark 1.5.
  • The fundamental group Γ\Gamma is not necessarily a lattice in Isom(X)\mathrm{Isom}(\mathrm{X}). The maximal entropy measure lies in the measure class μPS×μPS×𝐋𝐞𝐛\mu_{\mathrm{PS}}\times\mu_{\mathrm{PS}}\times\mathbf{Leb} modulo Γ\Gamma–action on X×X×\partial\mathrm{X}\times\partial\mathrm{X}\times\mathbb{R} (usually called the Bowen-Margulis-Sullivan measure in the pinched negatively curved case [25] or Knieper measure in the rank-1 case [45]). The finiteness of this measure can be characterized in several ways (e.g. [25] and [61]).

    If Γ\Gamma is a uniform lattice and X\mathrm{X} is a rank-1 symmetric space of noncompact type, then μPS\mu_{\mathrm{PS}} is the Lebesgue measure on X\partial\mathrm{X} invariant under Isom(X)\mathrm{Isom}(\mathrm{X}) (see [58]). In this restricted setting, if HH is confined by Isom(X)\mathrm{Isom}(\mathrm{X}) (equivalently by any uniform lattice, see Lemma 5.3), the strict inequality ωH>ωΓ/2\omega_{H}>\omega_{\Gamma}/2 was proved by Gekhtman-Levit [31] (without the finite confining subset assumption).

  • The item (1) is known in confined Kleinian groups for the Lebesgue measure on the sphere ([52, Theorem 2.11]). However, its proof does not generalize to Gromov hyperbolic spaces. Our proof is different and works in general metric spaces; see Theorem 1.11.

  • If HH is contained in Γ\Gamma, the assumptions after “Furthermore” are redundant, and the corresponding conclusions hold without them. A confined subgroup in a non-uniform lattice Γ\Gamma is not necessarily confined in Isom(X)\mathrm{Isom}(\mathrm{X}), so the inequality in item (2) does not follow from [31].

Another application is given for confined subgroups in the mapping class group Mod(Σg)\mathrm{Mod}(\Sigma_{g}) of a closed orientable surface Σg\Sigma_{g} (g2g\geq 2). The finitely generated group Mod(Σg)\mathrm{Mod}(\Sigma_{g}) is actually the orientation preserving isometry group of the Teichmüller space 𝒯g\mathcal{T}_{g} with Teichmüller metric, on which it acts properly with growth rate (6g6)(6g-6). Investigating the similarities between Mod(Σg)\mathrm{Mod}(\Sigma_{g}) and lattices in semi-simple Lie groups has been an active research theme. The following result fits into the rank-1 phenomenon of Mod(Σg)\mathrm{Mod}(\Sigma_{g}).

Theorem 1.6.

Consider the measure class preserving action of G=Mod(Σg)G=\mathrm{Mod}(\Sigma_{g}) on the Thurston boundary 𝒫\mathscr{PMF} of the Teichmüller space 𝒯g\mathcal{T}_{g} with Thurston measure μTh\mu_{\mathrm{Th}} (cf. [7] recalled in Example 2.26).

Let H<GH<G be a nontrivial confined subgroup. If g=2g=2, assume HH is infinite. Then the following holds

  1. (1)

    H(𝒫,μTh)H\curvearrowright(\mathscr{PMF},\mu_{\mathrm{Th}}) is conservative.

  2. (2)

    ωH>ωG/2=3g3\omega_{H}>\omega_{G}/2=3g-3.

  3. (3)

    ωH+ωG/H/26g6\omega_{H}+{\omega_{G/H}}/{2}\geq 6g-6.

The items (2) and (3) for normal subgroups were proved by Arzhantseva-Cashen [4], and Coulon [19] respectively. The item (1) is new even for the normal subgroup case.

Remark 1.7.

The Mumford compactness theorem implies that GG acts cocompactly on the ϵ\epsilon–thick part of 𝒯g\mathcal{T}_{g} for any fixed ϵ>0\epsilon>0. It follows that in the setting of Theorem 1.6, HH is a confined subgroup of GG if and only if 𝒯g/H\mathcal{T}_{g}/H has bounded injectivity radius from above over any fixed thick part of 𝒯g\mathcal{T}_{g}.

1.2. Ergodic properties of boundary actions

It is classical that any action of a countable group equipped with a quasi-invariant measure admits the Hopf decomposition into the conservative and dissipative parts. In 1939, Hopf found the dichotomy in the geodesic flow of Riemannian surface with constant curvature that the flow invariant measure is either ergodic (thus conservative) or completely dissipative. Hopf’s result was extended to higher dimensional hyperbolic manifolds by Sullivan [64], forming a still-growing collection of the now-called Hopf-Tsuji-Sullivan dichotomy results in a number of settings by many authors. For the convenience of the reader, we state the HTS dichotomy for conformal measures on the Gromov boundary for a discrete group on CAT(-1) space ([61]):

Theorem 1.8 (HTS dichotomy for CAT(-1) spaces).

Assume that HXH\curvearrowright\mathrm{X} is a proper action on CAT(-1) space. Let {μx:xX}\{\mu_{x}:x\in\mathrm{X}\} be a ω\omega-dimensional HH-conformal density on the Gromov boundary X\partial{\mathrm{X}} for some ω>0\omega>0. Then we have the dichotomy: either the following equivalent statements hold:

  1. (I.1)

    𝒫H(s,x,y)=hHesd(x,hy)\mathcal{P}_{H}(s,x,y)=\sum_{h\in H}\mathrm{e}^{-sd(x,hy)} is divergent at s=ωs=\omega.

  2. (I.2)

    μx\mu_{x} is supported on the set of conical points Λcon(Ho)\Lambda^{\mathrm{con}}(Ho).

  3. (I.3)

    μx×μx\mu_{x}\times\mu_{x} is ergodic. In particular, μx\mu_{x} is ergodic.

or the following equivalent statements hold:

  1. (II.1)

    𝒫H(s,x,y)=hHesd(x,hy)\mathcal{P}_{H}(s,x,y)=\sum_{h\in H}\mathrm{e}^{-sd(x,hy)} is convergent at s=ωs=\omega.

  2. (II.2)

    μx\mu_{x} is null on the set of conical points Λcon(Ho)\Lambda^{\mathrm{con}}(Ho).

  3. (II.3)

    μx×μx\mu_{x}\times\mu_{x} is completely dissipative.

Under the set (I) of conditions, we must have ω=ωH\omega=\omega_{H} and μx\mu_{x} is unique up to scaling.

This dichotomy and its variants have then been established for conformal densities on the visual boundary of CAT(0) spaces [46]; and in our context, groups with contracting elements acting on the convergence boundary X\partial{\mathrm{X}} in a sense in [72]. See Coulon’s recent works [19, 22] for a more complete statement on horofunction boundary.

The notion of convergence boundary X\partial{\mathrm{X}} (Definition 2.13) provides a unified framework for the following boundaries equipped with a conformal density (cf. Definition 2.24 and Examples 2.15):

Examples 1.9.
  • The prototype example is the rank-11 symmetric space X\mathrm{X} of non-compact type, with a family of Isom(X)\mathrm{Isom}(\mathrm{X})–equivariant conformal measures {μx:xX}\{\mu_{x}:x\in\mathrm{X}\} in the Lebesgue measure class on the visual boundary of X\mathrm{X}. This class of conformal measures can be realized as Patterson-Sullivan measures for any uniform lattice Γ<Isom(X)\Gamma<\mathrm{Isom}(\mathrm{X}). See [58].

  • The Gromov boundary of a hyperbolic group, with quasi-conformal density constructed first by Patterson [54] from a proper action of GG on a hyperbolic space X\mathrm{X}, and then studied extensively by Sullivan [64, 66, 65] in late 1970s, with a long array of results [18, 61] by many authors, to name a few.

  • Visual boundary of a CAT(0) space admitting a proper action with rank-1 elements. The Patterson’s construction equally applies to produce a conformal density on visual boundary.

  • Thurston boundary of Teichmüller space with a conformal density constructed from Thurston measures (in the Lebesgue measure class) on the space of measured foliations in [7]. The conformal density could be realized as Patterson-Sullivan measures. See Example 2.26 for more details.

  • In recent works [19, 72], the horofunction boundary hX\partial_{h}{\mathrm{X}} has been proven to be a convergence boundary for any proper action of a group with a contracting element; furthermore, on the horofunction boundary, a good theory of conformal density can be developed: in particular, the Shadow Lemma and HTS dichotomy are available.

Suppose that the space X\mathrm{X} admits a convergence boundary X\partial{\mathrm{X}} (or keep one of these examples in mind). Motivated by rank-1 symmetric space, we assume that there is an auxiliary proper action ΓX\Gamma\curvearrowright\mathrm{X}, in order to endow a ωΓ\omega_{\Gamma}–dimensional Γ\Gamma–equivariant conformal density {μx:xX}\{\mu_{x}:x\in\mathrm{X}\} on X\partial{\mathrm{X}}. Sometimes, we assume that ΓX\Gamma\curvearrowright\mathrm{X} is statistically convex-cocompact in the sense of [71] (SCC action; see Definition 2.38); this is the case in Theorem 4.2. Among the groups in Example 1.2, the first three classes are known to admit SCC actions; and all of these examples have purely exponential growth (PEG action):

(2) n0:NΓ(o,n)enωΓ.\forall n\geq 0:\;\sharp N_{\Gamma}(o,n)\asymp\mathrm{e}^{n\omega_{\Gamma}}.\quad

Most of results actually only assume a PEG action ΓX\Gamma\curvearrowright\mathrm{X}, or even a proper action of divergence type (DIV action) in a greater generality. The class of DIV actions is naturally featured, as one of the two alternatives, in the HTS dichotomy 1.8. To be clear, we shall make precise the actions ΓX\Gamma\curvearrowright\mathrm{X} assumed in our results stated in what follows, and refer to Fig. 1 for the implication between these various actions.

Our first main object is to analyze the Hopf decomposition for a measure-class-preserving action of a group HH on (X,μo)(\partial{\mathrm{X}},\mu_{o}).

In contrast to the HTS dichotomy of geodesic flow invariant measures, which can be applicable to the HH–action on (X×X,μo×μo)(\partial{\mathrm{X}}\times\partial{\mathrm{X}},\mu_{o}\times\mu_{o}), the ergodic behavior of actions on one copy of the boundary, which is intimately linked with the horocylic flow [42], exhibits more complicated situations. For Kleinian groups, Sullivan [65, Theorem IV] characterizes the conservative component with respect to the Lebesgue measure on the boundary in terms of the (small) horospheric limit set. This leads to several important applications in Ahlfors-Bers quasi-conformal deformation theory and Mostow-type rigidity theorems ([65, Sect. V &VI]). In a recent work [35], Grigorchuk-Kaimanovich-Nagnibeda carried out a detailed analysis of the boundary action of a subgroup HH in a free group 𝐅d\mathbf{F}_{d} with respect to the uniform measure on 𝐅d\partial\mathbf{F}_{d}. These works serve as a source of inspiration in our considerations in a more general framework.

The limit set Λ(Ho)\Lambda(Ho) denotes the set of accumulation points of HoHo in a convergence boundary X\partial{\mathrm{X}}. In the last two cases in Example 1.9, Λ(Ho)\Lambda(Ho) may depend on oXo\in\mathrm{X}. To avoid this difficulty, a nontrivial partition [][\cdot] on X\partial{\mathrm{X}} was introduced so that the set [Λ(Ho)][\Lambda(Ho)] of [][\cdot]–classes on Λ(Ho)\Lambda(Ho) is independent of the choice of base points. Analogously to the notions in the actions of the convergence group, we can define conical points and the big/small horospheric limit points. More details and precise definitions are provided in Section 2.

SCC Actions [71](Definition 2.38)PEG Actions (Definition 2)Strongly Positively Recurrent Actions [63]Positively RecurrentActions [56][25, 61][56, 70]DIV Actions in HTS Dicho. 1.8
Figure 1. Relations between assumptions on actions under consideration. The equivalence in the second column is expected to hold in groups with contracting elements (some known cases as cited).

First, we present a criterion when the action of a subgroup is completely dissipative. This is a vast generalization of [35, Theorem 4.2] for free groups, which can be traced back to the case of Fuchsian groups [55] (see [49]).

Theorem 1.10.

Let a group H<Isom(X)H<\mathrm{Isom}(\mathrm{X}) act properly with contracting elements, and μ0\mu_{0} be the PS measure constructed from a proper SCC action ΓX\Gamma\curvearrowright\mathrm{X}. Assume that ωH<ωΓ/2\omega_{H}<\omega_{\Gamma}/2. Then μo(ΛHor(Ho))=0\mu_{o}(\Lambda^{\mathrm{Hor}}{(Ho)})=0, where ΛHor(Ho)\Lambda^{\mathrm{Hor}}{(Ho)} denotes the big horospheric limit set of HoHo defined in (2.46).

In addition, assume that HH is torsion-free and preserves the measure class of μo\mu_{o} on X\partial{\mathrm{X}} (in particular, if HH is a subgroup of Γ\Gamma). Then the action of HH on (X,μo)(\partial{\mathrm{X}},\mu_{o}) is completely dissipative.

Note that, if X\mathrm{X} is hyperbolic and ΓX\Gamma\curvearrowright\mathrm{X} is co-compact, then any H<Isom(X)H<\mathrm{Isom}(\mathrm{X}) preserves the measure class of μo\mu_{o} (see Lemma 4.1). We expect Theorem 1.10 to hold under the weaker assumption that ΓX\Gamma\curvearrowright\mathrm{X} has purely exponential growth (see Remark 2.45 on particular situations where this is true).

Next, we prove that the action of a confined subgroup is conservative. Denote by E(Γ)E(\Gamma) the set of isometries in Isom(X)\mathrm{Isom}(\mathrm{X}) which fix every [][\cdot]–class in the limit set [Λ(Γo)][\Lambda(\Gamma o)] (Definition 5.5).

Theorem 1.11.

Suppose that a discrete subgroup H<Isom(X)H<\mathrm{Isom}(\mathrm{X}) is confined by Γ\Gamma with a compact confining subset PP. Assume that ΓX\Gamma\curvearrowright\mathrm{X} is of divergence type, and furthermore:

  1. (i)

    Either HH is torsion-free and X\mathrm{X} is a Gromov hyperbolic or geodesically complete CAT(0) space, or

  2. (ii)

    the confining subset PP is finite, and intersects trivially E(Γ)E(\Gamma).

Then μo(ΛHor(Ho))=1\mu_{o}(\Lambda^{\mathrm{Hor}}{(Ho)})=1. In particular, H(X,μo)H\curvearrowright(\partial{\mathrm{X}},\mu_{o}) is conservative.

Note that the action of HH on (X,μo)(\partial{\mathrm{X}},\mu_{o}) is not necessarily ergodic, with examples found in free groups [35] and Kleinian groups [52]. As an immediate corollary, we obtain the following non-strict inequality. With a further inequality in Theorem 1.15, it turns out that the assumption of the SCC action can be relaxed to be the DIV action.

Corollary 1.12.

Assume that ΓX\Gamma\curvearrowright\mathrm{X} is SCC. In the setting of Theorem 1.11, we have ωHωΓ/2\omega_{H}\geq\omega_{\Gamma}/2.

1.3. Growth inequalities for confined subgroups

Via a different approach, we can upgrade the result in Corollary 1.12 to a strict inequality. We work with PEG actions ΓX\Gamma\curvearrowright\mathrm{X} strictly larger than the SCC action considered in §1.2 (particularly in Theorem 4.2).

Theorem 1.13.

Assume that the proper action ΓX\Gamma\curvearrowright\mathrm{X} has purely exponential growth with a contracting element. If H<Isom(X)H<\mathrm{Isom}(\mathrm{X}) is a discrete subgroup confined by Γ\Gamma with a finite confining subset that intersects trivially with E(Γ)E(\Gamma), then ωH>ωΓ/2\omega_{H}>\omega_{\Gamma}/2.

If H<ΓH<\Gamma is normal, Coulon [19] and independently the third named author [72] proved this inequality using boundary measures in greater generality (only assuming ΓX\Gamma\curvearrowright\mathrm{X} is of divergence type). In addition, if HH is a normal subgroup of divergence type, then ωH=ωΓ\omega_{H}=\omega_{\Gamma}. Whether this holds for general confined subgroups remains open.

Together with Corollary 1.12, Theorem 1.13 has the following application to a finite tower of confined subgroups of Γ\Gamma.

Corollary 1.14.

Assume that ΓX\Gamma\curvearrowright\mathrm{X} is a PEG action with contracting elements. Let H:=H0<H1<Hs=:ΓH:=H_{0}<H_{1}\cdots<H_{s}=:\Gamma be a sequence of subgroups so that HiH_{i} is confined in Hi+1H_{i+1} for each 0i<s0\leq i<s. Then ωH>ωΓ/2s\omega_{H}>\omega_{\Gamma}/2^{s}.

This in particular applies to an ss–subnormal subgroup HH for some integer s1s\geq 1, for which there exists a strictly increasing subnormal series, H0:=HH1Hs=:ΓH_{0}:=H\unlhd H_{1}\cdots\unlhd H_{s}=:\Gamma, of length ss. This statement was known in free groups by Olshanskii [53], who also proved that the lower bound is sharp: for any ϵ>0\epsilon>0 and s1s\geq 1, there exists some ss–subnormal subgroup for which ωH<ϵ+ωΓ/2s\omega_{H}<\epsilon+\omega_{\Gamma}/2^{s}.

Next, consider the space X/H:={Hx:xX}\mathrm{X}/H:=\{Hx:x\in\mathrm{X}\} equipped with the quotient metric. That is, given Hx,HyX/HHx,Hy\in\mathrm{X}/H, define d¯(Hx,Hy)=inf{d(x,hy):hH}\bar{d}(Hx,Hy)=\inf\{d(x,hy):h\in H\}. Denote by π:XX/H\pi:\mathrm{X}\to\mathrm{X}/H the natural projection.

Fixing a point oXo\in\mathrm{X}, consider the ball-like set in the image of Γo\Gamma o in X/H\mathrm{X}/H:

NΓ/H(o,n)={Hgo:d¯(Ho,Hgo)n}N_{\Gamma/H}(o,n)=\{Hgo:\bar{d}(Ho,Hgo)\leq n\}

and define its growth rate ωΓ/H\omega_{\Gamma/H} as follows:

(3) ωΓ/H=lim supnlogNΓ/H(o,n)n.\omega_{\Gamma/H}=\limsup_{n\to\infty}\frac{\log\sharp N_{\Gamma/H}(o,n)}{n}.

As Γo\Gamma o and Γo\Gamma o^{\prime} have a finite Hausdorff distance, the growth rate ωΓ/H\omega_{\Gamma/H} does not depend on oXo\in\mathrm{X}. If X\mathrm{X} is the Cayley graph of Γ\Gamma and HH is a subgroup of Γ\Gamma, then X/H\mathrm{X}/H is the Schreier graph associated with HH. If in addition, HH is a normal subgroup of Γ\Gamma, then ωΓ/H\omega_{\Gamma/H} is the usual growth rate of Γ/H\Gamma/H with respect to the projected generating set, which explains the notation.

We obtain the following inequality relating the growth and co-growth of confined subgroups.

Theorem 1.15.

Assume that the proper action ΓX\Gamma\curvearrowright\mathrm{X} is of divergence type with a contracting element. If a discrete group H<Isom(X)H<\mathrm{Isom}(\mathrm{X}) is confined by Γ\Gamma with a finite confining subset that intersects trivially with E(Γ)E(\Gamma), then

ωH+ωΓ/H2ωΓ\omega_{H}+\frac{\omega_{\Gamma/H}}{2}\geq\omega_{\Gamma}

The inequality for normal subgroups was obtained earlier by Jaerisch and Matsuzaki in [39] for free groups, and by Coulon [19] in a proper action of divergence type with contracting elements. Our proof uses a shadow principle for confined subgroups, from which we deduce the inequality following closely Coulon’s proof. Compared to their works, it is worth noting that the confined subgroup HH is not required to be contained in Γ\Gamma.

As a corollary, we can relax the SCC action in Corollary 1.12 to be DIV actions.

Corollary 1.16.

In the setup of Theorem 1.15, we have ωHωΓ/2\omega_{H}\geq\omega_{\Gamma}/2.

Note that it follows from Theorem 1.15 that if Γ/H\Gamma/H has sub-exponential growth (i.e. ωΓ/H=0\omega_{\Gamma/H}=0), then ωH=ωΓ\omega_{H}=\omega_{\Gamma}. By [19][Cor. 4.27], for a normal subgroup HH, we have ωH=ωΓ\omega_{H}=\omega_{\Gamma} if Γ/H\Gamma/H is amenable. We do not know what condition on Γ/H\Gamma/H might characterize the equality ωH=ωΓ\omega_{H}=\omega_{\Gamma} for more general confined subgroups.

We are now in a position to give the the proofs of Main Applications in §1.1.

Proofs of Theorems 1.3, 1.4 and 1.6.

Theorems 1.3 and 1.4 follow immediately from Theorems 1.11, 1.13 and 1.15 together. We note here that assumption in (2) are redundant: the elliptic radical E(Γ)<Isom(X)E(\Gamma)<\mathrm{Isom}(\mathrm{X}) (i.e. fixing Λ(Γo)\Lambda(\Gamma o) pointwise) intersects Γ\Gamma in a finite subgroup. If H<ΓH<\Gamma is a torsion-free subgroup, then the confining subset of HH intersects trivially E(Γ)E(\Gamma), so the assumption for (2) holds automatically.

For mapping class groups, if g>2g>2, then E(G)E(G) is trivial, so Theorem 1.6 follows exactly as above. If g=2g=2, E(G)E(G) contains hyperelliptic involution, then an additional argument is needed to conclude the proof. Indeed, since GG is residually finite, we may pass to finite index torsion-free subgroups G^\hat{G} of GG, where G^H\hat{G}\cap H is infinite and torsion-free, so Theorem 1.6 follows. The growth rate remains the same after taking finite index subgroups, so the proof in the general case follows. ∎

1.4. Maximal quotient growth

We now relate the ergodic properties of the boundary actions studied in §1.2 to the growth of a Γ\Gamma orbit Γo\Gamma o in the quotient X/H\mathrm{X}/H. A proper action ΓX\Gamma\curvearrowright\mathrm{X} is called growth tight if ωΓ/H<ωΓ\omega_{\Gamma/H}<\omega_{\Gamma} holds for any infinite normal subgroup H<ΓH<\Gamma. In the setting of Theorem 1.11, the action of a torsion-free HH on the boundary is conservative. The inequality ωΓ/H<ωΓ\omega_{\Gamma/H}<\omega_{\Gamma} in particular implies that the quotient growth of HXH\curvearrowright\mathrm{X} is slower than ΓX\Gamma\curvearrowright\mathrm{X} in the following sense:

NΓ/H(o,n)NΓ(o,n)0.\frac{\sharp N_{\Gamma/H}(o,n)}{\sharp N_{\Gamma}(o,n)}\to 0.

Following Bahturin-Olshanski [8], we say that HXH\curvearrowright\mathrm{X} has maximal quotient growth 444In terms of [8], NΓ/H(o,n)N_{\Gamma/H}(o,n) is the growth function of the right (usually non-isometric) action of Γ\Gamma on H\ΓH\backslash\Gamma. The right action is not relevant here, so we take the term of quotient growth instead. if the following holds:

NΓ/H(o,n)NΓ(o,n).\sharp N_{\Gamma/H}(o,n)\asymp\sharp N_{\Gamma}(o,n).

We shall establish an equivalence of conservative action and slower quotient growth, and a characterization of maximal right coset growth in terms of boundary actions. These are only obtained here for co-compact actions on hyperbolic spaces. We expect the characterizations to hold in a much more general setup. See Question 1.18 and Remark 10.12 for conjectures for an extension to mapping class groups.

Theorem 1.17.

Assume that Γ\Gamma acts properly and co-compactly on a proper hyperbolic space X\mathrm{X} and μo\mu_{o} is the unique Patterson-Sullivan measure class on the Gromov boundary X\partial{\mathrm{X}}. Let H<ΓH<\Gamma be a subgroup. Then the small/big horospheric limit set is μo\mu_{o}–full, if and only if X/H\mathrm{X}/H grows slower than X\mathrm{X}. Moreover, μo(Λhor(Ho))<1\mu_{o}(\Lambda^{\mathrm{hor}}{(Ho)})<1 is equivalent to the following

{Hgo:d¯(Ho,Hgo)n,gΓ}eωΓn.\sharp\{Hgo:\bar{d}(Ho,Hgo)\leq n,g\in\Gamma\}\asymp\mathrm{e}^{\omega_{\Gamma}n}.

Before giving some corollaries, let us point out the following question which plays a key role in the proof of Theorem 1.17:

Question 1.18.

Let μo\mu_{o} be a conformal measure on a boundary X\partial{\mathrm{X}} without atoms (cf. Example 2.49). Under which conditions, do we have μo(ΛHor(Ho))=μo(Λhor(Ho))\mu_{o}(\Lambda^{\mathrm{Hor}}{(Ho)})=\mu_{o}(\Lambda^{\mathrm{hor}}{(Ho)})?

This question has been raised in Sullivan’s work [65], where he showed for the real hyperbolic space 𝐇n\mathbf{H}^{n} that a discrete subgroup H<Isom(𝐇n)H<\mathrm{Isom}(\mathbf{H}^{n}) is conservative with respect to (𝐒n1,𝐋𝐞𝐛)(\mathbf{S}^{n-1},\mathbf{Leb}) if and only if the volume of 𝐇n/H\mathbf{H}^{n}/H grows slower than 𝐇n\mathbf{H}^{n}. The key part of the proof is that the big and small horospheric limit sets differ in a 𝐋𝐞𝐛\mathbf{Leb}–null set. We elaborate this proof to answer the question positively for any subgroup HH in Theorem 1.17. See Theorem 10.9 for details.

For normal subgroups, adapting an argument of [48, 27], the question can be answered positively in the following specific situation.

Theorem 1.19.

Assume that ΓX\Gamma\curvearrowright\mathrm{X} has purely exponential growth with a contracting element. Let HH be an infinite normal subgroup of Γ\Gamma. Then μo(Λhor(Ho))=μo(ΛHor(Ho))=1\mu_{o}(\Lambda^{\mathrm{hor}}{(Ho)})=\mu_{o}(\Lambda^{\mathrm{Hor}}{(Ho)})=1.

As another application of Theorem 1.17, we obtain a characterization of conservative actions.

Corollary 1.20.

Under the assumption of Theorem 1.17, assume that HH is torsion-free. Then the action H(X,μo)H\curvearrowright(\partial{\mathrm{X}},\mu_{o}) is conservative if and only if the quotient growth of HXH\curvearrowright\mathrm{X} is slower than ΓX\Gamma\curvearrowright\mathrm{X}.

Another corollary is characterizing the purely exponential growth of right cosets as in the “moreover” statement.

Let H,K<ΓH,K<\Gamma be two subgroups with proper limit sets [Λ(Ho)],[Λ(Ko)][Λ(Γo)][\Lambda(Ho)],[\Lambda(Ko)]\subsetneq[\Lambda(\Gamma o)] (we say these are of second kind following the terminology in Kleinian groups). In [36], it is shown that the double cosets HgKHgK over gΓg\in\Gamma have purely exponential growth:

{HgKo:d(o,HgKo)n,gΓ}eωΓn.\sharp\{HgKo:d(o,HgKo)\leq n,g\in\Gamma\}\asymp\mathrm{e}^{\omega_{\Gamma}n}.

If K={1}K=\{1\}, this amounts to counting the right coset HgHg as above. In hyperbolic groups, this case is thus covered in Theorem 1.17, which furthermore gives a characterization of purely exponential growth of {Hg:gΓ}\{Hg:g\in\Gamma\} by μo(Λhor(Ho))<1\mu_{o}(\Lambda^{\mathrm{hor}}{(Ho)})<1. If H={1}H=\{1\}, a characterization remains unknown to us for purely exponential growth of left cosets gKgK.

To conclude our study, we present some examples of subgroups with nontrivial Hopf decomposition. First such examples are constructed in free groups ([35]). Here we are able to construct these subgroups for any SCC actions on hyperbolic spaces.

Theorem 1.21.

Suppose that Γ\Gamma admits a proper SCC action on a proper hyperbolic space X\mathrm{X}. Let fΓf\in\Gamma be a loxodromic element. Then there exist a large integer nn so that the normal closure \llanglefn\rrangle\llangle f^{n}\rrangle contains subgroups of second kind with nontrivial conservative part and dissipative part.

We expect this result extends to any SCC action. Indeed, denoting G=\llanglefn\rrangleG=\llangle f^{n}\rrangle, if ωG<ωΓ\omega_{G}<\omega_{\Gamma} is proved, X\mathrm{X} could be replaced with a general metric space. Amenability criterion in [21] for hyperbolic spaces says that ωG=ωΓ\omega_{G}=\omega_{\Gamma} is equivalent to the amenability of Γ/G\Gamma/G. If nn is sufficiently large, Γ/G\Gamma/G is non-amenable, so ωG<ωΓ\omega_{G}<\omega_{\Gamma}. If the amenability criterion holds for any SCC action, then ωG<ωΓ\omega_{G}<\omega_{\Gamma} follows and so does Theorem 1.21.

ΓX is DIVH(X,μo) is conservative\Gamma\curvearrowright\mathrm{X}\textrm{ is }\mathrm{DIV}\Rightarrow H\curvearrowright(\partial{\mathrm{X}},\mu_{o})\textrm{ is conservative}ΓX is PEGωH>ωΓ/2\Gamma\curvearrowright\mathrm{X}\textrm{ is }\mathrm{PEG}\Rightarrow\omega_{H}>\omega_{\Gamma}/2  (Thm 1.13)ΓX is DIVωH+ωΓ/H/2ωΓ\Gamma\curvearrowright\mathrm{X}\textrm{ is }\mathrm{DIV}\Rightarrow\omega_{H}+\omega_{\Gamma/H}/2\geq\omega_{\Gamma} (Thm 1.15)Thm 1.21: ΓX\Gamma\curvearrowright\mathrm{X} is SCC on hyp. spaceThen there exist H<ΓH<\Gamma of second kind withso that μo(𝐂𝐨𝐧𝐬(H))>0&μo(𝐃𝐢𝐬𝐬(H))>0\mu_{o}(\mathbf{Cons}(H))>0\ \&\ \mu_{o}(\mathbf{Diss}(H))>0Thm 1.11: ΓX is DIV\Gamma\curvearrowright\mathrm{X}\textrm{ is }\mathrm{DIV} + (i) or (ii)H(X,μo)\Longrightarrow\ H\curvearrowright(\partial{\mathrm{X}},\mu_{o}) is conservativeThm 1.10: ΓX\Gamma\curvearrowright\mathrm{X} is SCC and ωH<ωΓ/2\omega_{H}<\omega_{\Gamma}/2H(X,μo) is completely dissipative\Longrightarrow H\curvearrowright(\partial{\mathrm{X}},\mu_{o})\textrm{ is completely dissipative}Thm 1.17: ΓX\Gamma\curvearrowright\mathrm{X} cocompact on hyp. spaceμo(𝐂𝐨𝐧𝐬(H))=1\mu_{o}(\mathbf{Cons}(H))=1\Leftrightarrow Γ/H\Gamma/H has slower growthμo(𝐂𝐨𝐧𝐬(H))<1\mu_{o}(\mathbf{Cons}(H))<1\Leftrightarrow Γ/H\Gamma/H has max. growthH<Isom(X)H<\mathrm{Isom}(\mathrm{X}) confined by ΓX\Gamma\curvearrowright\mathrm{X}with a compact confining set PPPP is finiteHopf decomp of H(X,μo)H\curvearrowright(\partial{\mathrm{X}},\mu_{o}) w.r.t. conformal measure μo\mu_{o} from ΓX\Gamma\curvearrowright\mathrm{X}PP is compact
Figure 2. A flow chart of main results, where the assumptions DIV, PEG and SCC on the action ΓX\Gamma\curvearrowright\mathrm{X} are illustrated in Fig. 1

1.5. Ingredients in the proofs and organization of the paper

We highlight some tools and ingredients in obtaining the results presented above. The reader may restrict their attention to the special case H<ΓH<\Gamma for simplicity. Our standing assumption is that the space X\mathrm{X} admits a convergence boundary X\partial{\mathrm{X}}, and Γ\Gamma acts properly on X\mathrm{X}. Readers who are not familiar with the background on convergence boundary might keep in mind the example where X\mathrm{X} is hyperbolic space with Gromov boundary X\partial{\mathrm{X}}: for this special case we have illustrations of the main ideas throughout the text.

We refer to Fig. 2 for an overview of the main theorems and their logical relations. Fig. 1 illustrates the implications between assumptions on the actions.

Results on Hopf decomposition

The proof of results on the Hopf decomposition of H(X,μo)H\curvearrowright(\partial{\mathrm{X}},\mu_{o}), as presented above in §1.2, makes extensive use of the conformal measure theory {μx:xX}\{\mu_{x}:x\in\mathrm{X}\} developed in [72] on the convergence boundary, provided that HXH\curvearrowright\mathrm{X} is of divergence type. The case of horofunction boundary was independently obtained by Coulon [19]. The key facts we use include the following:

  1. (1)

    the Shadow Lemma 2.31 holds for μx\mu_{x}, and

  2. (2)

    μx\mu_{x} is supported on conical points (Lemma 2.35), and is unique up to bounded multiplicative constants on the reduced horofunction boundary (Lemma 2.37).

A well-prepared reader would notice that these are standard consequences of the HTS dichotomy (e.g. for CAT(-1) spaces stated in Theorem 1.8). In greater generality, this has been established for groups with contracting elements as mentioned above in [19, 72]. We recommend the reader to keep in mind the situation of CAT(-1) spaces in the sequel without losing the essentials.

The proof of Theorem 1.10 in the case that X\mathrm{X} is a Gromov hyperbolic space is obtained by recasting the combinatorial proof in free groups [35] in appropriate geometric terms. The general case mimics the same outline, with the additional input of the recent result [57] on full measures supported on regularly contracting limit sets. If HH is a confined subgroup, we push a generic set of limit points into the horospheric limit set of HH to prove the conservativity of the action of HH, Theorem 1.11. This strategy has appeared in the works [48, 27] for normal subgroups of groups acting on hyperbolic spaces.

Key technical results on confined subgroups

Toward the proof of Theorem 1.11, we prove the following key technical result stated in Lemma 5.8, which can be viewed as an enhanced version of the Extension Lemma 2.10.

Lemma 1.22 (=Lemma 5.8).

There exists a finite subset FF of contracting elements in Γ\Gamma with the following property: for any gΓg\in\Gamma, there exist fFf\in F and pPp\in P such that gfpf1g1gfpf^{-1}g^{-1} lies in HH and

|d(o,gfpf1g1o)2d(o,go)|D|d(o,gfpf^{-1}g^{-1}o)-2d(o,go)|\leq D

where DD depends only on FF and PP.

To facilitate the reader with perspective from hyperbolic spaces, we include a short section §3 to demonstrate the main ideas and tools without invoking much preliminary material. Complete proofs of Theorem 1.10 and Theorem 1.11 are provided for this special case.

This lemma provides the geometric property of confined subgroups that allows us to prove statements in a manner similar to that of normal subgroups. In particular, adapting the argument in [72] for normal subgroups, we prove the following.

Lemma 1.23 (Shadow Principle in §7).

Let {μx}xX\{\mu_{x}\}_{x\in\mathrm{X}} be a ωH\omega_{H}–dimensional HH–quasi-equivariant quasi-conformal density supported on X\partial{\mathrm{X}}. Then there exists r0>0r_{0}>0 such that

μyeωHd(x,y)λμx(Πx(y,r))λ,ϵ,rμyeωHd(x,y)\begin{array}[]{rl}\|\mu_{y}\|e^{-\omega_{H}\cdot d(x,y)}\quad\prec_{\lambda}\quad\mu_{x}(\Pi_{x}(y,r))\quad\prec_{\lambda,\epsilon,r}\quad\|\mu_{y}\|e^{-\omega_{H}\cdot d(x,y)}\\ \end{array}

for any x,yΓox,y\in\Gamma o and rr0r\geq r_{0}.

The usual Shadow Lemma asserts only the above inequalities on a smaller region of x,yHoΓox,y\in Ho\subseteq\Gamma o, where μy=1\|\mu_{y}\|=1. Hence, the Shadow Principle provides useful information in the case that Γo\Gamma o is a much larger set. If ΓX\Gamma\curvearrowright\mathrm{X} is cocompact, the above result holds over the whole space X\mathrm{X}.

Cogrowth tightness of confined subgroups

The nonstrict part of inequality ωH>ωΓ/2\omega_{H}>\omega_{\Gamma}/2 asserted by Theorem 1.13 follows immediately from Theorems 1.10 and 1.11. The strict part turns out to require a more subtle argument. The strategy follows the outline of the proof of strict inequality for normal subgroups of divergence type in [72]8). The proof for the convergence type is easier, so we omit the discussion here.

The main point in [72] is to obtain a uniform bound MM on the mass of μy\mu_{y} over yΓoy\in\Gamma o:

(4) goΓo,μgoM\forall go\in\Gamma o,\quad\|\mu_{go}\|\leq M

Once this is proved, the coefficient μy\|\mu_{y}\| in the Shadow Principle disappears, so we would obtain ωHωΓ\omega_{H}\geq\omega_{\Gamma} from a standard covering argument. That is, the number of shadows Πo(go,r))\Pi_{o}(go,r)), goAΓ(o,n,Δ)go\in A_{\Gamma}(o,n,\Delta) which contain a given point of X\partial X is bounded by a uniform constant CC, and therefore

eωHnAΓ(o,n,Δ)Ce^{-\omega_{H}n}\sharp A_{\Gamma}(o,n,\Delta)\leq C

which yields ωHωΓ\omega_{H}\geq\omega_{\Gamma} and thus ωH=ωΓ\omega_{H}=\omega_{\Gamma} follows for normal subgroups HH of divergence type.

To make life easier, let us assume X\mathrm{X} is a CAT(-1) space, so the set (I) of conditions in Theorem 1.8 holds. As HXH\curvearrowright\mathrm{X} is of divergence type, there exists a unique PS measure class {μx:xX}\{\mu_{x}:x\in\mathrm{X}\} of dimension ωΓ\omega_{\Gamma}, up to scaling. In particular, the PS measure μgo\mu_{go} is the unique limit point of the following one

μgos,y=1𝒫H(s,o,y)hHesd(go,hy)Dirac(hy)\mu_{go}^{s,y}=\frac{1}{\mathcal{P}_{H}(s,o,y)}\sum_{h\in H}\mathrm{e}^{-sd(go,hy)}{\mbox{Dirac}}{(hy)}

for any yXy\in\mathrm{X}, where 𝒫H(s,x,y)=hHesd(x,hy)\mathcal{P}_{H}(s,x,y)=\sum_{h\in H}\mathrm{e}^{-sd(x,hy)} is the Poincaré series associated to HH. Note that

(5) s>ωH:μgos,go=𝒫H(s,go,go)𝒫H(s,o,go)=[𝒫H(s,go,o)𝒫Hg(s,o,o)]1\forall s>\omega_{H}:\quad\|\mu_{go}^{s,go}\|=\frac{\mathcal{P}_{H}(s,go,go)}{\mathcal{P}_{H}(s,o,go)}=\Bigg{[}\frac{\mathcal{P}_{H}(s,go,o)}{\mathcal{P}_{H^{g}}(s,o,o)}\Bigg{]}^{-1}

The proof would be finished at this point, if Hg=HH^{g}=H is a normal subgroup: the RHS is the inverse of the mass μgos,o\|\mu_{go}^{s,o}\|. The uniqueness of the limit μgo\mu_{go} concludes that μgo=1\|\mu_{go}\|=1, so the proof for normal subgroups is finished.

Our effort indeed comes into this stage to handle a general confined subgroup. We have to take a different routine to show (4), through an argument by contradiction. Assuming that 2ωH=ωΓ2\omega_{H}=\omega_{\Gamma}, we make a crucial use of Lemma 1.22 to obtain the following estimates on growth function of each conjugate gHg1gHg^{-1}

CenωH(gHg1AΓ(o,n,Δ))CenωHC\mathrm{e}^{n\omega_{H}}\leq\sharp(gHg^{-1}\cap A_{\Gamma}(o,n,\Delta))\leq C^{\prime}\mathrm{e}^{n\omega_{H}}

in Lemma 8.5, which yields the coarse equality of the associated Poincaré series:

𝒫H(ωH,o,o)𝒫Hg(ωH,o,o)\mathcal{P}_{H}(\omega_{H},o,o)\asymp\mathcal{P}_{H^{g}}(\omega_{H},o,o)

The upper bound in the above inequality uses some ingredients in proving purely exponential growth in [71]. Substituting this equation into (5) proves the boundedness of μgo\|\mu_{go}\| as above, resulting in the equality ωH=ωΓ\omega_{H}=\omega_{\Gamma}. This contradicts our assumption, concluding the proof of the strict inequality.

The argument by contradiction leaves open whether equality ωH=ωΓ\omega_{H}=\omega_{\Gamma} holds for confined subgroups of divergence type, while it is known to hold for normal subgroups.

Subgroups with nontrivial Hopf decomposition

The construction of subgroups satisfying Theorem 1.21 starts in Section 11. We first take the normal closure G=\llanglefn\rrangleG=\llangle f^{n}\rrangle of a contracting element ff for some sufficiently large nn. It is well known that such a GG is a free group of infinite rank ([24]). Removing one generator gives a subgroup HH of second kind, for which we show that it has the desired properties. The proof relies on the recent adaption of rotating family theory to projection complex ([17, 11]). The proof of nontrivial conservative component uses the inequality ωG<ωΓ\omega_{G}<\omega_{\Gamma} by the Amenability Theorem [21] for SCC action on hyperbolic spaces. We mention that the equivalence of ωG<ωΓ\omega_{G}<\omega_{\Gamma} with non-amenability of Γ/G\Gamma/G is conjectured to hold for any SCC action with contracting elements.

A guide to the sections of the paper

In the preliminary §2, we introduce necessary materials on contracting elements, convergence boundary, quasiconformal density on it, and a brief discussion on Hopf decomposition for conformal measures. The main results of the paper are then grouped into three different but closely related parts. The first part is devoted to the study of ergodic properties on boundary, establishing completely dissipative actions for subgroups with small growth (Theorem 1.10) in §4, and conservative actions for confined subgroups (Theorem 1.11) in §6. To demonstrate the main idea in our general case, we include a short section 3 to explain their proof in hyperbolic setup. The growth of confined subgroups forms the main content of the second part. Using a key lemma 5.8 obtained in the first part, we prove the shadow principle for confined subgroups in §7 and then complete the proof of strict inequality in Theorem 1.13 in §8. As a further application of the Shadow principle, Section 9 shows an inequality in Theorem 1.15 relating the growth and co-growth of confined subgroups. The last part first explains a close relation between quotient growth and Hopf decomposition and then shows the existence of nontrivial Hopf decomposition. In §10), the conservative action is characterized by slower quotient growth in Theorem 1.17. The last section 11 constructs in abundance subgroups of second kind with non-trivial Hopf decomposition (Theorem 1.21).

2. Preliminaries

Let (X,d)(\mathrm{X},d) be a proper geodesic metric space. Let Isom(X,d)\mathrm{Isom}(\mathrm{X},d) be the isometry group endowed with compact open topology. It is well known that a subgroup Γ<Isom(X,d)\Gamma<\mathrm{Isom}(\mathrm{X},d) is discrete if and only if Γ\Gamma acts properly on X\mathrm{X} ([59, Theorem 5.3.5]).

Let α:[s,t]X\alpha:[s,t]\subseteq\mathbb{R}\to\mathrm{X} be a path parametrized by arc-length, from the initial point α:=α(s)\alpha^{-}:=\alpha(s) to the terminal point α+:=α(t)\alpha^{+}:=\alpha(t). If [s,t]=[s,t]=\mathbb{R}, the restriction of α\alpha to [a,+)[a,+\infty) for aa\in\mathbb{R} is referred to as a positive ray, and its complement a negative ray. By abuse of language, we often denote them by α+\alpha^{+} and α\alpha^{-} (in particular, when they represent boundary points to which the half rays converge as in Definition 2.13).

Given two parametrized points x,yαx,y\in\alpha, [x,y]α[x,y]_{\alpha} denotes the parametrized subpath of α\alpha going from xx to yy, while [x,y][x,y] is a choice of a geodesic between x,yXx,y\in\mathrm{X}.

A path α\alpha is called a cc–quasi-geodesic for c1c\geq 1 if for any rectifiable subpath β\beta,

(β)cd(β,β+)+c\ell(\beta)\leq c\cdot d(\beta^{-},\beta^{+})+c

where (β)\ell(\beta) denotes the length of β\beta.

Denote by αβ\alpha\cdot\beta (or simply αβ\alpha\beta) the concatenation of two paths α,β\alpha,\beta provided that α+=β\alpha^{+}=\beta^{-}.

Let f,gf,g be real-valued functions. Then fcigf\prec_{c_{i}}g means that there is a constant C>0C>0 depending on parameters cic_{i} such that f<Cgf<Cg. The symbol ci\succ_{c_{i}} is defined similarly, and ci\asymp_{c_{i}} means both ci\prec_{c_{i}} and ci\succ_{c_{i}} are true. The constant cic_{i} will be omitted if it is a universal constant.

2.1. Contracting geodesics

Let ZZ be a closed subset of X\mathrm{X} and xx be a point in X\mathrm{X}. By d(x,Z)d(x,Z) we mean the set-distance between xx and ZZ, i.e.

d(x,Z):=inf{d(x,y):yZ}.d(x,Z):=\inf\big{\{}d(x,y):y\in Z\big{\}}.

Let

πZ(x):={yZ:d(x,y)=d(x,Z)}\pi_{Z}(x):=\big{\{}y\in Z:d(x,y)=d(x,Z)\big{\}}

be the set of closet point projections from xx to ZZ. Since XX is a proper metric space, πZ(x)\pi_{Z}(x) is non empty. We refer to πZ(x)\pi_{Z}(x) as the projection set of xx to ZZ. Define dZ(x,y):=πZ(x)πZ(y)\textbf{d}_{Z}(x,y):=\|\pi_{Z}(x)\cup\pi_{Z}(y)\|, where \|\cdot\| denotes the diameter.

Definition 2.1.

We say a closed subset ZXZ\subseteq X is CC–contracting for a constant C>0C>0 if, for all pairs of points x,yXx,y\in X, we have

d(x,y)d(x,Z)dZ(x,y)C.d(x,y)\leq d(x,Z)\quad\Longrightarrow\quad\textbf{d}_{Z}(x,y)\leq C.

Any such CC is called a contracting constant for ZZ. A collection of CC–contracting subsets shall be referred to as a CC–contracting system.

An element hIsom(X)h\in\mathrm{Isom}(\mathrm{X}) is called contracting if it acts co-compactly on a contracting bi-infinite quasi-geodesic. Equivalently, the map nhnon\in\mathbb{Z}\longmapsto h^{n}o is a quasi-geodesic with a contracting image.

Unless explicitly stated, let us assume from now on that Γ<Isom(X)\Gamma<\mathrm{Isom}(\mathrm{X}) is a discrete group, so ΓX\Gamma\curvearrowright\mathrm{X} is a proper action (i.e. with discrete orbits and finite point stabilizers).

A group is called elementary if it is virtually \mathbb{Z} or a finite group. In a discrete group, a contracting element must be of infinite order and is contained in a maximal elementary subgroup as described in the next lemma.

Lemma 2.2.

[71, Lemma 2.11] For a contracting element hΓh\in\Gamma, we have

EΓ(h)={gΓ:n>0,(ghng1=hn)(ghng1=hn)}.E_{\Gamma}(h)=\{g\in\Gamma:\exists n\in\mathbb{N}_{>0},(\;gh^{n}g^{-1}=h^{n})\;\lor\;(gh^{n}g^{-1}=h^{-n})\}.

We shall suppress Γ\Gamma and write E(h)=EΓ(h)E(h)=E_{\Gamma}(h) if Γ\Gamma is clear in context.

Keeping in mind the basepoint oXo\in\mathrm{X}, the axis of hh is defined as the following quasi-geodesic

(6) Ax(h)={fo:fE(h)}.\mathrm{Ax}(h)=\{fo:f\in E(h)\}.

Notice that Ax(h)=Ax(k)\mathrm{Ax}(h)=\mathrm{Ax}(k) and E(h)=E(k)E(h)=E(k) for any contracting element kE(h)k\in E(h).

An element gΓg\in\Gamma preserves the orientation of a bi-infinite quasi-geodesic γ\gamma if α\alpha and gαg\alpha has finite Hausdorff distance for any half ray α\alpha of γ\gamma. Let E+(h)E^{+}(h) be the subgroup of E(h)E(h) with possibly index 2 which elements preserve the orientation of their axis. Then we have

E+(h)={gΓ:0n,ghng1=hn}.E^{+}(h)=\{g\in\Gamma:\exists 0\neq n\in\mathbb{Z},\;gh^{n}g^{-1}=h^{n}\}.

and E+(h)E^{+}(h) contains all contracting elements in E(h)E(h), and E(h)E+(h)E(h)\setminus E^{+}(h) consists of torsion elements.

Definition 2.3.

Two contracting elements h1,h2h_{1},h_{2} in a discrete group Γ\Gamma are called independent if the collection ={gAx(hi):gΓ;i=1,2}\mathscr{F}=\{g\mathrm{Ax}(h_{i}):g\in\Gamma;\ i=1,2\} is a contracting system with bounded intersection: for any r>0r>0, there exists L=L(r)L=L(r) so that

XY,Nr(X)Nr(Y)L.\forall X\neq Y\in\mathscr{F},\;\|N_{r}(X)\cap N_{r}(Y)\|\leq L.

This is equivalent to the bounded projection: πX(Y)B\|\pi_{X}(Y)\|\leq B for some BB independent of XYX\neq Y.

In a possibly nondiscrete group Γ\Gamma, we say that h1,h2Γh_{1},h_{2}\in\Gamma are weakly independent if {h1no:n}\{h_{1}^{n}o:n\in\mathbb{Z}\} and {h2no:n}\{h_{2}^{n}o:n\in\mathbb{Z}\} have infinite Hausdorff distance. Note that in some papers, weak independence is referred to as independence.

Remark 2.4.

Note that two conjugate contracting elements with disjoint fixed points are weakly independent, but not independent. In the current paper, we mainly work with independent contracting elements, though many technical results hold for weakly independent contracting ones.

Definition 2.5.

Fix r>0r>0 and a set FF in Γ\Gamma. A geodesic γ\gamma contains an (r,f)(r,f)–barrier for fFf\in F if there exists an element gΓg\in\Gamma so that

(7) max{d(go,γ),d(gfo,γ)}r.\max\{d(g\cdot o,\gamma),\;d(g\cdot fo,\gamma)\}\leq r.

By abuse of language, the point hoho or the axis hAx(f)h\mathrm{Ax}(f) is called (r,F)(r,F)–barrier on γ\gamma.

2.2. Extension Lemma

We fix a finite set FΓF\subseteq\Gamma of independent contracting elements and let ={gAx(f):fF,gΓ}\mathscr{F}=\{g\mathrm{Ax}(f):f\in F,g\in\Gamma\}. The following notion of an admissible path allows to construct a quasi-geodesic by concatenating geodesics via \mathscr{F}.

X0X_{0}XiX_{i}Xi+1X_{i+1}XnX_{n}πXi(qi),\|\pi_{X_{i}}(q_{i})\|,πXi+1(qi)τ\|\pi_{X_{i+1}}(q_{i})\|\leq\taupip_{i}p0p_{0}pi+1p_{i+1}qiq_{i}pnp_{n}L\geq LL\geq L
Figure 3. Admissible path
Definition 2.6 (Admissible Path).

Given L,τ0L,\tau\geq 0, a path γ\gamma is called (L,τ)(L,\tau)-admissible in X\mathrm{X}, if γ\gamma is a concatenation of geodesics p0q1p1qnpnp_{0}q_{1}p_{1}\cdots q_{n}p_{n} (n)(n\in\mathbb{N}), where the two endpoints of each pip_{i} lie in some XiX_{i}\in\mathscr{F}, and the following Long Local and Bounded Projection properties hold:

  1. (LL)

    Each pip_{i} for 1i<n1\leq i<n has length bigger than LL, and p0,pnp_{0},p_{n} could be trivial;

  2. (BP)

    For each XiX_{i}, we have XiXi+1X_{i}\neq X_{i+1} and max{πXi(qi),πXi(qi+1)}τ\max\{\|\pi_{X_{i}}(q_{i})\|,\|\pi_{X_{i}}(q_{i+1})\|\}\leq\tau, where q0:=γq_{0}:=\gamma_{-} and qn+1:=γ+q_{n+1}:=\gamma_{+} by convention.

The collection {Xi:1in}\{X_{i}:1\leq i\leq n\} is referred to as contracting subsets associated with the admissible path.

Remark 2.7.
  1. (1)

    The path qiq_{i} could be allowed to be trivial, so by the (BP) condition, it suffices to check XiXi+1X_{i}\neq X_{i+1}. It will be useful to note that admissible paths could be concatenated as follows: Let p0q1p1qnpnp_{0}q_{1}p_{1}\cdots q_{n}p_{n} and p0q1p1qnpnp_{0}^{\prime}q_{1}^{\prime}p_{1}^{\prime}\cdots q_{n}^{\prime}p_{n}^{\prime} be (L,τ)(L,\tau)–admissible. If pn=p0p_{n}=p_{0}^{\prime} has length bigger than LL, then the concatenation (p0q1p1qnpn)(q1p1qnpn)(p_{0}q_{1}p_{1}\cdots q_{n}p_{n})\cdot(q_{1}^{\prime}p_{1}^{\prime}\cdots q_{n}^{\prime}p_{n}^{\prime}) has a natural (L,τ)(L,\tau)–admissible structure.

  2. (2)

    In many situations, τ\tau could be chosen as the bounded projection constant BB of \mathscr{F}. In fact, if LL is large relative to τ\tau, an (L,τ)(L,\tau)–admissible path could be always truncated near contracting subsets XiX_{i} so that it becomes an (L^,B)(\hat{L},B)–admissible path. See [72, Lemma 2.14].

We frequently construct a path labeled by a word (g1,g2,,gn)(g_{1},g_{2},\cdots,g_{n}), which by convention means the following concatenation

[o,g1o]g1[o,g2o](g1gn1)[o,gno][o,g_{1}o]\cdot g_{1}[o,g_{2}o]\cdots(g_{1}\cdots g_{n-1})[o,g_{n}o]

where the basepoint oo is understood in context. With this convention, the paths labeled by (g1,g2,g3)(g_{1},g_{2},g_{3}) and (g1g2,g3)(g_{1}g_{2},g_{3}) respectively differ, depending on whether [o,g1o]g1[o,g2o][o,g_{1}o]g_{1}[o,g_{2}o] is a geodesic or not.

A sequence of points xix_{i} in a path pp is called linearly ordered if xi+1[xi,p+]px_{i+1}\in[x_{i},p^{+}]_{p} for each ii.

Definition 2.8 (Fellow travel).

Let γ=p0q1p1qnpn\gamma=p_{0}q_{1}p_{1}\cdots q_{n}p_{n} be an (L,τ)(L,\tau)-admissible path. We say γ\gamma has rr–fellow travel property for some r>0r>0 if for any geodesic α\alpha with the same endpoints as γ\gamma, there exists a sequence of linearly ordered points zi,wiz_{i},w_{i} (0in0\leq i\leq n) on α\alpha such that

d(zi,pi)r,d(wi,pi+)r.d(z_{i},p_{i}^{-})\leq r,\quad d(w_{i},p_{i}^{+})\leq r.

In particular, Nr(Xi)αL\|N_{r}(X_{i})\cap\alpha\|\geq L for each Xi(γ)X_{i}\in\mathscr{F}(\gamma).

The following result says that a local long admissible path enjoys the fellow travel property.

Proposition 2.9.

[69] For any τ>0\tau>0, there exist L,r>0L,r>0 depending only on τ,C\tau,C such that any (L,τ)(L,\tau)–admissible path has rr–fellow travel property. In particular, it is a cc–quasi-geodesic.

The next lemma gives a way to build admissible paths.

Lemma 2.10 (Extension Lemma).

For any independent contracting elements h1,h2,h3Γh_{1},h_{2},h_{3}\in\Gamma, there exist constants L,r,B>0L,r,B>0 depending only on CC with the following property.

Choose any element fihif_{i}\in\langle h_{i}\rangle for each 1i31\leq i\leq 3 to form the set FF satisfying FominL\|Fo\|_{\min}\geq L. Let g,hΓg,h\in\Gamma be any two elements.

  1. (1)

    There exists an element fFf\in F such that the path

    γ:=[o,go](g[o,fo])(gf[o,ho])\gamma:=[o,go]\cdot(g[o,fo])\cdot(gf[o,ho])

    is an (L,τ)(L,\tau)–admissible path relative to \mathscr{F}.

  2. (2)

    The point gogo is an (r,f)(r,f)–barrier for any geodesic [γ,γ+][\gamma^{-},\gamma^{+}].

Remark 2.11.

Since admissible paths are local conditions, we can connect via FF any number of elements gGg\in G to satisfy (1) and (2). We refer the reader to [71] for a precise formulation.

The following elementary fact will be invoked frequently.

Lemma 2.12.

Assume that two words (g1,f1,h1)(g_{1},f_{1},h_{1}) and (g2,f2,h2)(g_{2},f_{2},h_{2}) label two (L,τ)(L,\tau)–admissible paths with the same endpoints, where (C,τ,L)(C,\tau,L) satisfy Proposition 2.9. For any Δ>1\Delta>1, there exists R=R(Δ,C,τ)R=R(\Delta,C,\tau) with the following property. If |d(o,g1o)d(o,g2o)|Δ|d(o,g_{1}o)-d(o,g_{2}o)|\leq\Delta and d(g1o,g2o)>Rd(g_{1}o,g_{2}o)>R, then g1=g2g_{1}=g_{2}.

2.3. Horofunction boundary

We recall the notion of horofunction boundary.

Fix a basepoint oXo\in\mathrm{X}. For each yXy\in\mathrm{X}, we define a Lipschitz map by:XXb_{y}:\mathrm{X}\to\mathrm{X} by

xX:by(x)=d(x,y)d(o,y).\forall x\in\mathrm{X}:\quad b_{y}(x)=d(x,y)-d(o,y).

This family of 11–Lipschitz functions sits in the set of continuous functions on X\mathrm{X} vanishing at oo. Endowed with the compact-open topology, the Arzela-Ascoli Lemma implies that the closure of {by:yX}\{b_{y}:y\in\mathrm{X}\} gives a compactification of X\mathrm{X}. The complement of XX in this compactification is called the horofunction boundary of X\mathrm{X} and is denoted by hX\partial_{h}{\mathrm{X}}.

A Buseman cocycle Bξ:X×XB_{\xi}:\mathrm{X}\times\mathrm{X}\to\mathbb{R} (independent of oo) is given by

x1,x2X:Bξ(x1,x2)=bξ(x1)bξ(x2).\forall x_{1},x_{2}\in\mathrm{X}:\quad B_{\xi}(x_{1},x_{2})=b_{\xi}(x_{1})-b_{\xi}(x_{2}).

The topological type of horofunction boundary is independent of the choice of a basepoint. Every isometry ϕ\phi of X\mathrm{X} induces a homeomorphism on X¯\overline{\mathrm{X}}:

yX:ϕ(ξ)(y):=bξ(ϕ1(y))bξ(ϕ1(o)).\forall y\in\mathrm{X}:\quad\phi(\xi)(y):=b_{\xi}(\phi^{-1}(y))-b_{\xi}(\phi^{-1}(o)).

Depending on the context, we may use both ξ\xi and bξb_{\xi} to denote a point in the horofunction boundary.

Finite difference relation.

Two horofunctions bξ,bηb_{\xi},b_{\eta} have KK–finite difference for K0K\geq 0 if the LL^{\infty}–norm of their difference is KK–bounded:

bξbηK.\|b_{\xi}-b_{\eta}\|_{\infty}\leq K.

The locus of bξb_{\xi} consists of horofunctions bηb_{\eta} so that bξ,bηb_{\xi},b_{\eta} have KK–finite difference for some K>0K>0. The loci [bξ][b_{\xi}] of horofunctions bξb_{\xi} form a finite difference equivalence relation [][\cdot] on hX\partial_{h}{\mathrm{X}}. The locus [Λ][\Lambda] of a subset ΛhX\Lambda\subseteq\partial_{h}{\mathrm{X}} is the union of loci of all points in Λ\Lambda.

If xnXξXx_{n}\in\mathrm{X}\to\xi\in\partial{\mathrm{X}} and ynXηXy_{n}\in\mathrm{X}\to\eta\in\partial{\mathrm{X}} are sequences with supn1d(xn,yn)<\sup_{n\geq 1}d(x_{n},y_{n})<\infty, then [ξ]=[η][\xi]=[\eta].

2.4. Convergence boundary

Let (X,d)(\mathrm{X},d) be a proper metric space admitting an isometric action of a non-elementary countable group Γ\Gamma with a contracting element. Consider a metrizable compactification X¯:=XX\overline{\mathrm{X}}:=\partial{\mathrm{X}}\cup\mathrm{X}, so that X\mathrm{X} is open and dense in X¯\overline{\mathrm{X}}. We also assume that the action of Isom(X)\mathrm{Isom}(\mathrm{X}) extends by homeomorphism to X\partial{\mathrm{X}}.

We equip X\partial{\mathrm{X}} with a Isom(X)\mathrm{Isom}(\mathrm{X})–invariant partition [][\cdot]: [ξ]=[η][\xi]=[\eta] implies [gξ]=[gη][g\xi]=[g\eta] for any gIsom(X)g\in\mathrm{Isom}(\mathrm{X}). We say that ξ\xi is minimal if [ξ]={ξ}[\xi]=\{\xi\}, and a subset UU is [][\cdot]–saturated if U=[U]U=[U].

We say that [][\cdot] restricts to be a closed partition on a [][\cdot]–saturated subset UXU\subseteq\partial{\mathrm{X}} if xnUξXx_{n}\in U\to\xi\in\partial{\mathrm{X}} and ynUηXy_{n}\in U\to\eta\in\partial{\mathrm{X}} are two sequences with [xn]=[yn][x_{n}]=[y_{n}], then [ξ]=[η][\xi]=[\eta]. (The points ξ,η\xi,\eta are not necessarily in UU.) If U=XU=\partial{\mathrm{X}}, this is equivalent to saying that the relation {(ξ,η):[ξ]=[η]}\{(\xi,\eta):[\xi]=[\eta]\} is a closed subset in X×X\partial{\mathrm{X}}\times\partial{\mathrm{X}}, so the quotient space [X][\partial{\mathrm{X}}] is Hausdorff. In general, [][\cdot] may not be closed over the whole X\partial{\mathrm{X}} (e.g., the horofunction boundary with finite difference relation), but is closed when restricted to certain interesting subsets; see for example Assumption (C) below.

We say that xnx_{n} tends (resp. accumulates) to [ξ][\xi] if the limit point (resp. any accumulate point) is contained in the subset [ξ][\xi]. This implies that [xn][x_{n}] tends or accumulates to [ξ][\xi] in the quotient space [Λ(Γo)][\Lambda(\Gamma o)]. So, an infinite ray γ\gamma terminates at [ξ]X[\xi]\in\partial{\mathrm{X}} if any sequence of points in γ\gamma accumulates in [ξ][\xi].

πγ(yn)\pi_{\gamma}(y_{n})yny_{n}[ξ][\xi]ooγ1\gamma_{1}[ξ][\xi]γn\gamma_{n}y1Ωo(γ1)y_{1}\in\Omega_{o}(\gamma_{1})ynΩo(γn)y_{n}\in\Omega_{o}(\gamma_{n})ooooxnx_{n}yny_{n}[ξ]𝒞[\xi]\subseteq\mathcal{C}\Longrightarrowd(o,[xn,yn])d(o,[x_{n},y_{n}])\rightarrow\infty(A)(B)(C)
Figure 4. Assumptions (A)(B)(C) in convergence boundary

Recall that Ωx(A):={yX:[x,y]A}\Omega_{x}(A):=\{y\in\mathrm{X}:\exists[x,y]\cap A\neq\emptyset\} is the cone of a subset AXA\subseteq\mathrm{X} with light source at xx. A sequence of subsets AnXA_{n}\subseteq\mathrm{X} is escaping if d(o,An)d(o,A_{n})\to\infty for some (or any) oXo\in\mathrm{X}.

Definition 2.13.

We say that (X¯,[])(\overline{\mathrm{X}},[\cdot]) is a convergence compactification of X\mathrm{X} if the following assumptions hold.

  1. (A)

    Any contracting geodesic ray γ\gamma accumulates into a closed subset [ξ][\xi] for some ξX\xi\in\partial{\mathrm{X}}; and any sequence ynXy_{n}\in\mathrm{X} with escaping projections πγ(yn)\pi_{\gamma}(y_{n}) tends to [ξ][\xi].

  2. (B)

    Let {XnX:n1}\{X_{n}\subseteq\mathrm{X}:n\geq 1\} be an escaping sequence of CC–contracting quasi-geodesics for some C>0C>0. Then for any given oXo\in\mathrm{X}, there exists a subsequence of {Yn:=Ωo(Xn):n1}\{Y_{n}:=\Omega_{o}(X_{n}):n\geq 1\} (still denoted by YnY_{n}) and ξX\xi\in\partial{\mathrm{X}} such that YnY_{n} accumulates to [ξ][\xi]: any convergent sequence of points ynYny_{n}\in Y_{n} tends to a point in [ξ][\xi].

  3. (C)

    The set 𝒞\mathcal{C} of non-pinched points ξX\xi\in\partial{\mathrm{X}} is non-empty. We say ξ\xi is non-pinched if xn,ynXx_{n},y_{n}\in\mathrm{X} are two sequences of points converging to [ξ][\xi], then [xn,yn][x_{n},y_{n}] is an escaping sequence of geodesic segments. Moreover, for any ξ𝒞\xi\in\mathcal{C}, the partition [][\cdot] restricted to G[ξ]G[\xi] forms a closed relation.

Remark 2.14.

Assumption (C) is a non-triviality condition: any compactification with the coarsest partition asserting X\partial{\mathrm{X}} as one [][\cdot]–class satisfy (A) and (B). We require 𝒞\mathcal{C} to be non-empty to exclude such situations. The “moreover” statement is newly added relative to the one in [72]. If the restriction of the partition to 𝒞\mathcal{C} is maximal in the sense that every [][\cdot]–class is singleton, then the “moreover” assumption is always true.

By Assumption (A), a contracting quasi-geodesic ray γ\gamma determines a unique [][\cdot]–class of a boundary point denoted by [γ+][\gamma^{+}]. Similarly, the positive and negative rays of a bi-infinite contracting quasi-geodesic γ:(,)X\gamma:(-\infty,\infty)\to\mathrm{X} determine respectively two boundary [][\cdot]–classes denoted by [γ+][\gamma^{+}] and [γ][\gamma^{-}].

Examples 2.15.

The first three convergence boundaries below are equipped with a maximal partition [][\cdot] (that is, [][\cdot]–classes are singletons).

  1. (1)

    Hyperbolic space X\mathrm{X} with Gromov boundary X\partial{\mathrm{X}}, where all boundary points are non-pinched.

  2. (2)

    CAT(0) space X\mathrm{X} with visual boundary X\partial{\mathrm{X}} (homeomorphic to horofunction boundary), where all boundary points are non-pinched.

  3. (3)

    The Cayley graph X\mathrm{X} of a relatively hyperbolic group with Bowditch or Floyd boundary X\partial{\mathrm{X}}, where conical limit points are non-pinched.

    If X\mathrm{X} is infinitely ended, we could also take X\partial{\mathrm{X}} as the end boundary with the same statement.

  4. (4)

    Teichmüller space X\mathrm{X} with Thurston boundary X\partial{\mathrm{X}}, where [][\cdot] is given by Kaimanovich-Masur partition [41] and uniquely ergodic points are non-pinched.

  5. (5)

    Any proper metric space X\mathrm{X} with horofunction boundary X\partial{\mathrm{X}}, where [][\cdot] is given by finite difference partition and all boundary points are non-pinched. If X\mathrm{X} is the cubical CAT(0) space, the horofunction boundary is exactly the Roller boundary. If X\mathrm{X} is the Teichmüller space with Teichmüller metric, the horofunction boundary is the Gardiner-Masur boundary ([47, 68]).

From these examples, one can see that the convergence boundary is not a canonical object associated to a given proper metric space. In some sense, the horofunction boundary provides a “universal” convergence boundary with non-pinched points for any proper action.

Theorem 2.16.

[72, Theorem 1.1] The horofunction boundary is a convergence boundary with finite difference relation [][\cdot], where all boundary points are non-pinched.

Proof.

Only the “moreover” statement in Assumption (C) requires clarification. Under the finite difference relation [][\cdot], if [ξ][\xi] is KK–bounded for some K>0K>0, then any g[ξ]g[\xi] is KK–bounded as well. According to the topology of the horofunction compactification, if |bxn()byn()|K|b_{x_{n}}(\cdot)-b_{y_{n}}(\cdot)|\leq K and xnξ,ynηx_{n}\to\xi,y_{n}\to\eta, then |bξ()bη()|K|b_{\xi}(\cdot)-b_{\eta}(\cdot)|\leq K. This shows [ξ]=[η][\xi]=[\eta] (however, [ξ][\xi] may not be KK–bounded). ∎

If fIsom(X)f\in\mathrm{Isom}(\mathrm{X}) is a contracting element, the positive and negative rays of a contracting quasi-geodesic nfnon\in\mathbb{Z}\mapsto f^{n}o determine two boundary points denoted by [f+][f^{+}] and [f][f^{-}] respectively. Write [f±]=[f+][f][f^{\pm}]=[f^{+}]\cup[f^{-}]. We say that ff is non-pinched if [f±]𝒞[f^{\pm}]\subseteq\mathcal{C} are non-pinched. On the horofunction boundary, any contracting element is non-pinched by Theorem 2.16. In contrast, there exist pinched contracting elements in the Bowditch boundary of a relatively hyperbolic group; for instance, those contracting elements in a parabolic group are pinched. The following two results fail if the non-pinched condition is dropped.

Recall that Γ<Isom(X)\Gamma<\mathrm{Isom}(\mathrm{X}) is a discrete group, which we assume in the next results.

Lemma 2.17.

If fΓf\in\Gamma is a non-pinched contracting element, then E(f)E(f) coincides with the set stabilizer of the two fixed points [f±][f^{\pm}].

Proof.

By Assumption A, the two half ray of the axis Ax(f)\mathrm{Ax}(f) accumulates either to [f][f^{-}] or to [f+][f^{+}]. The same holds for gAx(f)g\mathrm{Ax}(f), which accumulates to g[f±]g[f^{\pm}]. By definition, E(f)E(f) is the set of elements gGg\in G which preserves Ax(f)\mathrm{Ax}(f) up to finite Hausdorff distance. Let gGg\in G preserve [f±][f^{\pm}] but gE(f)g\notin E(f). Choose two sequence of points xn,yngAx(f)x_{n},y_{n}\in g\mathrm{Ax}(f) so that xn[f],yn[f+]x_{n}\to[f^{-}],y_{n}\in[f^{+}], and let znπAx(f)(xn)z_{n}\in\pi_{\mathrm{Ax}(f)}(x_{n}) and wnπAx(f)(yn)w_{n}\in\pi_{\mathrm{Ax}(f)}(y_{n}). We claim that both znz_{n} and wnw_{n} are unbounded sequences. If not, assume for concreteness that znz_{n} is bounded. Choose unAx(f)u_{n}\in\mathrm{Ax}(f) tending to [f][f^{-}] ([f][f^{-}], if wnw_{n} is bounded). The contracting property implies [xn,un]B(zn,2C)[x_{n},u_{n}]\cap B(z_{n},2C)\neq\emptyset, where CC is the contracting constant of Ax(f)\mathrm{Ax}(f). As znz_{n} is bounded and xn,unx_{n},u_{n} tend to [f][f^{-}], this contradicts [f]𝒞[f^{-}]\subseteq\mathcal{C} by Definition 2.13(C). ∎

Lemma 2.18.

[72, Lemma 3.12] If f,gΓf,g\in\Gamma are two non-pinched independent contracting elements, then either [f±]=[g±][f^{\pm}]=[g^{\pm}] or [f±][g±]=[f^{\pm}]\cap[g^{\pm}]=\emptyset.

Lemma 2.19.

[72, Lemma 3.10] Let g,fΓg,f\in\Gamma be two independent contracting elements. Then for all n0n\gg 0, hn:=gnfnh_{n}:=g^{n}f^{n} is a contracting element so that [hn+][h_{n}^{+}] tends to [g+][g^{+}] and [hn][h_{n}^{-}] tends to [f][f^{-}].

The following lemma is often used to choose an independent contracting element.

Lemma 2.20.

Let H<Isom(X)H<\mathrm{Isom}(\mathrm{X}) be a non-elementary discrete group. Assume that fHf\in H is a contracting element. Then there exists a contracting element hHh\in H that is independent with ff.

Proof.

The set of translated axis :={gAx(f):gΓ}\mathscr{F}:=\{g\mathrm{Ax}(f):g\in\Gamma\} has bounded intersection. Pick two distinct conjugates denoted by f1,f2f_{1},f_{2} of ff. By Lemma 2.19, h:=f1nf2nh:=f_{1}^{n}f_{2}^{n} is a contracting element whose fixed points tend to [f1+][f_{1}^{+}] and [f2][f_{2}^{-}] respectively for any large n0n\gg 0. Note that the axis Ax(h)\mathrm{Ax}(h) of hh has an overlap with Ax(f1)\mathrm{Ax}(f_{1}) and Ax(f2)\mathrm{Ax}(f_{2}) of a large length d(o,fino)d(o,f_{i}^{n}o) for i=1,2i=1,2. Taking into account every pair of elements in \mathscr{F} has bounded intersection, we conclude that Ax(h)\mathrm{Ax}(h) has bounded intersection with every element in \mathscr{F}. That is, hh is independent with ff by definition. ∎

Let UU be a subset of X\mathrm{X}. The limit set of UU, denoted by Λ(U)\Lambda(U), is the set of all accumulation points of UU in X\partial{\mathrm{X}}. If H<ΓH<\Gamma is a subgroup, Λ(Ho)\Lambda(Ho) may depend on the choice of the basepoint oXo\in\mathrm{X}, but the [][\cdot]–loci [Λ(Ho)][\Lambda(Ho)] does not by Definition 2.13(B). For this reason, we shall refer [Λ(Ho)][\Lambda(Ho)] as the limit set of a subgroup HH. Moreover, the limit set of a discrete group Γ\Gamma enjoys the following desirable property as in the general theory of convergence group action.

Lemma 2.21.

[72, Lemma 3.9] Assume that Γ\Gamma is non-elementary. Let fΓf\in\Gamma be a contracting element. Then [Λ(Γo)]=[Γξ¯][\Lambda(\Gamma o)]=[\overline{\Gamma\xi}] for any ξ[f±]\xi\in[f^{\pm}].

2.5. Quasi-conformal density and Patterson’s construction

Let X¯:=XX\overline{\mathrm{X}}:=\partial{\mathrm{X}}\cup\mathrm{X} be a convergence compactification, with a nonempty set 𝒞\mathcal{C} of non-pinched points in Definition 2.13. We need to restrict to a smaller subset of 𝒞\mathcal{C}, on which some weak continuity of Busemann cocycles can be guaranteed.

Assumption D.

There exists a subset 𝒞hor𝒞\mathcal{C}^{\mathrm{hor}}\subseteq\mathcal{C} with a family of Buseman quasi-cocycles

{Bξ:X×X}ξ𝒞hor\big{\{}B_{\xi}:\quad\mathrm{X}\times\mathrm{X}\longrightarrow\mathbb{R}\big{\}}_{\xi\in\mathcal{C}^{\mathrm{hor}}}

so that for any x,yXx,y\in\mathrm{X}, we have

(8) lim supz[ξ]|Bξ(x,y)Bz(x,y)|ϵ,\displaystyle\limsup_{z\to[\xi]}|B_{\xi}(x,y)-B_{z}(x,y)|\leq\epsilon,

where ϵ0\epsilon\geq 0 may depend on [ξ][\xi] but not on x,yx,y.

For a given ϵ\epsilon, let 𝒞ϵhor\mathcal{C}^{\mathrm{hor}}_{\epsilon} denote the set of points ξ𝒞hor\xi\in\mathcal{C}^{\mathrm{hor}} for which (8) holds. Thus, 𝒞hor=ϵ0𝒞ϵhor\mathcal{C}^{\mathrm{hor}}=\bigcup_{\epsilon\geq 0}\mathcal{C}^{\mathrm{hor}}_{\epsilon}.

In practice, the assumption (8) could allow to take the following concrete definition:

Bξ(x,y)\displaystyle B_{\xi}(x,y) =lim supznξ[d(x,zn)d(y,zn)]\displaystyle=\limsup_{z_{n}\to\xi}\left[d(x,z_{n})-d(y,z_{n})\right]

The following facts shall be used implicitly.

(9) Bξ(x,y)d(x,y)Bξ(x,y)=Bgξ(gx,gy),gIsom(X)\begin{array}[]{l}B_{\xi}(x,y)\leq d(x,y)\\ B_{\xi}(x,y)=B_{g\xi}(gx,gy),\;\forall g\in\mathrm{Isom}(\mathrm{X})\end{array}
Horofunctions in convergence boundary

We now define Buseman quasi-cocycles at a [][\cdot]–class [ξ][\xi]. Namely, given ξ𝒞ϵhor\xi\in\mathcal{C}^{\mathrm{hor}}_{\epsilon}, define a Busemann quasi-cocycle at ξ\xi (resp. [ξ][\xi]) as follows

B[ξ](x,y)=lim supzn[ξ][d(x,zn)d(y,zn)]\displaystyle B_{[\xi]}(x,y)=\limsup_{z_{n}\to[\xi]}\left[d(x,z_{n})-d(y,z_{n})\right]

where the convergence znXξz_{n}\in\mathrm{X}\to\xi (resp. zn[ξ]z_{n}\to[\xi] ) takes place in X¯=XX\overline{\mathrm{X}}=\mathrm{X}\cup\partial{\mathrm{X}}. The equations in (9) hold for B[ξ](x,y)B_{[\xi]}(x,y), and moreover, for all ξ[ξ]\xi\in[\xi],

(10) |Bξ(x,y)B[ξ](x,y)|ϵ|B[ξ](x,y)B[ξ](x,z)B[ξ](z,y)|2ϵ\begin{array}[]{rl}|B_{\xi}(x,y)-B_{[\xi]}(x,y)|&\leq\epsilon\\ |B_{[\xi]}(x,y)-B_{[\xi]}(x,z)-B_{[\xi]}(z,y)|&\leq 2\epsilon\end{array}
Horoballs at [][\cdot]–classes

We give the following tentative definition of horoballs at a [][\cdot]–class in a general convergence boundary. Given a real number LL, we define the horoball of (algebraic) depth LL. We take the convention that the center ξ\xi has depth -\infty.

Definition 2.22.

Let ξ𝒞ϵhor\xi\in\mathcal{C}^{\mathrm{hor}}_{\epsilon} for ϵ0\epsilon\geq 0. We define a horoball centered at [ξ][\xi] of depth LL as follows

([ξ],L)={yX:B[ξ](y,o)L}\mathcal{HB}([\xi],L)=\{y\in\mathrm{X}:B_{[\xi]}(y,o)\leq L\}

We omit LL if L=0L=0 or it does not matter in context.

As the limit supper of continuous functions, B[ξ](x,y)B_{[\xi]}(x,y) is lower semi-continuous, so ([ξ],L)\mathcal{HB}([\xi],L) is a closed subset. By abuse of language, we say the level set {yX:B[ξ](y,o)=L}\{y\in\mathrm{X}:B_{[\xi]}(y,o)=L\} is a horosphere or the boundary of ([ξ],L)\mathcal{HB}([\xi],L).

Lemma 2.23.

For L1>L2L_{1}>L_{2}, we have

  1. (1)

    ([ξ],L1)NL(([ξ],L2))\mathcal{HB}([\xi],L_{1})\subseteq N_{L}(\mathcal{HB}([\xi],L_{2})) for L=L1L2L=L_{1}-L_{2}.

  2. (2)

    ([ξ],L2)\mathcal{HB}([\xi],L_{2}) has distance at least L2ϵL-2\epsilon to the complement X([ξ],L1)\mathrm{X}\setminus\mathcal{HB}([\xi],L_{1}).

Proof.

(1). Observe that, for any xXx\in\mathrm{X} and ξX\xi\in\partial{\mathrm{X}}, there exists a geodesic ray γ\gamma starting at xx so that Bξ(x,y)=d(x,y)B_{\xi}(x,y)=d(x,y) for any yγy\in\gamma. (γ\gamma is called a gradient line in some literature). By an Ascoli-Arzela argument, γ\gamma is obtained as a limiting ray of [x,zn][x,z_{n}] when znξz_{n}\to\xi. Thus, any point x([ξ],L1)x\in\mathcal{HB}([\xi],L_{1}) has a distance LL to some y([ξ],L2)y\in\mathcal{HB}([\xi],L_{2}).

(2). If d(x,y)L2ϵd(x,y)\leq L-2\epsilon for some x([ξ],L2)x\in\mathcal{HB}([\xi],L_{2}) and yXy\in\mathrm{X}, we obtain B[ξ](y,o)B[ξ](y,x)+B[ξ](x,o)+2ϵL+L2L1B_{[\xi]}(y,o)\leq B_{[\xi]}(y,x)+B_{[\xi]}(x,o)+2\epsilon\leq L+L_{2}\leq L_{1} so y([ξ],L1)y\in\mathcal{HB}([\xi],L_{1}). The conclusion follows. ∎

By abuse of language, we also call the set L([ξ],x)\mathcal{HB}_{L}([\xi],x) a horoball in any η[ξ]\eta\in[\xi]. In other words, all points in the same class [ξ][\xi] share a common horoball.

Let +(X¯)\mathcal{M}^{+}(\overline{\mathrm{X}}) be the set of finite positive Radon measures on X¯:=XX\overline{\mathrm{X}}:=\partial{\mathrm{X}}\cup\mathrm{X}, on which Γ\Gamma acts by push-forward: for any Borel set AA,

gμ(A)=μ(g1A).g_{\star}\mu(A)=\mu(g^{-1}A).
Definition 2.24.

Let ω[0,[\omega\in[0,\infty[. A map

μ:X\displaystyle\mu:\quad\mathrm{X}\quad\longrightarrow +(X¯)\displaystyle\quad\mathcal{M}^{+}(\overline{\mathrm{X}})
x\displaystyle x\quad\longmapsto μx\displaystyle\quad\mu_{x}

is a ω\omega–dimensional, Γ\Gamma–quasi-equivariant, quasi-conformal density if for any gΓg\in\Gamma and x,yXx,y\in\mathrm{X},

(11) μgxa.e.ξX:\displaystyle\mu_{gx}-\mathrm{a.e.}\;\xi\in\partial{\mathrm{X}}: dgμxdμgx(ξ)[1λ,λ],\displaystyle\quad\frac{dg_{\star}\mu_{x}}{d\mu_{gx}}(\xi)\in[\frac{1}{\lambda},\lambda],
(12) μya.e.ξX:\displaystyle\mu_{y}-\mathrm{a.e.}\;\xi\in\partial{\mathrm{X}}: 1λeωBξ(x,y)dμxdμy(ξ)λeωBξ(x,y)\displaystyle\quad\frac{1}{\lambda}e^{-\omega B_{\xi}(x,y)}\leq\frac{d\mu_{x}}{d\mu_{y}}(\xi)\leq\lambda e^{-\omega B_{\xi}(x,y)}

for a universal constant λ1\lambda\geq 1. We normalize μo\mu_{o} to be a probability: its mass μo=μo(X¯)=1\|\mu_{o}\|=\mu_{o}(\overline{\mathrm{X}})=1.

If {μx:xX}\{\mu_{x}:x\in\mathrm{X}\} is supported on non-pinched boundary points (i.e.: μo(𝒞)=1\mu_{o}(\mathcal{C})=1), we say it is a non-trivial quasi-conformal density.

Remark 2.25.

A non-trivial quasi-conformal density is much weaker than saying that μo\mu_{o} is supported on the conical points. Take for instance the horofunction boundary X\partial{\mathrm{X}}, where 𝒞=X\mathcal{C}=\partial{\mathrm{X}} is the whole boundary. Moreover, all Patterson-Sullivan measures are non-trivial quasi-conformal density in Examples 1.9.

If λ=1\lambda=1 for (11), the map μ:X+(X¯)\mu:\mathrm{X}\to\mathcal{M}^{+}(\overline{\mathrm{X}}) is Γ\Gamma–equivariant: that is, μgx=gμx\mu_{gx}=g_{\star}\mu_{x} (equivalently, μgx(gA)=μx(A)\mu_{gx}(gA)=\mu_{x}(A)). If in both (11) and (12), λ=1\lambda=1, we call μ\mu a conformal density.

Patterson’s construction of quasi-conformal density

Fix a basepoint oXo\in\mathrm{X}. Consider the orbital points in the ball of radius R>0R>0:

(13) NΓ(o,n)={vΓo:d(o,v)n}N_{\Gamma}(o,n)=\{v\in\Gamma o:d(o,v)\leq n\}

The critical exponent ωA\omega_{A} for a subset AΓA\subseteq\Gamma is independent of the choice of oXo\in\mathrm{X}:

(14) ωA=lim supRlog(NΓ(o,R)Ao)R,\omega_{A}=\limsup\limits_{R\to\infty}\frac{\log\sharp(N_{\Gamma}(o,R)\cap Ao)}{R},

which is intimately related to the (partial) Poincaré series

(15) s0,x,yX,𝒫A(s,x,y)=gAesd(x,gy)s\geq 0,x,y\in\mathrm{X},\quad\mathcal{P}_{A}(s,x,y)=\sum\limits_{g\in A}e^{-sd(x,gy)}

as 𝒫A(s,x,y)\mathcal{P}_{A}(s,x,y) diverges for s<ωAs<\omega_{A} and converges for s>ωAs>\omega_{A}. The action ΓX\Gamma\curvearrowright\mathrm{X} is called of divergence type (resp. convergence type) if 𝒫Γ(s,x,y)\mathcal{P}_{\Gamma}(s,x,y) is divergent (resp. convergent) at s=ωΓs=\omega_{\Gamma}.

Fix Δ1\Delta\geq 1. The family of the annulus-like sets of radius nn centered at oXo\in\mathrm{X}

(16) AΓ(o,n,Δ)={vΓo:|d(o,v)n|Δ}A_{\Gamma}(o,n,\Delta)=\{v\in\Gamma o:|d(o,v)-n|\leq\Delta\}

covers GoGo with multiplicity at most 2Δ2\Delta. It is useful to keep in mind that for s>ωA,x,yX,s>\omega_{A},x,y\in\mathrm{X},

(17) 𝒫A(s,o,o)ΔgAAΓ(o,n,Δ)esd(o,go)\quad\mathcal{P}_{A}(s,o,o)\asymp_{\Delta}\sum\limits_{g\in A\cap A_{\Gamma}(o,n,\Delta)}e^{-sd(o,go)}

Fix yXy\in\mathrm{X}. We start by constructing a family of measures {μxs,y}xX\{\mu_{x}^{s,y}\}_{x\in\mathrm{X}} supported on Γy\Gamma y for any given s>ωΓs>\omega_{\Gamma}. Assume that 𝒫Γ(s,x,y)\mathcal{P}_{\Gamma}(s,x,y) is divergent at s=ωΓs=\omega_{\Gamma}. Set

(18) μxs,y=1𝒫Γ(s,o,y)gΓesd(x,gy)Dirac(gy),\mu_{x}^{s,y}=\frac{1}{\mathcal{P}_{\Gamma}(s,o,y)}\sum\limits_{g\in\Gamma}e^{-sd(x,gy)}\cdot{\mbox{Dirac}}{(gy)},

where s>ωΓs>\omega_{\Gamma} and xXx\in\mathrm{X}. Note that μos,y\mu^{s,y}_{o} is a probability measure supported on Γy\Gamma y. If 𝒫Γ(s,x,y)\mathcal{P}_{\Gamma}(s,x,y) is convergent at s=ωΓs=\omega_{\Gamma}, the Poincaré series in (18) needs to be replaced by a modified series as in [54].

Choose siωΓs_{i}\to\omega_{\Gamma} such that for each xXx\in\mathrm{X}, μxsi,y\mu_{x}^{s_{i},y} is a convergent sequence in (Λ(Γo))\mathcal{M}(\Lambda(\Gamma o)). The family of limit measures μxy=limμxsi,y\mu_{x}^{y}=\lim\mu_{x}^{s_{i},y} (xXx\in\mathrm{X}) are called Patterson-Sullivan measures.

By construction, Patterson-Sullivan measures are by no means unique a priori, depending on the choice of yXy\in\mathrm{X} and siωΓs_{i}\to\omega_{\Gamma}. Note that μoy(Λ(Γo))=1\mu_{o}^{y}(\Lambda(\Gamma o))=1 is a normalized condition, where oXo\in\mathrm{X} is a priori chosen basepoint. In what follows, we usually write μx=μxo\mu_{x}=\mu_{x}^{o} for xXx\in\mathrm{X} (i.e.: y=oy=o).

There are other sources of conformal densities, not necessarily coming from Patterson’s construction. Here we describe in some details the construction of conformal densities on Thurston boundary. See [38] for other examples constructed on ends of hyperbolic 3-manifolds.

Example 2.26.

Consider the Teichmüller space 𝒯g\mathcal{T}_{g} of a closed oriented surface Σg\Sigma_{g} (g2g\geq 2). The space \mathscr{MF} of measured foliations on Σg\Sigma_{g}, which is homeomorphic to 6g6\mathbb{R}^{6g-6}, admits a Mod(Σg)\mathrm{Mod}(\Sigma_{g})–invariant ergodic measure, μTh\mu_{\mathrm{Th}}, called Thurston measure by the work of Masur and Veech. Positive reals >0\mathbb{R}_{>0} act on \mathscr{MF} by scaling, and let

π:𝒫=/>0\pi:\mathscr{MF}\longmapsto\mathscr{PMF}=\mathscr{MF}/\mathbb{R}_{>0}

be the natural quotient map.

𝒫\mathscr{PMF} is also called the Thurston boundary, and it gives a convergence boundary for 𝒯g\mathcal{T}_{g} with equipped with the Teichmüller metric. Here we take the Kaimanovich-Masur partition [][\cdot], whose precise description is not relevant here, but the set 𝒰\mathscr{UE} of uniquely ergodic points in 𝒫\mathscr{PMF} is partitioned into singletons. It is a well-known fact that μTh\mu_{\mathrm{Th}} is supported on π1(𝒰)\pi^{-1}(\mathscr{UE}).

The μTh\mu_{\mathrm{Th}} induces a (6g6)(6g-6)–dimensional, Mod(Σg)\mathrm{Mod}(\Sigma_{g})–equivariant, conformal density on 𝒫\mathscr{PMF} as follows. Given x𝒯gx\in\mathcal{T}_{g}, consider the extremal length function

Extx:0\mathrm{Ext}_{x}:\mathscr{MF}\longmapsto\mathbb{R}_{\geq 0}

which is square homogeneous, Extx(tξ)=t2Extx(ξ)\mathrm{Ext}_{x}(t\xi)=t^{2}\mathrm{Ext}_{x}(\xi). Take the ball-like set BExt(x)={ξ:Extx(ξ)1}.B_{\mathrm{Ext}}(x)=\{\xi\in\mathscr{MF}:\mathrm{Ext}_{x}(\xi)\leq 1\}. For any A𝒫A\subseteq\mathscr{PMF}, set μx(A)=μTh(BExt(x)π1A)\mu_{x}(A)=\mu_{\mathrm{Th}}(B_{\mathrm{Ext}}(x)\cap\pi^{-1}{A}). Direct computation gives

μya.e.ξ𝒫:dμxdμy(ξ)=[Extx(ξ)Exty(ξ)]6g6\mu_{y}-\mathrm{a.e.}\;\xi\in\mathscr{PMF}:\;\frac{d\mu_{x}}{d\mu_{y}}(\xi)=\left[\frac{\sqrt{\mathrm{Ext}_{x}(\xi)}}{\sqrt{\mathrm{Ext}_{y}(\xi)}}\right]^{6g-6}

The family {μx}\{\mu_{x}\} forms a conformal density of dimension 6g66g-6: the term in the right-hand bracket coincides with eBξ(x,y)\mathrm{e}^{-B_{\xi}(x,y)} for ξ𝒰\xi\in\mathscr{UE}, and μx\mu_{x} charges the full measure to 𝒰\mathscr{UE}. Note that this family is the same as the conformal density obtained from the action Mod(Σg)𝒯g\mathrm{Mod}(\Sigma_{g})\curvearrowright\mathcal{T}_{g} through the Patterson construction above (see [72]).

Theorem 2.27.

Suppose that GG acts properly on a proper geodesic space X\mathrm{X} compactified with horofunction boundary hX\partial_{h}{\mathrm{X}}. Then the family {μx:xX}\{\mu_{x}:x\in\mathrm{X}\} of Patterson-Sullivan measures is a ωG\omega_{G}-dimensional GG-equivariant conformal density supported on [ΛGo][\Lambda Go].

In the sequel, we write PS-measures as shorthand for Patterson-Sullivan measures.

Let {μx:xX}\{\mu_{x}:x\in\mathrm{X}\} be a nontrivial ω\omega–dimensional Γ\Gamma–equivariant quasiconformal density on a convergence compactification XX\mathrm{X}\cup\partial{\mathrm{X}}: μo\mu_{o} is supported on the set 𝒞\mathcal{C} of non-pinched points in Definition 2.13(C).

2.6. Shadow Principle

Let FΓF\subseteq\Gamma be a set of three (mutually) independent contracting elements fif_{i} (i=1,2,3i=1,2,3), which form a CC-contracting system

(19) ={gAx(fi):gΓ}\mathscr{F}=\{g\cdot\mathrm{Ax}(f_{i}):g\in\Gamma\}

where the axis Ax(fi)\mathrm{Ax}(f_{i}) in (6) is CC-contracting with CC depending on the choice of the basepoint oXo\in\mathrm{X}. We may often assume d(o,fo)d(o,fo) is large as possible, by taking sufficiently high power of fFf\in F. The contracting constant CC is not effected.

First of all, define the usual cone and shadow:

Ωx(y,r):={zX:[x,z]B(y,r)}\Omega_{x}(y,r):=\{z\in\mathrm{X}:\exists[x,z]\cap B(y,r)\neq\emptyset\}

and Πx(y,r)X\Pi_{x}(y,r)\subseteq\partial{\mathrm{X}} be the topological closure in X\partial{\mathrm{X}} of Ωx(y,r)\Omega_{x}(y,r).

The partial shadows ΠoF(go,r)\Pi_{o}^{F}(go,r) and cones ΩoF(go,r)\Omega_{o}^{F}(go,r) defined in the following depend on the choice of a contracting system \mathscr{F} in (19).

Definition 2.28 (Partial cone and shadow).

For xX,yΓox\in\mathrm{X},y\in\Gamma o, the (r,F)(r,F)–cone ΩxF(y,r)\Omega_{x}^{F}(y,r) is the set of elements zXz\in\mathrm{X} such that yy is a (r,F)(r,F)–barrier for some geodesic [x,z][x,z].

The (r,F)(r,F)–shadow ΠxF(y,r)X\Pi_{x}^{F}(y,r)\subseteq\partial{\mathrm{X}} is the topological closure in X\partial{\mathrm{X}} of the cone ΩxF(y,r)\Omega_{x}^{F}(y,r).

The follow terminology is from Roblin [60].

Definition 2.29.

We say that {μx:xX}\{\mu_{x}:x\in\mathrm{X}\} satisfies Shadow Principle over a subset ZXZ\subseteq\mathrm{X} if there is some large constant r0>0r_{0}>0 so that the following holds

μyeωd(x,y)μx(Πx(y,r))rμyeωd(x,y)\displaystyle\|\mu_{y}\|\mathrm{e}^{-\omega\cdot d(x,y)}\prec\mu_{x}(\Pi_{x}(y,r))\prec_{r}\|\mu_{y}\|\mathrm{e}^{-\omega\cdot d(x,y)}

for any x,yZx,y\in Z and r>r0r>r_{0}, where the implied constant depends on ZZ.

The most fundamental example is provided by the Sullivan’s Shadow Lemma, where ZZ could be taken to be any Γ\Gamma–cocompact subset (by Γ\Gamma–equivariance, μx\|\mu_{x}\| thus remains uniformly bounded over xZx\in Z). Here are some other examples.

Examples 2.30.
  1. (1)

    Roblin realized that ZZ could be enlarged to Z:=N(Γ)oZ:=N(\Gamma)o, provided that the normalizer N(Γ)N(\Gamma) is a discrete subgroup.

  2. (2)

    Let X\mathrm{X} be a rank-1 symmetric space. If ω\omega is the Hausdorff dimension of the visual boundary X\partial{\mathrm{X}}, there exists a unique (up to scaling) ω\omega–dimensional Γ\Gamma–equivariant conformal density {μx:xX}\{\mu_{x}:x\in\mathrm{X}\} on X\partial{\mathrm{X}}. It satisfies the Shadow Principle over the whole space Z=XZ=\mathrm{X}, where μx=1\|\mu_{x}\|=1 for any xXx\in\mathrm{X}.

  3. (3)

    We shall establish the Shadow Principle for confined subgroups in §7.

  4. (4)

    We conjecture that the unique (6g6)(6g-6)–dimensional conformal density on 𝒫\mathscr{PMF} also satisfies the Shadow Principle over Z=𝒯gZ=\mathcal{T}_{g}. See the relevance to Question 1.18 as explained in Remark 10.12.

We now recall the following shadow lemma on the convergence boundary.

Lemma 2.31 ([72, Lemma 6.3]).

Let {μx}xX\{\mu_{x}\}_{x\in\mathrm{X}} be a nontrivial ω\omega–dimensional GG–equivariant quasi-conformal density for some ω>0\omega>0 (i.e. supported on the set 𝒞X\mathcal{C}\subseteq\partial{\mathrm{X}} of non-pinched points). Then there exist r0,L0>0r_{0},L_{0}>0 with the following property.

Assume that d(o,fo)>L0d(o,fo)>L_{0} for each fFf\in F. For given rr0r\geq r_{0}, there exist C0=C0(F),C1=C1(F,r)C_{0}=C_{0}(F),C_{1}=C_{1}(F,r) such that

C0eωd(o,go)μo(ΠoF(go,r))μo(Πo(go,r))C1eωd(o,go)\begin{array}[]{rl}C_{0}\mathrm{e}^{-\omega\cdot d(o,go)}\leq\mu_{o}(\Pi_{o}^{F}(go,r))\leq\mu_{o}(\Pi_{o}(go,r))\leq C_{1}\mathrm{e}^{-\omega\cdot d(o,go)}\end{array}

for any goGogo\in Go.

Remark 2.32.

We emphasize that the ω\omega–dimensional Γ\Gamma–conformal density exists only for ωωΓ\omega\geq\omega_{\Gamma} ([72, Prop. 6.6]). Even if the action ΓX\Gamma\curvearrowright\mathrm{X} is of divergence action (i.e. a non-uniform lattice in Isom(n)\mathrm{Isom}(\mathbb{H}^{n})), a ω\omega–dimensional conformal density may exist for ω>ωΓ\omega>\omega_{\Gamma}. See [38] for relevant discussions.

In what follows, when speaking of ΠoF(go,r)\Pi_{o}^{F}(go,r), we assume r,Fr,F satisfy the Shadow Lemma 2.31.

2.7. Conical points

We now give the definition of a conical point. Recall that 𝒞X\mathcal{C}\subseteq\partial{\mathrm{X}} is the set of non-pinched boundary points in Definition 2.13. Roughly speaking, conical points are non-pinched and shadowed by infinitely many contracting segments with fixed parameter.

Definition 2.33.

A point ξ𝒞\xi\in\mathcal{C} is called (r,F)(r,F)–conical point if for some xΓox\in\Gamma o, the point ξ\xi lies in infinitely many (r,F)(r,F)–shadows ΠxF(yn,r)\Pi_{x}^{F}(y_{n},r) for ynΓoy_{n}\in\Gamma o. We denote by Λr,Fcon(Γo)\Lambda^{\textrm{con}}_{\textrm{r,F}}{(\Gamma o)} the set of (r,F)(r,F)–conical points.

We also denote by Λcon(Γo)\Lambda^{\textrm{con}}{(\Gamma o)} the set of conical points ξ𝒞\xi\in\mathcal{C} for which there exists gnoΓog_{n}o\in\Gamma o satisfying ξΠo(gno,r)\xi\in\Pi_{o}(g_{n}o,r) for some large r>0r>0. By definition, Λr,Fcon(Γo)Λcon(Γo)\Lambda^{\textrm{con}}_{\textrm{r,F}}{(\Gamma o)}\subseteq\Lambda^{\textrm{con}}{(\Gamma o)}.

gnog_{n}ognfnog_{n}f_{n}ognAx(fn)g_{n}\mathrm{Ax}(f_{n})ooΠoF(gno,r)ξ\Pi_{o}^{F}(g_{n}o,r)\ni\ \xir\leq rr\leq r
Figure 5. Conical points

The following is the Borel-Cantelli Lemma stated in terms of conical points.

Lemma 2.34.

Let μ0\mu_{0} be a probability measure on X\partial X that satisfies the shadow lemma as in Lemma 2.31, and let AΓA\subseteq\Gamma be a subset so that the partial Poincaré series 𝒫A(ωΓ,x,y)<\mathcal{P}_{A}(\omega_{\Gamma},x,y)<\infty. Then the following set

Λ:=lim supgAΠoF(go,r)\Lambda:=\limsup_{g\in A}\Pi_{o}^{F}(go,r)

is a μo\mu_{o}–null set, for any r0r\gg 0.

Proof.

By definition, we can write Λ=n1Λn\Lambda=\cap_{n\geq 1}\Lambda_{n} where

Λn=gA:d(o,go)nΠoF(go,r)\Lambda_{n}=\bigcup_{g\in A:d(o,go)\geq n}\Pi_{o}^{F}(go,r)

If μ1(Λ)>0\mu_{1}(\Lambda)>0 then μ1(Λn)>μ1(Λ)/2\mu_{1}(\Lambda_{n})>\mu_{1}(\Lambda)/2 for all n0n\gg 0. On the other hand, the shadow lemma implies

μo(Λn)gA:d(o,go)nμo(ΠoF(go,r))gA:d(o,go)neωΓd(o,go)\mu_{o}(\Lambda_{n})\leq\sum_{g\in A:d(o,go)\geq n}\mu_{o}(\Pi_{o}^{F}(go,r))\leq\sum_{g\in A:d(o,go)\geq n}\mathrm{e}^{-\omega_{\Gamma}d(o,go)}

so the tails of the series 𝒫A(ωΓ,x,y)\mathcal{P}_{A}(\omega_{\Gamma},x,y) has a uniform positive lower bound, contradicting to the assumption. The proof is complete. ∎

We list the following useful properties from [72, Lemma 4.6, Theorem 1.10, Lemma 5.5] about conical points.

Lemma 2.35.

The following holds for any ξΛr,Fcon(Γo)\xi\in\Lambda^{\textrm{con}}_{\textrm{r,F}}{(\Gamma o)}:

  1. (1)

    ξ\xi is visual: for any basepoint oXo\in\mathrm{X} there exists a geodesic ray starting at oo ending at [ξ][\xi].

  2. (2)

    If GXG\curvearrowright\mathrm{X} is of divergence type, μo\mu_{o} charges full measure on [Λr,Fcon(Γo)][\Lambda^{\textrm{con}}_{\textrm{r,F}}{(\Gamma o)}] and every [ξ][\xi]–class is μo\mu_{o}–null.

  3. (3)

    Let yn,znXy_{n},z_{n}\in\mathrm{X} be two sequences tending to [ξ][\xi] in X¯=XX\overline{\mathrm{X}}=\mathrm{X}\cup\partial{\mathrm{X}}. Then for any xXx\in\mathrm{X},

    lim supn|Byn(x)Bzn(x)|20C.\limsup_{n\to\infty}|B_{y_{n}}(x)-B_{z_{n}}(x)|\leq 20C.

    In particular, [Λr,Fcon(Γo)][\Lambda^{\textrm{con}}_{\textrm{r,F}}{(\Gamma o)}] is a subset of 𝒞20Chor\mathcal{C}^{\mathrm{hor}}_{20C} (defined in Assumption D).

It is well known that in hyperbolic spaces, a horoball centered at a point in the Gromov boundary has the unique limit point at the center. This could be proved for uniquely ergodic points in Thurston boundary of Teichmüller spaces. Analogous to these facts, we have the following.

Lemma 2.36.

Let ξΛr,Fcon(Γo)\xi\in\Lambda^{\textrm{con}}_{\textrm{r,F}}{(\Gamma o)}. Consider a horoball ([ξ])\mathcal{HB}([\xi]) centered at [ξ][\xi] defined as Definition 2.22. Then any escaping sequence xn([ξ])x_{n}\in\mathcal{HB}([\xi]) has the accumulation points in [ξ][\xi].

Proof.

According to definition, there exists a sequence of elements gnΓg_{n}\in\Gamma so that ξΠrF(gno)\xi\in\Pi_{r}^{F}(g_{n}o). By [72, Lemma 4.4], there exists a geodesic ray γ\gamma starting from oo terminating at [ξ][\xi] so that

γNr(gnAx(fn))L\|\gamma\cap N_{r}(g_{n}\mathrm{Ax}(f_{n}))\|\geq L

where fnFf_{n}\in F and L:=min{d(o,fo):fF}2rL:=\min\{d(o,fo):f\in F\}-2r.

Fix any X:=gnAx(fn)X:=g_{n}\mathrm{Ax}(f_{n}), and consider the exit point yy of γ\gamma at NC(X)N_{C}(X). Let zmγ[ξ]z_{m}\in\gamma\to[\xi] as mm\to\infty. Given xn([ξ])x_{n}\in\mathcal{HB}([\xi]), the following holds for m0m\gg 0 by Lemma 2.35(3):

(20) Bzm(xn)=d(xn,zm)d(o,zm)21C.B_{z_{m}}(x_{n})=d(x_{n},z_{m})-d(o,z_{m})\leq 21C.

We claim that the projection ynXy_{n}\in X of all but finitely many xnx_{n} to XX lies in a 10C10C–neighborhood of yy. Indeed, if not, d(yn,y)>10Cd(y_{n},y)>10C holds for infinitely many nn. The contracting property of XX implies that [zm,xn][z_{m},x_{n}] intersects NC(X)N_{C}(X) for all n,m0n,m\gg 0, so d(y,[xn,xm])2Cd(y,[x_{n},x_{m}])\leq 2C. Combined with (20), we obtain d(y,xn)d(y,o)+100Cd(y,x_{n})\leq d(y,o)+100C. This is a contradiction, as xnx_{n} is a unbounded sequence in a fixed ball around yy.

If L>100CL>100C is assumed, the above claim implies the corresponding projections of oo and xnx_{n} to XX have distance at least 90C90C, so the contracting property shows [o,xn]NC(X)[o,x_{n}]\cap N_{C}(X)\neq\emptyset. That is, xnΩo(NC(X))x_{n}\in\Omega_{o}(N_{C}(X)) for infinitely many xnx_{n}. As X=gnAx(fn)X=g_{n}\mathrm{Ax}(f_{n}) is chosen arbitrarily along γ\gamma, the assumption (B) in Definition 2.13 implies that xnx_{n} and gnog_{n}o have the same limit, so xn[ξ]x_{n}\to[\xi]. ∎

Uniqueness of quasi-conformal measures

Suppose that ΓX\Gamma\curvearrowright\mathrm{X} is of divergence type, where X\mathrm{X} is compactified by the horofunction boundary hX\partial_{h}{\mathrm{X}}. The quasi-conformal measures on hX\partial_{h}{\mathrm{X}} are in general neither ergodic nor unique. To remedy this, we shall push the measures to the reduced horofunction boundary [hX][\partial_{h}{\mathrm{X}}], modding out the finite difference relation. However, [hX][\partial_{h}{\mathrm{X}}] is a pathological topological object (i.e. may be non-Hausdorff). The remedy is to consider the reduced Myrberg limit set, which provides better topological property. Define

(21) ΛMyr(Γo)=Λr,Fcon(Γo)\Lambda^{\mathrm{Myr}}{(\Gamma o)}=\bigcap\Lambda^{\textrm{con}}_{\textrm{r,F}}{(\Gamma o)}

where the intersection is taken over all possible choice of three independent contracting elements FF in Γ\Gamma. By [72, Lemma 2.27 & Lemma 5.6], the quotient map

ΛMyr(Γo)\displaystyle\Lambda^{\mathrm{Myr}}{(\Gamma o)}\quad\longrightarrow [ΛMyr(Γo)]\displaystyle\quad[\Lambda^{\mathrm{Myr}}{(\Gamma o)}]
ξ\displaystyle\xi\quad\longmapsto [ξ]\displaystyle\quad[\xi]

is a closed map with compact fibers, thus [ΛMyr(Γo)][\Lambda^{\mathrm{Myr}}{(\Gamma o)}] is a completely metrizable and second countable topological space.

Let {μx}xX\{\mu_{x}\}_{x\in\mathrm{X}} be a ωΓ\omega_{\Gamma}–dimensional Γ\Gamma–equivariant conformal density on hX\partial_{h}{\mathrm{X}}. This is pushed forward to a ωΓ\omega_{\Gamma}–dimensional Γ\Gamma–quasi-equivariant quasi-conformal density, denoted by {[μx]}xX\{[\mu_{x}]\}_{x\in\mathrm{X}}, on [ΛMyr(Γo)][\Lambda^{\mathrm{Myr}}{(\Gamma o)}].

Lemma 2.37 ([72, Lemma 8.4]).

Let {[μx]}xX\{[\mu_{x}]\}_{x\in\mathrm{X}} and {[μx]}xX\{[\mu_{x}^{\prime}]\}_{x\in\mathrm{X}} be two ωΓ\omega_{\Gamma}–dimensional Γ\Gamma–quasi-equivariant quasi-conformal densities on the quotient [ΛMyr(Γo)][\Lambda^{\mathrm{Myr}}{(\Gamma o)}]. Then the Radon-Nikodym derivative d[μo]/d[μo]d[\mu_{o}]/d[\mu_{o}^{\prime}] is bounded from above and below by a constant depending only on λ\lambda in Definition 2.24.

2.8. Regularly contracting limit points

In this subsection, we first introduce a class of statistically convex-cocompact actions in [71] as a generalization of convex-cocompact actions, encompassing Examples in 1.2. This notion was independently introduced by Schapira-Tapie [63] as strongly positively recurrent manifold in a dynamical context.

Given constants 0M1M20\leq M_{1}\leq M_{2}, let 𝒪M1,M2\mathcal{O}_{M_{1},M_{2}} be the set of elements gΓg\in\Gamma such that there exists some geodesic γ\gamma between NM2(o)N_{M_{2}}(o) and NM2(go)N_{M_{2}}(go) with the property that the interior of γ\gamma lies outside NM1(Γo)N_{M_{1}}(\Gamma o).

Definition 2.38 (SCC Action).

If there exist positive constants M1,M2>0M_{1},M_{2}>0 such that ω𝒪M1,M2<ωΓ<\omega_{\mathcal{O}_{M_{1},M_{2}}}<\omega_{\Gamma}<\infty, then the proper action of Γ\Gamma on X\mathrm{X} is called statistically convex-cocompact (SCC).

Remark 2.39.

The motivation for defining the set 𝒪M1,M2\mathcal{O}_{M_{1},M_{2}} comes from the action of the fundamental group of a finite volume negatively curved Hadamard manifold on its universal cover. In that situation, it is rather easy to see that for appropriate constants M1,M2>0M_{1},M_{2}>0, the set 𝒪M1,M2\mathcal{O}_{M_{1},M_{2}} coincides with the union of the orbits of cusp subgroups up to a finite Hausdorff distance. The assumption in SCC actions is called the parabolic gap condition by Dal’bo, Otal and Peigné in [25]. The growth rate ω𝒪M1,M2\omega_{\mathcal{O}_{M_{1},M_{2}}} is called complementary growth exponent in [5] and entropy at infinity in [63].

SCC actions have purely exponential growth, and thus are of divergence type. Therefore, we have the unique conformal density on a convergence boundary by Lemma 2.37.

Lemma 2.40.

Suppose that ΓX\Gamma\curvearrowright\mathrm{X} is a non-elementary SCC action with contracting elements. Then Γ\Gamma has purely exponential growth: for any n1n\geq 1, NΓ(o,n)eωΓn\sharp N_{\Gamma}(o,n)\asymp\mathrm{e}^{\omega_{\Gamma}n}.

For the remainder of this subsection, assume that ΓX\Gamma\curvearrowright\mathrm{X} is an SCC action with a contracting element. Fix a set FF of contracting elements in Γ\Gamma.

We now recall another specific class of conical limit points, introduced in [57], called regularly contracting limit points, which have been shown to be generic for PS measures there. This notion is modeled on the following purely metric notion of regularly contracting geodesics introduced earlier in [32]. For any ratio θ[0,1]\theta\in[0,1], a θ\theta–interval of a geodesic segment γ\gamma means a connected subsegment of γ\gamma with length θ(γ)\theta\ell(\gamma).

Definition 2.41.

Fix constants θ,r,C,L>0\theta,r,C,L>0. We say that a geodesic γ\gamma is (r,C,L)(r,C,L)–contracting at θ\theta–frequency if every θ\theta–interval of γ\gamma contains a segment of length LL that is rr–close to a CC–contracting geodesic.

A geodesic ray γ\gamma is (r,C,L)(r,C,L)–contracting at θ\theta–frequency if any sufficiently long initial segment of γ\gamma (i.e. γ[0,t]\gamma[0,t] for t0t\gg 0) is (r,C,L)(r,C,L)–contracting at θ\theta–frequency.

Furthermore, γ\gamma is frequently (r,C,L)(r,C,L)–contracting if it is (r,C,L)(r,C,L)–contracting at θ\theta–frequency for any θ(0,1)\theta\in(0,1).

The above notion is not used in this paper, but motivates the following analogous notion involving a proper action ΓX\Gamma\curvearrowright\mathrm{X}.

Definition 2.42.

Fix θ(0,1]\theta\in(0,1] and r>0r>0 and fΓf\in\Gamma. We say that a geodesic γ\gamma contains (r,f)(r,f)–barriers at θ\theta–frequency if for every θ\theta–segment of γ\gamma has (r,f)(r,f)–barriers.

An element gGg\in G has (r,f)(r,f)–barriers at θ\theta–frequency if there exists a geodesic γ\gamma between B(o,M)B(o,M) and B(go,M)B(go,M) such that γ\gamma has (r,f)(r,f)–barriers at θ\theta–frequency.

Let 𝒲(θ,r,F)\mathcal{W}(\theta,r,F) denote the set of elements in Γ\Gamma having no (r,f)(r,f)–barriers at θ\theta–frequency for some fFf\in F. That is, an element gg of Γ\Gamma belongs to 𝒲(θ,r,F)\mathcal{W}(\theta,r,F) if and only if any geodesic between B(o,M)B(o,M) and B(go,M)B(go,M) contains a θ\theta–interval that has no (r,f)(r,f)–barriers.

Lemma 2.43.

[57, Lemma 4.7] Fix θ(0,1],r>M\theta\in(0,1],r>M. Then 𝒲(θ,r,F)\mathcal{W}(\theta,r,F) is growth tight.

Set V:=Γ𝒲(θ,r,F)V:=\Gamma\setminus\mathcal{W}(\theta,r,F). Let 𝒞(θ,r,F)\mathcal{FC}(\theta,r,F) denote the set of limit points ξΛ(Γo)\xi\in\Lambda(\Gamma o) that is contained in infinitely many shadows at elements in VV. More precisely, 𝒞(θ,r,F)\mathcal{FC}({\theta,r,F}) is the limit superior of the following sequence, as nn\to\infty,

vAΓ(o,n,Δ)VΠoF(v,r)\bigcup\limits_{v\in A_{\Gamma}(o,n,\Delta)\cap{V}}\Pi_{o}^{F}(v,r)

where AΓ(o,n,Δ)A_{\Gamma}(o,n,\Delta) is defined in (16). Hence,

Λn:=kn(vAΓ(o,k,Δ)VΠoF(v,r))𝒞(θ,r,F)as n.\Lambda_{n}:=\bigcup\limits_{k\geq n}\Bigg{(}\bigcup\limits_{v\in A_{\Gamma}(o,k,\Delta)\cap{V}}\Pi_{o}^{F}(v,r)\Bigg{)}\searrow\mathcal{FC}(\theta,r,F)\quad\textrm{as $n\rightarrow\infty$}.

Recall Fn={fn:fF}F^{n}=\{f^{n}:f\in F\} for an integer n1n\geq 1. Define the set of regularly contracting points as follows

Λr,Freg(Γo):=nθ(0,1]𝒞(θ,r,Fn)\Lambda^{\textrm{reg}}_{\textrm{r,F}}{(\Gamma o)}:=\bigcap_{n\in\mathbb{N}}\bigcap_{\theta\in(0,1]\cap\mathbb{Q}}\mathcal{FC}(\theta,r,F^{n})

We could take a further countable intersection over all possible FF of three independent contracting elements. In general, this is a proper subset of Mryberg set in [72].

It is proved in [57, Lemma 4.12] that the set of (r,C,L)(r,C,L)–regularly contracting rays lies in 𝒞(θ,r,F)\mathcal{FC}(\theta,r,F) for certain FF. The following result thus implies [57, Theorem A], saying that (r,C,L)(r,C,L)–regularly contracting rays is μo\mu_{o}–full.

Proposition 2.44.

[57] Assume that ΓX\Gamma\curvearrowright\mathrm{X} is an SCC action with contracting elements. Then the regularly contracting limit set [Λr,Freg(Γo)][\Lambda^{\textrm{reg}}_{\textrm{r,F}}{(\Gamma o)}] is a μo\mu_{o}–full subset of [Λr,Fcon(Γo)][\Lambda^{\textrm{con}}_{\textrm{r,F}}{(\Gamma o)}].

Proof.

Recall that [Λr,Fcon(Γo)][\Lambda^{\textrm{con}}_{\textrm{r,F}}{(\Gamma o)}] is the limit super of {ΠoF(go,r):gΓ}\{\Pi_{o}^{F}(go,r):g\in\Gamma\}, and it is μo\mu_{o}–full by Lemma 2.35. The set W=𝒲(θ,r,F)W=\mathcal{W}(\theta,r,F) is growth tight by Lemma 2.43, so the limit super Λ\Lambda of {ΠoF(go,r):gW}\{\Pi_{o}^{F}(go,r):g\in W\} is μo\mu_{o}–null by Lemma 2.34. Noting V=ΓWV=\Gamma\setminus W as in the above definition, [Λr,Freg(Γo)][\Lambda^{\textrm{reg}}_{\textrm{r,F}}{(\Gamma o)}] is contained in the complementary subset of Λ\Lambda in [Λr,Fcon(Γo)][\Lambda^{\textrm{con}}_{\textrm{r,F}}{(\Gamma o)}]. Hence, μo([Λr,Freg(Γo)])=1\mu_{o}([\Lambda^{\textrm{reg}}_{\textrm{r,F}}{(\Gamma o)}])=1 is proved. ∎

Remark 2.45.

If X\mathrm{X} is a negatively curved Riemannian manifold with finite BMS measure on geodesic flow, then the PS measure is supported on the frequently contracting limit points. The same is true for certain CAT(0) groups with rank-1 elements and mapping class groups in [32, Theorems 5.1 & 7.1]. In these settings, finiteness of the BMS measure is equivalent to having purely exponential growth ([61]).

In the current coarse setting, we expect that the SCC action assumption could be replaced in Proposition 2.44 with purely exponential growth.

2.9. Hopf decomposition

In this subsection, we consider a (possibly non-discrete) countable subgroup H<Isom(X)H<\mathrm{Isom}(\mathrm{X}). In particular, HXH\curvearrowright\mathrm{X} is not necessarily a proper action with discrete orbits.

Assume HH admits a measure class preserving action on (X,m)(\partial{\mathrm{X}},m). We say that the action H(X,m)H\curvearrowright(\partial{\mathrm{X}},m) is (infinitely) conservative, if for any AXA\subseteq\partial{\mathrm{X}} with m(A)>0m(A)>0, there are (infinitely many) hH{1}h\in H\setminus\{1\} so that hAA.hA\cap A\neq\emptyset. It is called dissipative if there is a measurable wandering set AA: gAA=gA\cap A=\emptyset for each 1hH1\neq h\in H. If in addition, hHhA=X\cup_{h\in H}hA=\partial{\mathrm{X}}, then it is completely dissipative.

Let 𝐂X\mathbf{C}\subseteq\partial{\mathrm{X}} be the union of purely atomic ergodic components, and 𝐃=X𝐂\mathbf{D}=\partial{\mathrm{X}}\setminus\mathbf{C}. Denote 𝐃1\mathbf{D}_{1} (resp. 𝐃>1\mathbf{D}_{>1}) the subset of 𝐃\mathbf{D} with trivial (resp. nontrivial) stabilizers, 𝐃\mathbf{D}_{\infty} the subset of 𝐃\mathbf{D} with infinite stabilizers. The partition 𝐂𝐨𝐧𝐬=𝐂𝐃>1\mathbf{Cons}=\mathbf{C}\cup\mathbf{D}_{>1} and 𝐃𝐢𝐬𝐬=𝐃1\mathbf{Diss}=\mathbf{D}_{1} (mod 0) gives the Hopf decomposition H(X,m)H\curvearrowright(\partial{\mathrm{X}},m) as the disjoint union of conservative and completely dissipative components. If HXH\curvearrowright\mathrm{X} is proper on a hyperbolic space, 𝒟\mathcal{D}_{\infty} consists of parabolic points with infinite stabilizer. If HH is torsion-free, parabolic points are countable.

By [40, Theorem 29], the infinite conservative part 𝐂𝐨𝐧𝐬=𝐂𝐃\mathbf{Cons}_{\infty}=\mathbf{C}\cup\mathbf{D}_{\infty} coincides (mod 0) with the following set

𝐂𝐨𝐧𝐬\displaystyle\mathbf{Cons}_{\infty} ={ξX:t>0{dgm(ξ)/dm(ξ)>t}=}\displaystyle=\{\xi\in\partial{\mathrm{X}}:\exists t>0\sharp\{dg_{\star}m(\xi)/dm(\xi)>t\}=\infty\}
={ξX:gHdgm(ξ)/dm(ξ)=}\displaystyle=\{\xi\in\partial{\mathrm{X}}:\sum_{g\in H}dg_{\star}m(\xi)/dm(\xi)=\infty\}

We now consider the Γ\Gamma–equivariant ωΓ\omega_{\Gamma}–dimensional conformal density {μx:xX}\{\mu_{x}:x\in\mathrm{X}\} on X\partial{\mathrm{X}}. We emphasize that HH is not necessarily contained in Γ\Gamma, but is assumed to preserve the measure class of m:=μom:=\mu_{o}, so that the following holds for gμo=μgog_{\star}\mu_{o}=\mu_{go}:

dgμodμo(ξ)eωΓBξ(go,o)\frac{dg_{\star}\mu_{o}}{d\mu_{o}}(\xi)\asymp\mathrm{e}^{-\omega_{\Gamma}B_{\xi}(go,o)}

Recall that μo\mu_{o} is supported on [Λ(Γo)][\Lambda(\Gamma o)].

We do not assume HH to be a discrete group, so HoHo may not be a discrete subset in X\mathrm{X}. We still denote by Λ(Ho)\Lambda(Ho) the accumulation points of the orbit HoHo in the convergence boundary X\partial{\mathrm{X}}. It depends on the basepoint oo, but [Λ(Ho)][\Lambda(Ho)] does not, because of Assumption (B) in Definition 2.13.

Let 𝒞hor\mathcal{C}^{\mathrm{hor}} be defined in Assumption D as the set of points ξ𝒞\xi\in\mathcal{C} at which Busemann cocycles satisfy some weak form of continuity. According to the above discussion, 𝐂𝐨𝐧𝐬\mathbf{Cons}_{\infty} is exactly the so-called big horospheric limit point defined as follows.

Definition 2.46.

A point ξ[Λ(Ho)]𝒞hor\xi\in[\Lambda(Ho)]\cap\mathcal{C}^{\mathrm{hor}} is called big horospheric limit point if there exists hnHh_{n}\in H such that B[ξ](hno,o)LB_{[\xi]}(h_{n}o,o)\geq L for some LL\in\mathbb{R}. If, in addition, B[ξ](hno,o)LB_{[\xi]}(h_{n}o,o)\geq L holds for any given LL\in\mathbb{R}, then ξ\xi is called small horospheric limit point. Denote by ΛHor(Ho)\Lambda^{\mathrm{Hor}}{(Ho)} (resp. Λhor(Ho)\Lambda^{\mathrm{hor}}{(Ho)}) the set of the big (resp. small) horospheric limit points.

Remark 2.47.

By definition, a big or small horospheric limit point is a property of a [][\cdot]–class. Thus, ΛHor(Ho)=[ΛHor(Ho)]\Lambda^{\mathrm{Hor}}{(Ho)}=[\Lambda^{\mathrm{Hor}}{(Ho)}] is [][\cdot]–saturated. In terms of horoballs (2.22), ξ\xi is a big (resp. small) horospheric limit point if some (resp. any) horoball centered at [ξ][\xi] contains infinitely many hnoh_{n}o with hnoξh_{n}o\to\xi.

Let X>1\partial{\mathrm{X}}_{>1} (resp. X1\partial{\mathrm{X}}_{1}) denote the set of boundary points with nontrivial (resp. trivial) stabilizers in HH. Summarizing the above discussion, we have the Hopf decomposition for conformal measures.

Lemma 2.48.

Let HH act properly on X\mathrm{X} and preserve the measure class of μo\mu_{o}. Then ΛHor(Ho)\Lambda^{\mathrm{Hor}}{(Ho)} is the infinite conservative part, 𝐂𝐨𝐧𝐬=ΛHor(Ho)X>1\mathbf{Cons}=\Lambda^{\mathrm{Hor}}{(Ho)}\cup\partial{\mathrm{X}}_{>1} and 𝐃𝐢𝐬𝐬=X𝐂𝐨𝐧𝐬=X1ΛHor(Ho)\mathbf{Diss}=\partial{\mathrm{X}}\setminus\mathbf{Cons}=\partial{\mathrm{X}}_{1}\setminus\Lambda^{\mathrm{Hor}}{(Ho)}.

It is wide open whether the big horospheric limit set differs from the small one only in a negligible set (cf. Question 1.18). This has been confirmed for Kleinian groups [65] for spherical measures, free subgroups [35] for visual measures and normal subgroups of divergent type actions [27] for general conformal measures without atoms. Here is an example of atomic conformal measures in which the large and small horospheric sets differ by a positive measure set.

Example 2.49.

Let Γ\Gamma be a non-uniform lattice in Isom(m)\mathrm{Isom}(\mathbb{H}^{m}) with ωΓ=m11\omega_{\Gamma}=m-1\geq 1. We can put a conformal measure on the boundary at infinity, supported on the set 𝐏\mathbf{P} of countably many parabolic fixed points. Indeed, fix any ω>(m1)\omega>(m-1) and μo(ξ)=c>0\mu_{o}(\xi)=c>0 for given ξ𝐏\xi\in\mathbf{P}. The value of cc will be adjusted below. The other points η=gξ\eta=g\xi in Γξ\Gamma\xi is then determined by

μo(η)=ceωBξ(g1o,o)\mu_{o}(\eta)=c\cdot\mathrm{e}^{-\omega B_{\xi}(g^{-1}o,o)}

where η=gξ\eta=g\xi which does not depend on gΓg\in\Gamma (for Bξ(,)B_{\xi}(\cdot,\cdot) is invariant under the stabilizer Γξ\Gamma_{\xi}). Fix gηΓg_{\eta}\in\Gamma so that gηξ=ηg_{\eta}\xi=\eta. Observe that μo(Γξ)\mu_{o}(\Gamma\xi) can realize any value in (0,1](0,1], by adjusting cc with the following fact for given ω>ωΓ\omega>\omega_{\Gamma},

ηΓξeωBη(o,gηo)<\sum_{\eta\in\Gamma\xi}\mathrm{e}^{-\omega B_{\eta}(o,g_{\eta}o)}<\infty

which, in turn, follows from purely exponential growth of double cosets of the parabolic subgroup Γξ\Gamma_{\xi} (e.g. [36])

{g(ξ):d((ξ),g(ξ))n}eωΓn\sharp\{g\mathcal{HB}(\xi):d(\mathcal{HB}(\xi),g\mathcal{HB}(\xi))\leq n\}\asymp\mathrm{e}^{\omega_{\Gamma}n}

As 𝐏/Γ\sharp\mathbf{P}/\Gamma is finite, we are able to achieve that μo(𝐏)=1\mu_{o}(\mathbf{P})=1.

The set of parabolic points is the infinite conservative part. If the number of cusps is at least 2, the action is not ergodic. Moreover, parabolic points are the big horospheric limit set (mod 0), but not small horospheric limit points. So μo(ΛHor(Ho))=1\mu_{o}(\Lambda^{\mathrm{Hor}}{(Ho)})=1 and μo(Λhor(Ho))=0\mu_{o}(\Lambda^{\mathrm{hor}}{(Ho)})=0.

We shall prove that μo(ΛHor(Ho)Λhor(Ho))=0\mu_{o}(\Lambda^{\mathrm{Hor}}{(Ho)}\setminus\Lambda^{\mathrm{hor}}{(Ho)})=0 for subgroups in a hyperbolic group, where μo\mu_{o} is the Patterson-Sullivan measure for Γ\Gamma.

At last, we collect the notations of various limit sets studied in this paper:

  • Λ(Γo)\Lambda(\Gamma o) is the limit set of Γo\Gamma o, while [Λ(Γo)][\Lambda(\Gamma o)] is the [][\cdot]–classes over Λ(Γo)\Lambda(\Gamma o).

  • Λr,Fcon(Γo)\Lambda^{\textrm{con}}_{\textrm{r,F}}{(\Gamma o)} is the set of (r,F)(r,F)–conical points, and Λcon(Γo)\Lambda^{\textrm{con}}{(\Gamma o)} denotes the set of usual conical points.

  • ΛHor(Γo)\Lambda^{\mathrm{Hor}}(\Gamma o) (resp. Λhor(Γo)\Lambda^{\mathrm{hor}}(\Gamma o)) is the big/small horospheric limit set.

  • ΛMyr(Γo)\Lambda^{\mathrm{Myr}}{(\Gamma o)} is the Myrberg limit set, and Λr,Freg(Γo)\Lambda^{\textrm{reg}}_{\textrm{r,F}}{(\Gamma o)} is the regularly contracting limit set.

They are related as

Λr,Freg(Γo)ΛMyr(Γo)Λr,Fcon(Γo)Λcon(Γo)Λhor(Γo)ΛHor(Γo)Λ(Γo)\Lambda^{\textrm{reg}}_{\textrm{r,F}}{(\Gamma o)}\subseteq\Lambda^{\mathrm{Myr}}{(\Gamma o)}\subseteq\Lambda^{\textrm{con}}_{\textrm{r,F}}{(\Gamma o)}\subseteq\Lambda^{\textrm{con}}{(\Gamma o)}\subseteq\Lambda^{\mathrm{hor}}(\Gamma o)\subseteq\Lambda^{\mathrm{Hor}}(\Gamma o)\subseteq\Lambda(\Gamma o)

where each inclusion is proper in general.

Part I Hopf decomposition of confined subgroups

3. Prelude: proof of selected results in the hyperbolic setting

The goal of this section is two-fold: to illustrate some of our key tools via well-known facts in hyperbolic spaces; and present in this setting the proofs for (essentially all) theorems in §4 and §6 and a key tool used in subsequent sections. This section could be skipped without affecting the remaining ones.

Throughout this section, we assume that X\mathrm{X} is a Gromov hyperbolic space and HH, Γ\Gamma, GG are subgroups of Isom(X)\mathrm{Isom}(\mathrm{X}) with HH and Γ\Gamma being discrete. Let {μx:xX}\{\mu_{x}:x\in\mathrm{X}\} be Patterson-Sullivan measures on the Gromov boundary X\partial{\mathrm{X}} constructed from ΓX\Gamma\curvearrowright\mathrm{X}. If ΓX\Gamma\curvearrowright\mathrm{X} is of divergence type, then μo\mu_{o} is atomless, and charges on the conical limit set (Lemma 2.35), and Shadow Lemma 2.31 holds. We assume the reader is familiar with these facts (e.g. which could be found in [50]).

Since X\mathrm{X} is Gromov hyperbolic, there exists δ>0\delta>0 such that any geodesic triangle admits a δ\delta–center, that is, a point within a δ\delta–distance to each side.

3.1. Preliminary

The notion of an admissible path (Def. 2.6) is a generalization of the LL–local quasi-geodesic paths:

p0q1p1qnpnp_{0}q_{1}p_{1}\cdots q_{n}p_{n}

where any two consecutive paths give a quasi-geodesic by (BP), of length at least (pi)>L\ell(p_{i})>L (LL). It is well known that, in hyperbolic spaces, if L0L\gg 0, a local quasi-geodesic is a global quasi-geodesic.

Assume that ΓX\Gamma\curvearrowright\mathrm{X} is a proper non-elementary action, so there exist at least three loxodromic elements with pairwise disjoint fixed points. The following result, which is a special case of the Extension Lemma 2.10, is a consequence of the fact on quasi-geodesics mentioned above.

Lemma 3.1.

There exists a set FF of three loxodromic elements of Γ\Gamma and constants L,c>0L,c>0 that depend only on FF with the following property. For any g,hΓg,h\in\Gamma, there exists an element fFf\in F such that the path

γ:=[o,go](g[o,fo])(gf[o,ho])\gamma:=[o,go]\cdot(g[o,fo])\cdot(gf[o,ho])

is a cc–quasigeodesic.

This result is well known to experts in the field and, to the best of our knowledge, appeared first in [6, Lemma 3]. It was subsequently reproved or implicitly used in many works, which we do not attempt to track down here.

3.2. Completely dissipative actions

In this subsection, we further assume that X\mathrm{X} is proper and Γ\Gamma acts on X\mathrm{X} cocompactly. The goal is to prove the conclusion of Theorem 4.2 in this setting.

Lemma 3.2.

Let ξΛHor(Ho)\xi\in\Lambda^{\mathrm{Hor}}{(Ho)} be a big horospheric point. Then for any ϵ>0\epsilon>0, there exist infinitely many orbit points hnoHoh_{n}o\in Ho such that d(o,zn)>d(o,hno)(1/2ϵ)d(o,z_{n})>d(o,h_{n}o)(1/2-\epsilon), where znz_{n} is a δ\delta–center of Δ(o,ξ,hno)\Delta(o,\xi,h_{n}o) for some δ>0\delta>0.

Proof.

By the definition of a big horospheric point, there is an infinite sequence of points hnoHoξh_{n}o\in Ho\to\xi such that hnoh_{n}o lies in some horoball (ξ,L)\mathcal{HB}(\xi,L) for all n1n\geq 1 and for some LL depending on ξ\xi. Recall the definition of a horoball: (ξ,L)={zX:Bξ(z,o)L}\mathcal{HB}(\xi,L)=\{z\in\mathrm{X}:B_{\xi}(z,o)\leq L\}, so we obtain

Bz(hno,o)=d(z,hno)d(z,o)L+10δB_{z}(h_{n}o,o)=d(z,h_{n}o)-d(z,o)\leq L+10\delta

for z[o,ξ]z\in[o,\xi] with d(o,z)d(o,z) sufficiently large. By hyperbolicity, if zn[o,ξ]z_{n}\in[o,\xi] is a δ\delta–center of Δ(o,z,hno)\Delta(o,z,h_{n}o) for some δ>0\delta>0, then d(z,zn)+d(zn,hno)2δ+d(z,hno)d(z,z_{n})+d(z_{n},h_{n}o)\leq 2\delta+d(z,h_{n}o). Combining these inequalities gives

d(z,zn)+d(zn,hno)\displaystyle d(z,z_{n})+d(z_{n},h_{n}o) 12δ+L+d(z,o)\displaystyle\leq 12\delta+L+d(z,o)
12δ+L+d(o,zn)+d(zn,z)\displaystyle\leq 12\delta+L+d(o,z_{n})+d(z_{n},z)

yielding d(zn,hno)12δ+L+d(o,zn)d(z_{n},h_{n}o)\leq 12\delta+L+d(o,z_{n}). Thus, d(o,hno)d(o,zn)+d(zn,hno)12δ+L+2d(o,zn)d(o,h_{n}o)\leq d(o,z_{n})+d(z_{n},h_{n}o)\leq 12\delta+L+2d(o,z_{n}). Note however that LL depends on ξ\xi. Given any ϵ>0\epsilon>0, we can still obtain (1/2ϵ)d(o,hno)d(o,zn)(1/2-\epsilon)d(o,h_{n}o)\leq d(o,z_{n}) after dropping finitely many hnoh_{n}o. The proof is complete. ∎

Theorem 3.3 (Theorem 1.10 in hyp. setup).

If ωH<ωΓ/2\omega_{H}<\omega_{\Gamma}/2, then μo(ΛHor(Ho))=0\mu_{o}(\Lambda^{\mathrm{Hor}}{(Ho)})=0.

Proof.

Fix a small ϵ>0\epsilon>0, so ωH<ωΓ(1/2ϵ)\omega_{H}<\omega_{\Gamma}(1/2-\epsilon). Let Z^\hat{Z} be the set of points zXz\in\mathrm{X} that satisfy d(o,z)>d(o,ho)(1/2ϵ)d(o,z)>d(o,ho)(1/2-\epsilon), where zz is a δ\delta center of the triangle Δ(o,ho,ξ)\Delta(o,ho,\xi) for some hHh\in H and some ξΛHor(Ho)\xi\in\Lambda^{\mathrm{Hor}}{(Ho)}.

Choose a sufficiently large constant r>18δr>18\delta such that the Shadow Lemma 2.31 applies. If ZZ^Z\subseteq\hat{Z} is a maximal rr–separated net in Z^\hat{Z}, then by Lemma 3.2, ΛHor(Ho)\Lambda^{\mathrm{Hor}}{(Ho)} is contained in the limit superior lim supzZΠo(z,2r)\limsup_{z\in Z}\Pi_{o}(z,2r) of shadows. As the action ΓX\Gamma\curvearrowright\mathrm{X} is cocompact, we may assume ZΓoZ\subseteq\Gamma o upon increasing rr again.

Note that, for each hHh\in H, the corresponding zz in the first paragraph is lying on the δ\delta–neighborhood of [o,ho][o,ho]. This neighborhood contains at most Cd(o,ho)Cd(o,ho) points in the rr–separated set ZZ where C>0C>0 is a universal constant.

By the Shadow Lemma 2.31 for the Γ\Gamma–conformal density, we compute

zZμo(Πo(z,2r))\displaystyle\sum_{z\in Z}\mu_{o}(\Pi_{o}(z,2r)) hHd(o,ho)eωΓd(o,ho)(1/2ϵ)\displaystyle\prec\sum_{h\in H}d(o,ho)\cdot\mathrm{e}^{-\omega_{\Gamma}d(o,ho)(1/2-\epsilon)}
hHeωd(o,ho)<\displaystyle\prec\sum_{h\in H}\mathrm{e}^{-\omega d(o,ho)}<\infty

which is finite by definition of ωH\omega_{H} and ωH<ωΓ(1/2ϵ)\omega_{H}<\omega_{\Gamma}(1/2-\epsilon). The Borel-Cantelli lemma implies that lim supzZΠo(z,2r))\limsup_{z\in Z}\Pi_{o}(z,2r)) is μo\mu_{o}–null. Hence, μo(ΛHor(Ho))=0\mu_{o}(\Lambda^{\mathrm{Hor}}{(Ho)})=0, and the theorem is proved. ∎

3.3. Conservative actions for confined subgroups

This subsection is an abridged version of Sections §5 and §6 in hyperbolic setup. In this restricted setting, we will give two proofs that the horospheric limit set of a confined subgroup HH has full Patterson-Sullivan measure. One of the proofs will work whenever HH has a compact confining set, whereas the other requires a finite confining set. However, the latter proof is more readily adapted to the setting of general actions on metric spaces with contracting elements, which we will consider in Section 6. Therefore, we find it instructive to include both proofs.

We first state a key tool that allows us to use a finite confining set geometrically in hyperbolic spaces.

Lemma 3.4 (\ll Lemma 5.7).

Let PIsom(X)P\subseteq\mathrm{Isom}(\mathrm{X}) be a finite set, so that no element in PP fixes pointwise the limit set Λ(Γo)\Lambda(\Gamma o). Then there exists a finite set FΓF\subseteq\Gamma of loxodromic elements and a constant D>0D>0 with the following property.

For any g,hGg,h\in G, one can find fFf\in F so that for any pPp\in P, the word (g,f,p,f1,h)(g,f,p,f^{-1},h) labels a quasi-geodesic:

|d(o,gfpf1ho)d(o,go)d(o,ho)|D|d(o,gfpf^{-1}ho)-d(o,go)-d(o,ho)|\leq D
Proof.

Let pPp\in P. By assumption, as all fixed points of loxodromic elements are dense in Λ(Γo)\Lambda(\Gamma o), pp moves the two fixed points of some ff. If qPq\in P moves g±g^{\pm} for another gg, then p,qp,q must move the two fixed points of fngnf^{n}g^{n}, which tends to f+f^{+} and gg^{-} as nn\to\infty. As PP is a finite set, a common ff could be chosen for each pPp\in P: pf±f±pf^{\pm}\cap f^{\pm}\neq\emptyset. Equivalently, pAx(f)p\mathrm{Ax}(f) has bounded projection with Ax(f)\mathrm{Ax}(f) for each fP1f\in P\setminus 1.

With a bit more effort, we produce F={f1,f2,f3}F=\{f_{1},f_{2},f_{3}\} where pPp\in P moves the two endpoints of each fFf\in F, and the set of axis {pAx(f):pP,fF}\{p\mathrm{Ax}(f):p\in P,f\in F\} has bounded projection. This implies that each gΓg\in\Gamma has bounded projection to at least two of Ax(f1),Ax(f2),Ax(f3)\mathrm{Ax}(f_{1}),\mathrm{Ax}(f_{2}),\mathrm{Ax}(f_{3}). Consequently, for any g,hΓg,h\in\Gamma, there exists a common fFf\in F so that

max{πAx(f)([o,go]),πAx(f)([o,ho])}τ\max\{\pi_{\mathrm{Ax}(f)}([o,go]),\pi_{\mathrm{Ax}(f)}([o,ho])\}\leq\tau

Set L=min{d(o,fo):fF}L=\min\{d(o,fo):f\in F\}. The word (g,f,p,f1,h)(g,f,p,f^{-1},h) labels an LL–local cc–quasi-geodesic path, denoted by γ\gamma, where cc depends on τ\tau. This concludes the proof. ∎

The following is an immediate consequence of Lemma 3.4 applied to (g,g1)(g,g^{-1}), and pPp\in P is chosen for gfgf according to the definition of confined subgroups.

Lemma 3.5 (\ll Lemma 5.8).

Under the assumption of Lemma 3.4, for any gGg\in G, there exist fFf\in F and pPp\in P such that gfpf1g1gfpf^{-1}g^{-1} lies in HH and

|d(o,gfpf1g1o)2d(o,go)|D|d(o,gfpf^{-1}g^{-1}o)-2d(o,go)|\leq D

Recall that a point ξΛcon(Γo)\xi\in\Lambda^{\textrm{con}}{(\Gamma o)} is a conical point if there exist a sequence of gnΓg_{n}\in\Gamma so that gnog_{n}o lies in a rr–neighborhood of a geodesic ray γ=[o,ξ]\gamma=[o,\xi].

The next two results prove Theorem 1.11 in hyperbolic setup under the condition (i) and (ii) accordingly. The following does not use Lemmas 3.5 and 3.4.

Theorem 3.6 (Theorem 1.11(i) in hyp. setup).

Assume that H<Isom(X)H<\mathrm{Isom}(\mathrm{X}) is a torsion-free discrete subgroup confined by Γ\Gamma with a compact confining subset PP. Then the big horospheric limit set ΛHor(Ho)\Lambda^{\mathrm{Hor}}{(Ho)} contains all but countably many points of Λcon(Γo)\Lambda^{\textrm{con}}{(\Gamma o)}.

Proof.

We denote by Λ0Λcon(Γo)\Lambda_{0}\in\Lambda^{\textrm{con}}{(\Gamma o)} the countable union of fixed points of all loxodromic and parabolic elements hHh\in H. We shall prove that any conical point ξΛcon(Γo)Λ0\xi\in\Lambda^{\textrm{con}}{(\Gamma o)}\setminus\Lambda_{0} is contained in ΛHor(Ho)\Lambda^{\mathrm{Hor}}{(Ho)}. By definition, there exists a sequence of gnΓg_{n}\in\Gamma so that gnog_{n}o lies in a rr–neighborhood of a geodesic ray γ=[o,ξ]\gamma=[o,\xi].

Set D:=max{d(o,po):pP}<D:=\max\{d(o,po):p\in P\}<\infty. The confined subgroup HH implies the existence of pnPp_{n}\in P so that hn:=gnpngn1Hh_{n}:=g_{n}p_{n}g_{n}^{-1}\in H. Now, let z[o,ξ]z\in[o,\xi] so that d(gno,z)rd(g_{n}o,z)\leq r. Thus, d(hno,z)r+D+d(o,gno)2r+D+d(o,z)d(h_{n}o,z)\leq r+D+d(o,g_{n}o)\leq 2r+D+d(o,z). Direct computation shows that hnoHoh_{n}o\in Ho lies in the horoball (ξ,o,2r+D+ϵ)\mathcal{HB}(\xi,o,2r+D+\epsilon) (or see Lemma 6.1 for this general fact).

A horoball (ξ)\mathcal{HB}(\xi) only accumulates at the center ξ\xi. As the orbit HoXHo\subset X is discrete, it suffices to prove that {hno:n1}\{h_{n}o:n\geq 1\} is an infinite subset. Arguing by contradiction, assume now that {hno:n1}\{h_{n}o:n\geq 1\} is a finite set. By taking a subsequence, we may assume that h:=hn=hmh:=h_{n}=h_{m} for any n,m1n,m\geq 1.

By assumption, HH is torsion-free, so pnp_{n} must be of infinite order. According to the classification of isometries, pnp_{n} is either hyperbolic or parabolic, so is hnh_{n}.

As d(hgno,gno)=d(o,pno)<Dd(hg_{n}o,g_{n}o)=d(o,p_{n}o)<D, the convergence gnoξg_{n}o\to\xi implies that hgnoξhg_{n}o\to\xi and then hh fixes ξ\xi. So ξΛ0\xi\in\Lambda_{0} gives a contradiction, which completes the proof. ∎

We now demonstrate how to use Lemma 3.5 to deal with finite confining subset PP.

Theorem 3.7 (Theorem 1.11(ii) in hyp. setup).

Assume that H<Isom(X)H<\mathrm{Isom}(\mathrm{X}) is a discrete subgroup confined by Γ\Gamma with a finite confining subset PP. Assume that no element in PP fixes pointwise Λ(Γo)\Lambda(\Gamma o). Then the big horospheric limit set ΛHor(Ho)\Lambda^{\mathrm{Hor}}{(Ho)} contains Λcon(Γo)\Lambda^{\textrm{con}}{(\Gamma o)}.

Proof.

Given ξΛcon(Γo)\xi\in\Lambda^{\textrm{con}}{(\Gamma o)}, there exists a sequence of gnΓg_{n}\in\Gamma such that d(gno,[o,ξ])rd(g_{n}o,[o,\xi])\leq r. By Lemma 3.5, we can choose fnFf_{n}\in F and pnPp_{n}\in P so that hn:=gnfnpnfn1gn1Hh_{n}:=g_{n}f_{n}p_{n}f_{n}^{-1}g_{n}^{-1}\in H labels a quasi-geodesic path.

If d(gno,gmo)0d(g_{n}o,g_{m}o)\gg 0 for any nmn\neq m, then hnohmoh_{n}o\neq h_{m}o follows by Morse Lemma. Thus {hno:n1}\{h_{n}o:n\geq 1\} is an infinite subset. Setting

D=maxfF{d(o,fo)}+maxpP{d(o,po)}D=\max_{f\in F}\{d(o,fo)\}+\max_{p\in P}\{d(o,po)\}

we argue exactly as in the proof of Theorem 6.5 and obtain that hnoHoh_{n}o\in Ho lies in the horoball (ξ,2r+2D+ϵ)\mathcal{HB}(\xi,2r+2D+\epsilon). Hence, {hno:n1}\{h_{n}o:n\geq 1\} converges to ξ\xi, so ξ\xi is a big horospheric limit point. ∎

4. Completely dissipative actions

Our goal of this section is to prove Theorem 1.10 under the following setup.

  • The auxiliary proper action ΓX\Gamma\curvearrowright\mathrm{X} is assumed to be SCC, in particular it is of divergence type.

  • Let X\partial{\mathrm{X}} be a convergence boundary for X\mathrm{X} and {μx:xX}\{\mu_{x}:x\in\mathrm{X}\} be the unique quasi-conformal, Γ\Gamma–equivariant density of dimension ωΓ\omega_{\Gamma} on X\partial{\mathrm{X}}.

  • Let H<Isom(X)H<\mathrm{Isom}(\mathrm{X}) be a discrete subgroup, which is confined by Γ\Gamma with a compact confining subset PP in Isom(X)\mathrm{Isom}(\mathrm{X}).

We emphasize that HH is not necessarily contained in Γ\Gamma, but preserves the measure class of μo\mu_{o}. This is motivated by the following example. If X\mathrm{X} is the rank-1 symmetric space equipped with the Lebesgue measure μo\mu_{o} on the visual boundary, any subgroup H<Isom(X)H<\mathrm{Isom}(\mathrm{X}) preserves the measure class of μ0\mu_{0}. More generally, we have.

Lemma 4.1.

Suppose that X\mathrm{X} is a hyperbolic space on which Γ<Isom(X)\Gamma<\mathrm{Isom}(\mathrm{X}) acts properly and co-compactly. Let {μx:xX}\{\mu_{x}:x\in\mathrm{X}\} be the unique Patterson-Sullivan measure class of dimension ωΓ\omega_{\Gamma} on X\partial{\mathrm{X}}. Then Isom(X)\mathrm{Isom}(\mathrm{X}) preserves the measure class of μo\mu_{o}.

Proof.

It is known that μo\mu_{o} coincides with the Hausdorff measure of X\partial{\mathrm{X}} with respect to the visual metric. Namely, if ρϵ\rho_{\epsilon} is the visual metric for parameter ϵ\epsilon, then μo(A)ρϵd(A)\mu_{o}(A)\asymp\mathcal{H}^{d}_{\rho_{\epsilon}}(A) where d=ωΓ/ϵd=\omega_{\Gamma}/\epsilon is the Hausdorff dimension. As an isometry on hyperbolic spaces induces a bi-Lipschitz map on boundary, it preserves the Hausdorff measure class. The proof is complete.∎

If HH contains no torsion, the set 𝒟<\mathcal{D}_{<\infty} of points with nontrivial stabilizer is empty. Thus the conservative component is exactly the infinite conservative part, which by Lemma 2.48 coincides with the big horospherical limit set. Hence, we shall prove μo(ΛHor(Ho))=0\mu_{o}(\Lambda^{\mathrm{Hor}}{(Ho)})=0, provided that ωH<ωΓ/2\omega_{H}<\omega_{\Gamma}/2.

Our argument is essentially a geometric interpretation of the corresponding one in [35], where the same conclusion is proven for free groups. The proof in [35] is more combinatorial: it crucially uses the so-called Nielsen-Schreier system given by a spanning tree in the Schreier graph.

Recall that 𝒞ϵhor\mathcal{C}_{\epsilon}^{\mathrm{hor}} is a subset in 𝒞\mathcal{C} (Assumption D) on which the Busemann cocycles converge up to an additive error ϵ\epsilon in (8).

Theorem 4.2.

If ωH<ωΓ/2\omega_{H}<\omega_{\Gamma}/2, then μo(ΛHor(Ho))=0\mu_{o}(\Lambda^{\mathrm{Hor}}{(Ho)})=0. In particular, if HH is torsion-free, the action of HH on (X,μo)(\partial{\mathrm{X}},\mu_{o}) is completely dissipative.

Proof.

By Proposition 2.44, μo\mu_{o} has full measure on Λr,Freg(Γo)\Lambda^{\textrm{reg}}_{\textrm{r,F}}{(\Gamma o)}, and Λr,Freg(Γo)\Lambda^{\textrm{reg}}_{\textrm{r,F}}{(\Gamma o)} is a subset of 𝒞20Chor\mathcal{C}^{\mathrm{hor}}_{20C}. By taking the intersection Λr,Freg(Γo)ΛHor(Ho)\Lambda^{\textrm{reg}}_{\textrm{r,F}}{(\Gamma o)}\cap\Lambda^{\mathrm{Hor}}{(Ho)}, we may assume in addition that every point ξΛHor(Ho)\xi\in\Lambda^{\mathrm{Hor}}{(Ho)} is a frequently contracting point. This is only the place in the proof where we use the SCC action.

For any sufficiently small θ(0,1]\theta\in(0,1], instead of Lemma 3.2, we now prove

Claim.

Let ξΛHor(Ho)\xi\in\Lambda^{\mathrm{Hor}}{(Ho)}. There exist two sequences of elements hn,tnGh_{n},t_{n}\in G such that

(22) d(o,tno)>(1/2θ)d(o,hno)\displaystyle d(o,t_{n}o)>(1/2-\theta)d(o,h_{n}o)
(23) d(tno,[o,ξ])r\displaystyle d(t_{n}o,[o,\xi])\leq r
(24) d(tno,[o,hno])C+Fo\displaystyle d(t_{n}o,[o,h_{n}o])\leq C+\|Fo\|
Proof of the claim.

As ξ\xi is a big horospheric point, there exists hno([ξ],L0)h_{n}o\in\mathcal{HB}([\xi],L_{0}) tending to ξ\xi, where L0L_{0} depends on ξ\xi. That is, B[ξ](hno,o)L0B_{[\xi]}(h_{n}o,o)\leq L_{0} where ξ𝒞20Chor\xi\in\mathcal{C}^{\mathrm{hor}}_{20C}. Setting M=L0+20CM=L_{0}+20C, we obtain from (8):

(25) z[o,ξ]:d(hno,z)d(o,z)+M\forall z\in[o,\xi]:\quad d(h_{n}o,z)\leq d(o,z)+M

Fix a big LML\gg M to be decided below, which also depends on ξ\xi.

Choose a point z[o,ξ]z\in[o,\xi] so that d(o,z)=d(o,hno)Ld(o,z)=d(o,h_{n}o)-L for n0n\gg 0. Consider the initial segment α:=[o,z]\alpha:=[o,z] of [o,ξ][o,\xi] and one θ\theta–interval [x,y]α[x,y]\subseteq\alpha at the middle satisfying d(o,x)=(α)/2d(o,x)=\ell(\alpha)/2 and d(y,α+)=(1/2+θ)(α)d(y,\alpha_{+})=(1/2+\theta)\ell(\alpha). By definition of ξΛr,Freg(Γo)\xi\in\Lambda^{\textrm{reg}}_{\textrm{r,F}}{(\Gamma o)}, any θ\theta–interval of α\alpha contains (r,f)(r,f)–barrier tGt\in G, so

max{d(to,[x,y]),d(tfo,[x,y])}r\max\{d(to,[x,y]),d(tfo,[x,y])\}\leq r

where fFf\in F is a contracting element. Up to enlarge rr, we assume [to,tfo]Nr([x,y]),[to,tfo]\subseteq N_{r}([x,y]), so (23) is satisfied.

xxyyzzhnoh_{n}ouuvvtototfotfoooξ\xiθ(α)\theta\ell(\alpha)α=[o,z]\alpha=[o,z]
Figure 6. Schematic configuration in the proof of Theorem 4.2

Look at the triangle with vertices (o,hno,z)(o,h_{n}o,z) for z[o,ξ]z\in[o,\xi]. See Fig. (6). By the contracting property, tAx(f)t\mathrm{Ax}(f) intersects non-trivially either NC([o,hno])N_{C}([o,h_{n}o]) or NC([z,hno])N_{C}([z,h_{n}o]). Indeed, if this is false, the projection of [o,hno][o,h_{n}o] and [z,hno][z,h_{n}o] to tAx(f)t\mathrm{Ax}(f) together has diameter at most 2C2C. Meanwhile, the corresponding initial and terminal segments of α\alpha entering and leaving NC(tAx(f))N_{C}(t\mathrm{Ax}(f)) both project to tAx(f)t\mathrm{Ax}(f) with diameter at most CC as well. Hence, we obtain αNC(tAx(f))4C\|\alpha\cap N_{C}(t\mathrm{Ax}(f))\|\leq 4C. On the other hand, noting the above αNr(tAx(f))d(o,fo)\|\alpha\cap N_{r}(t\mathrm{Ax}(f))\|\geq d(o,fo), we would get a contradiction if d(o,fo)4C+2rd(o,fo)\gg 4C+2r.

Moreover, as [to,tfo]Nr(α)tAx(f)[to,tfo]\subseteq N_{r}(\alpha)\cap t\mathrm{Ax}(f), the above argument further shows that [to,tfo][to,tfo] intersects non-trivially either NC([o,hno])N_{C}([o,h_{n}o]) or NC([z,hno])N_{C}([z,h_{n}o]).

We now claim that [to,tfo]NC([z,hno])=[to,tfo]\cap N_{C}([z,h_{n}o])=\emptyset. Indeed, if not, we have d(u,v)Cd(u,v)\leq C for some u[to,tfo]u\in[to,tfo] and v[z,hno]v\in[z,h_{n}o]. Thus,

d(v,hno)\displaystyle d(v,h_{n}o) =d(z,hno)d(z,v)\displaystyle=d(z,h_{n}o)-d(z,v)
d(z,hno)d(z,u)+C\displaystyle\leq d(z,h_{n}o)-d(z,u)+C
d(z,o)d(z,u)+C+M\displaystyle\leq d(z,o)-d(z,u)+C+M

where the last inequality uses (25). On the other hand, as d(o,hno)=d(o,z)+Ld(o,h_{n}o)=d(o,z)+L, we have

d(v,hno)\displaystyle d(v,h_{n}o) d(o,hno)d(o,v)\displaystyle\geq d(o,h_{n}o)-d(o,v)
d(o,z)d(o,u)+LC\displaystyle\geq d(o,z)-d(o,u)+L-C

If L>2C+ML>2C+M is assumed, this would give a contradiction, so the above claim is proved.

By the claim, let us choose u[to,tfo]NC([o,hno])u\in[to,tfo]\cap N_{C}([o,h_{n}o]). Thus, we have d(to,[o,hno])C+d(o,fo)d(to,[o,h_{n}o])\leq C+d(o,fo), so (24) is satisfied.

Recall that (α)=d(o,hno)L\ell(\alpha)=d(o,h_{n}o)-L and d(to,tfo)d(x,y)+2rd(to,tfo)\leq d(x,y)+2r. For u[to,tfo]u\in[to,tfo], we have

d(u,to)d(x,y)+2rθα+2rθ(d(o,hno)L)+2rd(u,to)\leq d(x,y)+2r\leq\theta\alpha+2r\leq\theta(d(o,h_{n}o)-L)+2r

and

d(o,u)d(o,x)r(α)/2r(d(o,hno)L)/2rd(o,u)\geq d(o,x)-r\geq\ell(\alpha)/2-r\geq(d(o,h_{n}o)-L)/2-r

which together give

d(o,to)d(o,u)d(u,to)(d(o,hno)L)(1/2θ)3r.d(o,to)\geq d(o,u)-d(u,to)\geq(d(o,h_{n}o)-L)(1/2-\theta)-3r.

As d(o,hno)d(o,h_{n}o)\to\infty, we may drop finitely many hnoh_{n}o to remove LL (which depends on ξ\xi though), so that d(o,to)d(o,hno)(1/22θ).d(o,to)\geq d(o,h_{n}o)(1/2-2\theta). This shows that tn:=tt_{n}:=t is the desired element with (22,23,24). ∎

Now we choose θ,ϵ>0\theta,\epsilon>0 so that ωH+ϵ<ωΓ(1/2θ)\omega_{H}+\epsilon<\omega_{\Gamma}(1/2-\theta). Let ZZ be the set of points tΓt\in\Gamma with the above property in the above Claim. Note that Λr,Freg(Γo)ΛHor(Ho)\Lambda^{\textrm{reg}}_{\textrm{r,F}}{(\Gamma o)}\cap\Lambda^{\mathrm{Hor}}{(Ho)} is contained in the limit superior of {ΠoF(to,r):tZ}\{\Pi_{o}^{F}(to,r):t\in Z\}. Note again that hh is over used by tt at most C0d(o,ho)C_{0}d(o,ho) times for a universal constant C0C_{0}, as toto in (24) lies in a fixed neighborhood of [o,ho][o,ho]. Shadow lemma hence allows to compute

tZμo(ΠoF(to,r))hHd(o,ho)eωΓ(1/2θ)d(o,ho)hHd(o,ho)e(ωH+ϵ)d(o,ho)<.\sum_{t\in Z}\mu_{o}(\Pi_{o}^{F}(to,r))\leq\sum_{h\in H}d(o,ho)\mathrm{e}^{-\omega_{\Gamma}(1/2-\theta)d(o,ho)}\prec\sum_{h\in H}d(o,ho)\mathrm{e}^{-(\omega_{H}+\epsilon)d(o,ho)}<\infty.

Hence,

μo(Λr,Freg(Γo)ΛHor(Ho))=0\mu_{o}(\Lambda^{\textrm{reg}}_{\textrm{r,F}}{(\Gamma o)}\cap\Lambda^{\mathrm{Hor}}{(Ho)})=0

by Borel-Cantelli Lemma. Thus, μo(Λr,Freg(Γo))=1\mu_{o}(\Lambda^{\textrm{reg}}_{\textrm{r,F}}{(\Gamma o)})=1 implies μo(ΛHor(Ho))=0\mu_{o}(\Lambda^{\mathrm{Hor}}{(Ho)})=0, completing the proof. ∎

5. Preliminaries on confined subgroups

From this section to §9, we study the boundary actions of confined subgroups, growth and co-growth inequalities. In this section we first recall the notion of confined subgroups and review some known facts for later reference. Then we show an enhanced version of the Extension Lemma, which will be crucial in the later sections.

5.1. Confined subgroups

Definition 5.1.

Let GG be a locally compact, metrizable, topological group and Γ<G\Gamma<G. We say that H<GH<G is confined by Γ\Gamma if there exists a compact subset PP in GG such that PP intersects non-trivially with any conjugate g1Hgg^{-1}Hg for gΓg\in\Gamma:

g1Hg(P{1}).g^{-1}Hg\cap(P\setminus\{1\})\neq\emptyset.

We refer PP as a confining subset for HH. If Γ=G\Gamma=G, we say that HH is a confined subgroup in GG.

In this paper, we are mainly interested in the case where G=Isom(X)G=\mathrm{Isom}(\mathrm{X}) or G=ΓG=\Gamma.

Note that any nontrivial normal subgroup is confined. Here are a few remarks in order.

Remark 5.2.

If PP is the confining subset, so is Pn={pn:pP}P^{n}=\{p^{n}:p\in P\} for n1n\geq 1. If PP is finite, we may further assume that PP is minimal: for any pPp\in P, there exists gGg\in G so that gpg1Hgpg^{-1}\in H.

If HH is a confined subgroup in GG, then any subgroup containing HH is also confined. Note that being confined is inherited to finite index subgroups. Indeed, if KK is a finite index subgroup of HH, then any element hHKh\in H\setminus K admits a sufficiently large power in KK, so KK is a confined subgroup in GG.

Let HH be a discrete subgroup of Isom(X)\mathrm{Isom}(\mathrm{X}). We define the injectivity radius of the action at a point xXx\in\mathrm{X} as follows:

InjH(x)=min{d(x,hx):hHHx}\textrm{Inj}_{H}(x)=\min\{d(x,hx):h\in H\setminus H_{x}\}

where Hx:={hH:hx=x}H_{x}:=\{h\in H:hx=x\} is finite by the proper action. By definition, InjH(hx)=InjH(x)\textrm{Inj}_{H}(hx)=\textrm{Inj}_{H}(x) for hHh\in H, so the injectivity radius function descends to the quotient X/H\mathrm{X}/H. When HH is torsion free, this coincides with the usual notion of injectivity radius, that is the maximal radius of the disk centered at xx in X\mathrm{X} injecting into X/H\mathrm{X}/H. It is clear that InjG(Hx)InjG(Gx)\textrm{Inj}_{G}(Hx)\geq\textrm{Inj}_{G}(Gx) for any H<GH<G.

Lemma 5.3.

Assume that Γ<Isom(X)\Gamma<\mathrm{Isom}(\mathrm{X}) acts cocompactly (and not necessarily properly) on a Γ\Gamma–invariant subset ZXZ\subseteq\mathrm{X}. Denote π:XX/H\pi:\mathrm{X}\to\mathrm{X}/H. Then the subset π(Z)\pi(Z) of X/H\mathrm{X}/H has injectivity radius bounded from above if and only if HH is confined by Γ\Gamma.

Proof.

(1). Suppose that InjH:X/H\textrm{Inj}_{H}:\mathrm{X}/H\to\mathbb{R} is bounded from above by D>0D>0 over the set π(Z)\pi(Z). That is, for any xZx\in Z, there exists hHHxh\in H\setminus H_{x} such that d(x,hx)Dd(x,hx)\leq D. Fixing a basepoint oZo\in Z, as Γ\Gamma acts cocompactly on XX, there exists a constant M>0M>0 such that ZNM(Γo)Z\subseteq N_{M}(\Gamma o). Given fΓf\in\Gamma, let xZx\in Z so that d(fo,x)Md(fo,x)\leq M, so we have d(o,f1hfo)2D+Md(o,f^{-1}hfo)\leq 2D+M. The set P:={gIsom(X):d(o,go)2D+M}P:=\{g\in\mathrm{Isom}(\mathrm{X}):d(o,go)\leq 2D+M\} is compact, so HH is confined by Γ\Gamma.

(2). Suppose that HH is confined by Γ\Gamma with a confining subset PP, we need to bound InjG(Z/H)\textrm{Inj}_{G}(Z/H). For any xZx\in Z, the cocompact action of Γ\Gamma on ZZ implies the existence of gΓg\in\Gamma such that d(go,x)Md(go,x)\leq M. As HH is confined by Γ\Gamma, there exist hHh\in H and pPp\in P such that g1hg=pg^{-1}hg=p. Thus, we have d(x,hx)d(go,hgo)+2Md(o,po)+2Md(x,hx)\leq d(go,hgo)+2M\leq d(o,po)+2M. Setting D=2M+max{d(o,po):pP}<D=2M+\max\{d(o,po):p\in P\}<\infty completes the other direction of the proof: InjH(Hx)D\textrm{Inj}_{H}(Hx)\leq D over xZx\in Z. ∎

A direct corollary is the following.

Lemma 5.4.

Assume that a group Γ<Isom(X)\Gamma<\mathrm{Isom}(\mathrm{X}) acts cocompactly on X\mathrm{X}. Then for a discrete subgroup H<G:=Isom(X)H<G:=\mathrm{Isom}(\mathrm{X}), the following are equivalent:

  1. (1)

    HH is confined by Γ\Gamma;

  2. (2)

    HH is confined in GG;

  3. (3)

    X/H\mathrm{X}/H has injectivity radius bounded from above.

5.2. Elliptic radical

According to the definition, the product of any subgroup and a non-trivial finite normal subgroup is confined. To eliminate such pathological examples, we consider the notion of elliptic radical which, roughly speaking, consists of elements that fix the boundary pointwise.

Consider a convergence boundary X\partial{\mathrm{X}} of X\mathrm{X}, where the action of Isom(X)\mathrm{Isom}(\mathrm{X}) extends to X\partial{\mathrm{X}} by homeomorphisms and preserves the partition [][\cdot] (see Definition 2.13). Assume that Γ\Gamma is a non-elementary discrete subgroup of Isom(X)\mathrm{Isom}(\mathrm{X}) with non-pinched contracting elements.

Definition 5.5.

Let G<Isom(X)G<\mathrm{Isom}(\mathrm{X}) be a group. The elliptic radical EG(Γ)E_{G}(\Gamma) consists of elements gGg\in G that induces an identity map on the limit set [Λ(Γo)][\Lambda(\Gamma o)]:

EG(Γ):={gG:g[ξ]=[ξ],[ξ][Λ(Γo)]}.E_{G}(\Gamma):=\{g\in G:g[\xi]=[\xi],\forall[\xi]\subseteq[\Lambda(\Gamma o)]\}.

Note that gEG(Γ)g\in E_{G}(\Gamma) may not fix pointwise Λ(Γo)\Lambda(\Gamma o). If G=Isom(X)G=\mathrm{Isom}(\mathrm{X}) we write E(Γ)=EG(Γ)E(\Gamma)=E_{G}(\Gamma).

Remark 5.6.

Recall that, given a contracting element fΓf\in\Gamma, EG(f)E_{G}(f) is the maximal elementary subgroup in GG that contains ff. By Lemma 2.17, if ff is non-pinched, then EG(f)E_{G}(f) is the set stabilizer of [f±][f^{\pm}] in GG.

If G=ΓG=\Gamma, EG(Γ)E_{G}(\Gamma) is the intersection of the maximal elementary subgroups EG(f)E_{G}(f) (in Lemma 2.2) for all contracting elements fGf\in G. This is the unique maximal finite normal subgroup of GG.

5.3. An enhanced extension lemma

The following result will be a crucial tool in the next sections, which could be thought of as an enhanced version of Lemma 2.10.

Lemma 5.7.

Let PIsom(X)P\subseteq\mathrm{Isom}(\mathrm{X}) be a finite set disjoint with E(Γ)E(\Gamma). Then there exists a finite set FΓF\subseteq\Gamma of contracting elements and constants τ,D>0\tau,D>0 with the following property. Set L=min{d(o,fo):fF}L=\min\{d(o,fo):f\in F\}.

For any g,hΓg,h\in\Gamma, one can find fFf\in F so that for any pPp\in P, the word (g,f,p,f1,h)(g,f,p,f^{-1},h) labels an (L,τ)(L,\tau)–admissible path with associated contracting subsets gAx(f)g\mathrm{Ax}(f) and gfpAx(f)gfp\mathrm{Ax}(f). Moreover,

|d(o,gfpf1ho)d(o,go)d(o,ho)|D|d(o,gfpf^{-1}ho)-d(o,go)-d(o,ho)|\leq D
Proof.

First of all, fixing a non-pinched contracting element fΓf\in\Gamma, we claim that for any pPp\in P, there exists gΓg\in\Gamma so that pg[f±]g[f±]pg[f^{\pm}]\neq g[f^{\pm}].

Indeed, if not, then [pgη]=[gη][pg\eta]=[g\eta] for any gΓg\in\Gamma and for any η[f±]\eta\in[f^{\pm}]. By Lemma 2.21, the set of Γ\Gamma–translates of its fixed points [f±][f^{\pm}] is dense in Λ(Γo)\Lambda(\Gamma o): [Λ(Γo)]=[Γη¯][\Lambda(\Gamma o)]=[\overline{\Gamma\eta}]. To be more precise, for any ξΛ(Γo)\xi\in\Lambda(\Gamma o), there exists gnΓg_{n}\in\Gamma such that gnηξg_{n}\eta\to\xi^{\prime} for some ξ[ξ]\xi^{\prime}\in[\xi]. By assumption, [pgnη]=[gnη][pg_{n}\eta]=[g_{n}\eta]. According to Definition 2.13(C), [][\cdot] on Γ[g±]\Gamma[g^{\pm}] forms a closed relation, this implies [pξ]=[ξ][p\xi^{\prime}]=[\xi^{\prime}]: that is, pp fixes every [][\cdot]–class in [Λ(Γo)][\Lambda(\Gamma o)], so pE(Γ)p\in E(\Gamma). This gives a contradiction. Hence, we see that pPp\in P does not stabilize g[f±]g[f^{\pm}] for any gΓg\in\Gamma.

By Lemma 2.19, if f,gf,g are non-pinched contracting elements in Γ\Gamma, then hn:=fngnh_{n}:=f^{n}g^{n} for any n0n\gg 0 is a contracting element so that the attracting fixed point [h+][h^{+}] tends to [f+][f^{+}] and the repelling fixed point [h+][h^{+}] tends to [g][g^{-}] as nn\to\infty. By the previous paragraphs, for any p,qE(Γ)p,q\notin E(\Gamma), we can find two non-pinched contracting elements f,gΓf,g\in\Gamma, so that pp does not fix [f+][f^{+}] and qq does not fix [g][g^{-}]. As [f+][f_{+}] and [pf+][pf_{+}] are disjoint closed subsets in X\partial{\mathrm{X}} by Definition 2.13(A), choose disjoint open neighbourhoods such that Lemma 2.19 shows that p,qp,q do not stabilize the fixed points [hn±][h_{n}^{\pm}] of hnh_{n} for any n0n\gg 0. We do not conclude here that p,qEΓ(hn)p,q\in E_{\Gamma}(h_{n}), as EΓ(hn)E_{\Gamma}(h_{n}) is the subgroup in Γ\Gamma (not in Isom(X)\mathrm{Isom}(\mathrm{X})) stabilizing [hn±][h_{n}^{\pm}].

As PP is a finite set, a repeated application of the previous paragraph produces an infinite set F~Γ\tilde{F}\subseteq\Gamma of independent contracting elements such that p[f±][f±]=p[f^{\pm}]\cap[f^{\pm}]=\emptyset for any pPp\in P and fF~f\in\tilde{F}.

Now to conclude the proof, it suffices to invoke the same arguments in the proof of Lemma 2.10. Namely, for any finite set FF~F\subseteq\tilde{F}, there is a bounded intersection function τ:\tau:\mathbb{R}\to\mathbb{R} so that for any ffFf\neq f^{\prime}\in F and any R>0R>0,

NR(Ax(f))NR(Ax(f))τ(R).\|N_{R}(\mathrm{Ax}(f))\cap N_{R}(\mathrm{Ax}(f^{\prime}))\|\leq\tau(R).

The function τ\tau depends only on the configuration of the axis Ax(f)=EΓ(f)o\mathrm{Ax}(f)=E_{\Gamma}(f)\cdot o for fFf\in F. That is to say, τ\tau remains the same if FF^{\prime} is chosen as a different set of representatives fEΓ(f)f^{\prime}\in E_{\Gamma}(f) for fFf\in F. Recall that EΓ(f)<ΓE_{\Gamma}(f)<\Gamma is the maximal elementary subgroup in Γ\Gamma defined in Lemma 2.2.

By the contracting property, for any gΓg\in\Gamma and any given R0R\gg 0, the diameter πAx(f)([o,go])\|\pi_{\mathrm{Ax}(f)}([o,go])\| is comparable with NR(Ax(f))[o,go]\|N_{R}(\mathrm{Ax}(f))\cap[o,go]\|. Hence, with at most one exception f0Ff_{0}\in F,

(26) fF{f0}:πAx(f)([o,go])τ,\forall f\in F\setminus\{f_{0}\}:\quad\pi_{\mathrm{Ax}(f)}([o,go])\leq\tau,

and, since pAx(f)Ax(f)p\mathrm{Ax}(f)\neq\mathrm{Ax}(f) for any fFf\in F and a finite set of pPp\in P,

(27) fF:πAx(f)([o,po])τ,\forall f\in F:\quad\pi_{\mathrm{Ax}(f)}([o,po])\leq\tau,

where τ=τ(R)\tau=\tau(R) is a constant depending on the axis Ax(f)\mathrm{Ax}(f) for fFf\in F as above. Consequently, for any g,hΓg,h\in\Gamma, there exists a common fFf\in F satisfying (26) so that

(28) max{πAx(f)([o,go]),πAx(f)([o,ho])}τ\max\{\pi_{\mathrm{Ax}(f)}([o,go]),\pi_{\mathrm{Ax}(f)}([o,ho])\}\leq\tau

Setting L=min{d(o,fo):fF}L=\min\{d(o,fo):f\in F\}, the above equations (26)(27)(28) verify that for any pPp\in P, the word (g,f,p,f1,h)(g,f,p,f^{-1},h) labels an (L,τ)(L,\tau)–admissible path, denoted by γ\gamma, associated with the contracting sets gAx(f)g\mathrm{Ax}(f) and gfpAx(f)gfp\mathrm{Ax}(f). Taking a high power of fFf\in F if necessary, the constant LL can be large enough to satisfy Lemma 2.10, so any geodesic α\alpha with same endpoints as γ\gamma rr–fellow travels γ\gamma, so go,gfogo,gfo and gfpo,gfpf1ogfpo,gfpf^{-1}o have at most a distance rr to α\alpha. Letting D=8r+L+max{d(o,po):pP}D=8r+L+\max\{d(o,po):p\in P\} concludes the proof of the lemma. ∎

5.4. First consequences on confined subgroups

Until the end of this subsection, consider a subgroup H<GH<G, which is confined by Γ\Gamma with a finite confining subset PP. Assume that EG(Γ)P{1}E_{G}(\Gamma)\cap P\setminus\{1\}\neq\emptyset. The main situations we keep in mind are G=Isom(X)G=\mathrm{Isom}(\mathrm{X}) or G=ΓG=\Gamma.

The following is an immediate consequence of Lemma 5.7 applied to (g,g1)(g,g^{-1}), and pPp\in P is chosen for gfgf according to the definition of confined subgroups.

Lemma 5.8.

There exists a finite subset FF of contracting elements in Γ\Gamma with the following property: for any gΓg\in\Gamma, there exist fFf\in F and pPp\in P such that gfpf1g1gfpf^{-1}g^{-1} lies in HH and

|d(o,gfpf1g1o)2d(o,go)|D|d(o,gfpf^{-1}g^{-1}o)-2d(o,go)|\leq D

where DD depends only on FF and PP.

This allows us to derive the following desirable property.

Lemma 5.9.

There exists a finite set FF of pairwise independent contracting elements in Γ\Gamma with the following property. Fix a conjugate gHg1gHg^{-1} for gΓg\in\Gamma. For any f1f2Ff_{1}\neq f_{2}\in F there exists p1,p2Pp_{1},p_{2}\in P such that fipifi1gHg1f_{i}p_{i}f_{i}^{-1}\in gHg^{-1} and their product c(f1,f2):=(f1p1f11)(f2p2f21)c(f_{1},f_{2}):=(f_{1}p_{1}f_{1}^{-1})\cdot(f_{2}p_{2}f_{2}^{-1}) is a contracting element. Moreover, any two such c(f1,f2)c(f_{1},f_{2}) for distinct pairs (f1,f2)F×F(f_{1},f_{2})\in F\times F are independent.

Proof.

For i=1,2i=1,2, applying Lemma 5.8 to g1g^{-1} with fif_{i} yields the choice of pip_{i} in PP so that hi:=fipifi1h_{i}:=f_{i}p_{i}f_{i}^{-1} lies in gHg1gHg^{-1}. It remains to note that h:=h1h2h:=h_{1}\cdot h_{2} is contracting. To this end, as Ax(f1)\mathrm{Ax}(f_{1}) and Ax(f2)\mathrm{Ax}(f_{2}) have bounded intersection, setting L:=min{d(o,fo):fF}L:=\min\{d(o,fo):f\in F\}, we have that the power hnh^{n} labels an (L,τ)(L,\tau)–admissible path as follows

γ:=nhn([o,f1o]f1[o,p1o]f1p1[o,f11o]h1[o,f2o]f2[o,p2o]f2p2[o,f21o])\gamma:=\cup_{n\in\mathbb{Z}}h^{n}([o,f_{1}o]f_{1}[o,p_{1}o]f_{1}p_{1}[o,f_{1}^{-1}o]\cdot h_{1}[o,f_{2}o]f_{2}[o,p_{2}o]f_{2}p_{2}[o,f_{2}^{-1}o])

where the associated contracting subsets are given by the corresponding translated axis of f1f_{1} and f2f_{2}. Choosing LL sufficiently large, the path γ\gamma is contracting, so hh is contracting.

Finally, if h=c(f1,f2)h=c(f_{1},f_{2}) and h=c(f1,f2)h^{\prime}=c(f_{1}^{\prime},f_{2}^{\prime}) for (f1,f2)(f1,f2)(f_{1},f_{2})\neq(f_{1}^{\prime},f_{2}^{\prime}), then we need to show that hh and hh^{\prime} are independent. That amounts to proving that their axes γ\gamma and γ\gamma^{\prime} have infinite Hausdorff distance. Indeed, if not, then γ\gamma lies in a finite neighborhood of γ\gamma^{\prime}, so they fellow travel a common bi-infinite geodesic α\alpha (which could be obtained by applying Cantor argument with Ascoli-Arzela Lemma). For definiteness, say f1f1f2f_{1}^{\prime}\neq f_{1}\neq f_{2}^{\prime}. The fellow travel property by Proposition 2.9 then forces a large intersection of some axis of f1f_{1} with the axis of f1f_{1}^{\prime} or the axis of f2f_{2}^{\prime}. This would lead to a contraction with bounded intersection of FF. ∎

We get the following corollary on confined subgroups.

Lemma 5.10.

Assume that H<GH<G is a subgroup confined by Γ\Gamma with a finite confining subset PP that intersects trivially EG(Γ)E_{G}(\Gamma). Then HH must be a non-elementary subgroup with a contracting element. Moreover, Λ(Γo)[Λ(Ho)]\Lambda(\Gamma o)\subseteq[\Lambda(Ho)].

Proof.

Let gnoξΛ(Γo)g_{n}o\to\xi\in\Lambda(\Gamma o) for gnΓg_{n}\in\Gamma. As above, we then choose fnFf_{n}\in F and pnPp_{n}\in P so that hn:=gnfnpnfn1gn1Hh_{n}:=g_{n}f_{n}p_{n}f_{n}^{-1}g_{n}^{-1}\in H labels an (L,τ)(L,\tau)–admissible path. Consider the sequence of contracting quasi-geodesic Xn=gnAx(fn)X_{n}=g_{n}\mathrm{Ax}(f_{n}), so it follows that hnoNr(Xn)h_{n}o\cap N_{r}(X_{n})\neq\emptyset where rr is given by Proposition 2.9. Hence, we proved that gno,hnoΩ(Nr(Xn))g_{n}o,h_{n}o\in\Omega(N_{r}(X_{n})), so the Assumption (B) in Definition 2.13 implies that hnoh_{n}o also tends to [ξ][\xi]. The proof is now complete. ∎

6. Conservativity of boundary actions of confined subgroups

Our goal is to prove Theorem 1.11. In this section we retain the same setting as §4.

6.1. Some preliminary geometric lemmas

We first prepare some geometric lemmas. Recall that 𝒞ϵhor\mathcal{C}_{\epsilon}^{\mathrm{hor}} is a subset in 𝒞\mathcal{C} (Assumption D) on which the Busemann cocycles converge up to an additive error ϵ\epsilon in (8).

Lemma 6.1.

Let γ=[o,ξ]\gamma=[o,\xi] be a geodesic ray ending at [ξ][\xi] for some ξ𝒞ϵhor\xi\in\mathcal{C}_{\epsilon}^{\mathrm{hor}}. If yXy\in\mathrm{X} is a point that satisfies d(y,z)d(o,z)+Dd(y,z)\leq d(o,z)+D for some zγz\in\gamma and a real number DD\in\mathbb{R}, then y(ξ,ϵ+D)y\in\mathcal{HB}(\xi,\epsilon+D).

Proof.

According to Definition 2.22, the horoball (ξ,L)\mathcal{HB}(\xi,L) consists of points xXx\in\mathrm{X} such that B[ξ](x,o)LB_{[\xi]}(x,o)\leq L. Noting that d(γ(t),z)+d(z,o)=td(\gamma(t),z)+d(z,o)=t for zγz\in\gamma and td(o,z)t\geq d(o,z), so d(γ(t),y)t[d(γ(t),z)+d(z,y)]d(γ(t),z)d(z,o)d(z,y)d(z,o)D.d(\gamma(t),y)-t\leq[d(\gamma(t),z)+d(z,y)]-d(\gamma(t),z)-d(z,o)\leq d(z,y)-d(z,o)\leq D. Taking tt\to\infty shows B[ξ](y,o)ϵ+DB_{[\xi]}(y,o)\leq\epsilon+D by (8). That is, y(ξ,ϵ+D)y\in\mathcal{HB}(\xi,\epsilon+D). ∎

We now arrive at a crucial observation in the proof of Theorem 6.5.

Lemma 6.2.

Given τ,C>0\tau,C>0 there exists L=L(C,τ)L=L(C,\tau) with the following property. Assume that Ax(f)\mathrm{Ax}(f) is CC–contracting. Consider a distinct axis tAx(f)Ax(f)t\mathrm{Ax}(f)\neq\mathrm{Ax}(f) for some tGt\in G. Let g1oAx(f),g2otAx(f)g_{1}o\in\mathrm{Ax}(f),g_{2}o\in t\mathrm{Ax}(f) so that [g1o,g2o][g_{1}o,g_{2}o] intersects Ax(f)\mathrm{Ax}(f) in a diameter at least LL, and g2og_{2}o lies in a τ\tau–neighborhood of the projection of Ax(f)\mathrm{Ax}(f) to tAx(f)t\mathrm{Ax}(f). Then h:=g11g2h:=g_{1}^{-1}g_{2} is a contracting element.

Proof.

The CC–contracting property of Ax(f)\mathrm{Ax}(f) implies that any geodesic outside NC(Ax(f))N_{C}(\mathrm{Ax}(f)) has CC–bounded projection, so [o,ho][o,ho] has 2C2C–bounded projection to Ax(f)\mathrm{Ax}(f). According to the assumption, [o,ho][o,ho] has τ\tau–bounded projection to hAx(f)h\mathrm{Ax}(f), so this verifies that the path

γ=nhn[o,ho]\gamma=\cup_{n\in\mathbb{Z}}h^{n}[o,ho]

is an (L,τ+2C)(L,\tau+2C)–admissible path with associated contracting subsets {hnAx(f),hntAx(f):n}\{h^{n}\mathrm{Ax}(f),h^{n}t\mathrm{Ax}(f):n\in\mathbb{Z}\}. If LL is sufficiently large, then γ\gamma is a contracting quasi-geodesic, concluding the proof that gg is contracting. ∎

As a corollary, we have.

Lemma 6.3.

Assume that ξΛr,Fcon(Γo)\xi\in\Lambda^{\textrm{con}}_{\textrm{r,F}}{(\Gamma o)} is an (r,F)(r,F)–conical point. Let [o,ξ][o,\xi] be a geodesic ray ending at [ξ][\xi] with two distinct (r,f)(r,f)–barriers XYX\neq Y. Let g1oX,g2oYg_{1}o\in X,g_{2}o\in Y be CC–close to the corresponding entry points of [o,ξ][o,\xi] in NC(X)N_{C}(X) and NC(Y)N_{C}(Y). Then g11g2g_{1}^{-1}g_{2} is a contracting element.

Recall that a geodesic metric space is geodesically complete, if any geodesic segment extends to a (possibly non-unique) bi-infinite geodesic. A smooth Hadamard manifold is geodesically complete.

Lemma 6.4.

Assume that X\mathrm{X} is a proper, geodesically complete, CAT(0) space. Let pIsom(X)p\in\mathrm{Isom}(\mathrm{X}) be a parabolic isometry. If pp fixes a conical point ξ\xi in Λr,Fcon(Γo)\Lambda^{\textrm{con}}_{\textrm{r,F}}{(\Gamma o)} for a proper action of Γ\Gamma on X\mathrm{X}, then pp has the unique fixed point.

Proof.

By [30], under the assumption on X\mathrm{X}, the fixed point set Fix(p)\mathrm{Fix}(p) of a parabolic element has diameter at most π/2\pi/2 in the Tits metric on the visual boundary X\partial{\mathrm{X}}. A conical point ξ\xi is visible from any other point ξηX\xi\neq\eta\in\partial{\mathrm{X}}; that is, there is a bi-infinite geodesic between ξ\xi and η\eta. Thus, the angular metric between ξ\xi and η\eta is π\pi, so being the induced length metric, the Tits metric from ξ\xi to η\eta is at least π\pi. Hence, Fix(p)\mathrm{Fix}(p) is singleton, completing the proof. ∎

6.2. Proof of Theorem 1.11

Assume that HH preserves the measure class of μo\mu_{o}. By Lemma 2.35, μo\mu_{o} is supported on the set of conical points [Λr,Fcon(Γo)][\Lambda^{\textrm{con}}_{\textrm{r,F}}{(\Gamma o)}]. By Lemma 2.35, [Λr,Fcon(Γo)][\Lambda^{\textrm{con}}_{\textrm{r,F}}{(\Gamma o)}] is a subset of 𝒞20Chor\mathcal{C}_{20C}^{\mathrm{hor}} (defined in Assumption D).

The conservativity of the action of HH on (X,μo)(\partial{\mathrm{X}},\mu_{o}) as stated in Theorem 1.11, follows from a combination of the next two theorems 6.5 and 6.7, which applies assuming the first condition (i) and second condition (ii) respectively. Note that there are situations (torsion-free HH with a finite confining set) where both theorems apply, but we stress that the arguments are of different flavors. The first result applies under the assumption of a compact confining subset PP without torsion; the second result crucially uses the Lemma 5.8 based on the finiteness of PP with torsion allowed.

Theorem 6.5.

Assume that H<Isom(X)H<\mathrm{Isom}(\mathrm{X}) is a torsion-free discrete confined subgroup. If X\mathrm{X} is neither hyperbolic nor geodesically complete CAT(0), assume, in addition, that PP is finite. Then the big horospheric limit set ΛHor(Ho)\Lambda^{\mathrm{Hor}}{(Ho)} contains a μo\mu_{o}–full subset of [Λr,Fcon(Γo)][\Lambda^{\textrm{con}}_{\textrm{r,F}}{(\Gamma o)}].

Remark 6.6.

The proof shows that the conclusion of Theorem 6.5 remains valid for any atomless measure ν\nu supported on [Λr,Fcon(Γo)][\Lambda^{\textrm{con}}_{\textrm{r,F}}{(\Gamma o)}]. If X\mathrm{X} is hyperbolic, Λr,Fcon(Γo)\Lambda^{\textrm{con}}_{\textrm{r,F}}{(\Gamma o)} could be replaced with Λcon(Γo)\Lambda^{\textrm{con}}{(\Gamma o)}. In particular, harmonic measures on horofunction boundary which can arise as the hitting measure of a random walk on Γ\Gamma are among such examples.

Proof.

Let Λ0\Lambda_{0} be the countable union of fixed points [h±][h^{\pm}] of all contracting elements hHh\in H. If X\mathrm{X} is hyperbolic or CAT(0), we adjoin into Λ0\Lambda_{0} the fixed points of all parabolic elements in HH, which are countably many by Lemma 6.4. Note that μo(Λ0)=0\mu_{o}(\Lambda_{0})=0 for μ0\mu_{0} has no atom. By Lemma 2.35, we will prove the μo\mu_{o}–full set of points ξ[Λr,Fcon(Γo)]Λ0\xi\in[\Lambda^{\textrm{con}}_{\textrm{r,F}}{(\Gamma o)}]\setminus\Lambda_{0} is contained in [ΛHor(Ho)][\Lambda^{\mathrm{Hor}}{(Ho)}]. By definition of an (r,F)(r,F)–conical point, there exists a sequence of (r,fn)(r,f_{n})–barriers gnΓg_{n}\in\Gamma on a geodesic ray γ=[o,ξ]\gamma=[o,\xi], where fnFf_{n}\in F is taken over the finite set FΓF\subseteq\Gamma. By definition, we have d(gno,[o,ξ]),d(gnfno,[o,ξ])rd(g_{n}o,[o,\xi]),d(g_{n}f_{n}o,[o,\xi])\leq r. Passing to a subsequence, we may assume that f:=fnf:=f_{n} for all n1n\geq 1. Moreover, we can assume that XnX_{n} are all distinct; otherwise ξ\xi lies in the limit set of the same XnX_{n} for n0n\gg 0, so is fixed by a contracting element gnfgn1g_{n}fg_{n}^{-1}. This implies ξΛ0\xi\in\Lambda_{0}, contradicting the assumption.

Let PIsom(X,d)P\subseteq\mathrm{Isom}(\mathrm{X},d) be a compact confining subset for HH, so D:=max{d(o,po):pP}<D:=\max\{d(o,po):p\in P\}<\infty.

Choose pnPp_{n}\in P so that hn:=gnpngn1Hh_{n}:=g_{n}p_{n}g_{n}^{-1}\in H. Now, let z[o,ξ]z\in[o,\xi] so that d(gno,z)rd(g_{n}o,z)\leq r. Thus, d(hno,z)r+D+d(o,gno)2r+D+d(o,z)d(h_{n}o,z)\leq r+D+d(o,g_{n}o)\leq 2r+D+d(o,z). According to Lemma 6.1, we see that hnoHoh_{n}o\in Ho lies in the horoball (ξ,o,2r+D+ϵ)\mathcal{HB}(\xi,o,2r+D+\epsilon).

We shall prove that {hno:n1}\{h_{n}o:n\geq 1\} is an infinite (discrete) subset, which thus converges to ξ\xi by Lemma 2.36. Arguing by contradiction, assume now that {hno:n1}\{h_{n}o:n\geq 1\} is a finite set. By taking a subsequence, the proper action of HH allows us to assume that h:=hn=hmh:=h_{n}=h_{m} for any n,m1n,m\geq 1.

Case 1. PP is finite. We may then assume that pn=pp_{n}=p for all n1n\geq 1. We thus obtain gm1gnh=hgm1gng_{m}^{-1}g_{n}h=hg_{m}^{-1}g_{n}. By Lemma 6.3, gm1gng_{m}^{-1}g_{n} is a contracting element for any two nm1n\neq m\geq 1. If hh is of infinite order, hE(gm1gn)h\in E(g_{m}^{-1}g_{n}) gives a contradiction, as any infinite order element in E(gm1gn)E(g_{m}^{-1}g_{n}) is a contracting element.

Case 2. PP may be an infinite compact set, but X\mathrm{X} is assumed to be hyperbolic or geodesically complete CAT(0). Note that pnp_{n} might be distinct in general. As HH is torsion-free, pnp_{n} must be of infinite order. According to classification of isometries, pnp_{n} is either hyperbolic or parabolic.

As d(hgno,gno)=d(o,pno)<Dd(hg_{n}o,g_{n}o)=d(o,p_{n}o)<D, the convergence gno[ξ]g_{n}o\to[\xi] implies that hgno[ξ]hg_{n}o\to[\xi] and then hh fixes [ξ][\xi]. So ξΛ0\xi\in\Lambda_{0} gives a contradiction. The proof is complete. ∎

In the following statement, we emphasize that H<Isom(X)H<\mathrm{Isom}(\mathrm{X}) is not necessarily a discrete subgroup. The proof relies on Lemma 5.8 applied with G=Isom(X)G=\mathrm{Isom}(\mathrm{X}) here, which does not use the proper action HXH\curvearrowright\mathrm{X} as well. Compared with Theorem 6.5, HH may contain torsion elements, but PE(Γ)P\cap E(\Gamma) is assumed to be trivial.

Theorem 6.7.

Assume that H<Isom(X)H<\mathrm{Isom}(\mathrm{X}) is a subgroup confined by Γ\Gamma with a finite confining subset PP that intersects trivially E(Γ)E(\Gamma). Then ΛHor(Ho)\Lambda^{\mathrm{Hor}}{(Ho)} contains [Λr,Fcon(Γo)][\Lambda^{\textrm{con}}_{\textrm{r,F}}{(\Gamma o)}] as a subset. Moreover, Λ(Γo)[Λ(Ho)]\Lambda(\Gamma o)\subseteq[\Lambda(Ho)].

Proof.

By definition of an (r,F)(r,F)–conical point, there exists a sequence of (r,F)(r,F)–barriers gnΓg_{n}\in\Gamma on a geodesic ray γ=[o,ξ]\gamma=[o,\xi]. By definition, we have d(gno,[o,ξ])rd(g_{n}o,[o,\xi])\leq r.

By Lemma 5.8, we can choose fnFf_{n}\in F and pnPp_{n}\in P so that hn:=gnfnpnfn1gn1Hh_{n}:=g_{n}f_{n}p_{n}f_{n}^{-1}g_{n}^{-1}\in H labels an (L,τ)(L,\tau)–admissible path. If d(gno,gmo)0d(g_{n}o,g_{m}o)\gg 0 for any nmn\neq m, then hnohmoh_{n}o\neq h_{m}o by Lemma 2.12. Thus {hno:n1}\{h_{n}o:n\geq 1\} is an infinite subset. Setting

D=maxfF{d(o,fo)}+maxpP{d(o,po)}D=\max_{f\in F}\{d(o,fo)\}+\max_{p\in P}\{d(o,po)\}

we argue exactly as in the proof of Theorem 6.5 and obtain that hnoHoh_{n}o\in Ho lies in the horoball ([ξ],2r+2D+ϵ)\mathcal{HB}([\xi],2r+2D+\epsilon) (the main issue there was proving the infiniteness of {hno:n1}\{h_{n}o:n\geq 1\}). Hence, {hno:n1}\{h_{n}o:n\geq 1\} converges to [ξ][\xi] by Lemma 2.36, so ξ\xi is a big horospheric limit point. ∎

Corollary 6.8.

In the setting of Theorems 6.5 or 6.7, if HH is a subgroup of Γ\Gamma, then [Λ(Γo)]=[Λ(Ho)][\Lambda(\Gamma o)]=[\Lambda(Ho)].

Proof.

Under the assumptions of Theorem 6.5, we proved that a μo\mu_{o}–full subset Λ\Lambda of [Λr,Fcon(Γo)][\Lambda^{\textrm{con}}_{\textrm{r,F}}{(\Gamma o)}] is contained in the horospheric limit set ΛHor(Ho)\Lambda^{\mathrm{Hor}}{(Ho)}. Taking a countable intersection Γ:=gΓgΛ\Gamma:=\cap_{g\in\Gamma}g\Lambda allows to assume that Λ\Lambda is Γ\Gamma–invariant. If X\mathrm{X} is hyperbolic or CAT(0), it is well-known that the limit set Λ(Γo)\Lambda(\Gamma o) is a Γ\Gamma–invariant minimal subset. Hence, the topological closure of Λ\Lambda recovers Λ(Γo)\Lambda(\Gamma o), thus verifying Λ(Γo)Λ(Ho)\Lambda(\Gamma o)\subseteq\Lambda(Ho), so the proof is completed in this case.

Otherwise, PP is a finite set by assumption. As H<ΓH<\Gamma is assumed, the conclusion follows immediately from Lemma 5.10. ∎

To conclude this subsection, we record a stronger statement, provided that HH is a normal subgroup. The proof strategy is due to [27].

Theorem 6.9.

Suppose that HH is an infinite normal subgroup of Γ\Gamma. Then Λhor(Ho)\Lambda^{\mathrm{hor}}{(Ho)} contains ΛMyr(Γo)\Lambda^{\mathrm{Myr}}{(\Gamma o)} as a subset.

Proof.

Given ξΛMyr(Γo)\xi\in\Lambda^{\mathrm{Myr}}{(\Gamma o)}, let γ\gamma be a geodesic ray starting at oo and ending at [ξ][\xi]. Let us take a contracting element fHf\in H, as an infinite normal subgroup contains infinitely many ones. By definition of ΛMyr(Γo)\Lambda^{\mathrm{Myr}}{(\Gamma o)} in (21), γ\gamma contains infinitely many (r,fn)(r,f^{n})–barriers gng_{n} for any n1n\geq 1. By normality of HH, we have gnfngn1Hg_{n}f^{n}g_{n}^{-1}\in H forms a sequence of elements, which enters any given horoball based on [ξ][\xi]. With Lemma 2.36, this implies that ΛMyr(Γo)[Λhor(Ho)]\Lambda^{\mathrm{Myr}}{(\Gamma o)}\subseteq[\Lambda^{\mathrm{hor}}{(Ho)}]. ∎

Part II Growth inequalities for confined subgroups

7. Shadow Principle for confined subgroups

We shall establish in this section a Shadow Principle for confined subgroups, which is crucial for the next two sections §8 and 9. The basic setup is as follows.

  • The auxiliary proper action ΓX\Gamma\curvearrowright\mathrm{X} is assumed to be of divergence type with contracting elements.

  • Let X\partial{\mathrm{X}} be a convergence boundary for X\mathrm{X} and {μx:xX}\{\mu_{x}:x\in\mathrm{X}\} be a quasi-conformal, Γ\Gamma–equivariant density of dimension ωΓ\omega_{\Gamma} on X\partial{\mathrm{X}}.

  • Let HH be a discrete subgroup of a group G<Isom(X)G<\mathrm{Isom}(\mathrm{X}), confined by Γ\Gamma, with a finite confining subset PP that intersects trivially E(Γ){1}E(\Gamma)\setminus\{1\}.

Recall that E(Γ)E(\Gamma) is the [][\cdot]–class stabilizer of [Λ(Γo)][\Lambda(\Gamma o)] in Isom(X)\mathrm{Isom}(\mathrm{X}) (Def. 5.5).

In this section, let FΓF\subseteq\Gamma be a set of three independent contracting elements, and the constants L,τ,rL,\tau,r given by Lemma 5.7.

7.1. Shadow Principle for conformal density supported on boundary

Below, the constant λ\lambda is given in Definition 2.24, and ϵ\epsilon is the convergence error of the horofunction in Assumption D. Note that ϵ=0\epsilon=0 for the horofunction boundary. By Corollary 6.8, we have [Λ(Ho)=[Λ(Γo)][\Lambda(Ho)=[\Lambda(\Gamma o)].

Lemma 7.1 (Shadow Principle).

Let {μx}xX\{\mu_{x}\}_{x\in\mathrm{X}} be a ωH\omega_{H}–dimensional HH–quasi-equivariant quasi-conformal density supported on X\partial{\mathrm{X}}. Then there exists r0>0r_{0}>0 such that

μyeωHd(x,y)λμx(ΠxF(y,r))λ,ϵ,rμyeωHd(x,y)\begin{array}[]{rl}\|\mu_{y}\|e^{-\omega_{H}\cdot d(x,y)}\quad\prec_{\lambda}\quad\mu_{x}(\Pi_{x}^{F}(y,r))\quad\prec_{\lambda,\epsilon,r}\quad\|\mu_{y}\|e^{-\omega_{H}\cdot d(x,y)}\\ \end{array}

for any x,yΓox,y\in\Gamma o and rr0r\geq r_{0}.

Proof.

We start by proving the lower bound, which is the key part of the proof. Write explicitly x=g1ox=g_{1}o and y=g2oy=g_{2}o for g1,g2Γg_{1},g_{2}\in\Gamma. Let g=g11g2g=g_{1}^{-1}g_{2}.

1. Lower Bound. Fix a Borel subset U:=[Λ(Ho)]U:=[\Lambda(Ho)] such that μo(U)>0\mu_{o}(U)>0. For given ξU\xi\in U, there exists a sequence of points hnoXh_{n}o\in\mathrm{X} such that hnoξh_{n}o\to\xi. Since FF is a finite set, up to taking a subsequence of hnoh_{n}o, there exist fFf\in F and r>0r>0 given by Lemma 5.7 such that for any pPp\in P, gg is an (r,f)(r,f)–barrier for any geodesic [o,gfpf1hno][o,gfpf^{-1}h_{n}o]. This implies gfpf1hnoΩoF(go,r)gfpf^{-1}h_{n}o\in\Omega_{o}^{F}(go,r), which tends to gfpf1ξgfpf^{-1}\xi. In addition, we can choose pfPp_{f}\in P according to definition of the confined subgroup HH so that g2fpf(g2f)1Hg_{2}fp_{f}(g_{2}f)^{-1}\in H. By definition of the shadow,

(29) gfpff1ξΠoF(go,r)gfp_{f}f^{-1}\xi\in\Pi_{o}^{F}(go,r)

Note that fFf\in F depends on ξU\xi\in U, and FF consists of three elements.

Consequently, the set UU can be decomposed as a disjoint union of three sets U1,U2,U3U_{1},U_{2},U_{3}: for each UiU_{i}, there exist fiF,piPf_{i}\in F,p_{i}\in P such that gfipifi1UiΠoF(go,r)gf_{i}p_{i}f_{i}^{-1}U_{i}\subseteq\Pi_{o}^{F}(go,r); equivalently,

(30) g2fipifi1UiΠg1oF(g2o,r).g_{2}f_{i}p_{i}f_{i}^{-1}U_{i}\subseteq\Pi_{g_{1}o}^{F}(g_{2}o,r).

Denote L=max{2d(o,fo)+d(o,po):fF,pP}<L=\max\{2d(o,fo)+d(o,po):f\in F,p\in P\}<\infty. As d(o,fipifi1o)L,d(o,f_{i}p_{i}f_{i}^{-1}o)\leq L, there exists θ=θ(L,ωH)>0\theta=\theta(L,\omega_{H})>0 by quasi-conformality (12) such that

(31) μg2o(g2fipifi1Ui)θμg2fipifi1o(g2fipifi1Ui).\mu_{g_{2}o}(g_{2}f_{i}p_{i}f_{i}^{-1}U_{i})\geq\theta\cdot\mu_{g_{2}f_{i}p_{i}f_{i}^{-1}o}(g_{2}f_{i}p_{i}f_{i}^{-1}U_{i}).
gogogfogfoooΠoF(go,r)gfpf1hnξ\Pi_{o}^{F}(go,r)\ni\ gfpf^{-1}h_{n}\xir\leq rr\leq rhnoξUh_{n}o\rightarrow\xi\in Ugfpf1hnogfpf^{-1}h_{n}of1f^{-1}pPp\in P
Figure 7. Upper bound in Shadow Principle: transporting a large set UU into the shadow by Lemma 5.7.

We apply the HH–quasi-equivariance (11) with g2fipi1fi1g21Hg_{2}f_{i}p_{i}^{-1}f_{i}^{-1}g_{2}^{-1}\in H to the right-hand side:

(32) μg2fipifi1o(g2fipifi1Ui)λ1μg2o(g2Ui).\mu_{g_{2}f_{i}p_{i}f_{i}^{-1}o}(g_{2}f_{i}p_{i}f_{i}^{-1}U_{i})\geq\lambda^{-1}\cdot\mu_{g_{2}o}(g_{2}U_{i}).

Combining together (30), (31) and (32), we have

μg2o(Πg1oF(g2o,r))\displaystyle\mu_{g_{2}o}(\Pi_{g_{1}o}^{F}(g_{2}o,r)) 1i3μg2o(g2Ui)/3\displaystyle\geq\sum_{1\leq i\leq 3}\mu_{g_{2}o}(g_{2}U_{i})/3
λ1θμg2o(U)/3\displaystyle\geq\lambda^{-1}\theta\cdot\mu_{g_{2}o}(U)/3

where the last line uses the HH–invariance of U=[Λ(Ho)]=[Λ(Γo)]U=[\Lambda(Ho)]=[\Lambda(\Gamma o)] by Lemma 5.10.

Denote M:=λ2θ/3M:=\lambda^{-2}\cdot\theta/3. We now conclude the proof for the lower bound via quasi-conformality (12):

μg1o(Πg1oF(g2o,r))λ1eωd(g1o,g2o)μg2o(Πg1oF(g2o,r))Mμg2o(U)eωd(g1o,g2o)\begin{array}[]{rl}\mu_{g_{1}o}(\Pi_{g_{1}o}^{F}(g_{2}o,r))&\geq\lambda^{-1}\cdot e^{-\omega\cdot d(g_{1}o,g_{2}o)}\cdot\mu_{g_{2}o}(\Pi_{g_{1}o}^{F}(g_{2}o,r))\\ \\ &\geq M\cdot\mu_{g_{2}o}(U)\cdot e^{-\omega\cdot d(g_{1}o,g_{2}o)}\end{array}

2. Upper Bound. Fix rr0r\geq r_{0}. Given ξΠg1oF(g2o,r)\xi\in\Pi_{g_{1}o}^{F}(g_{2}o,r), there is a sequence of znXz_{n}\in\mathrm{X} tending to ξ\xi such that γnB(g2o,r)\gamma_{n}\cap B(g_{2}o,r)\neq\emptyset for γn:=[g11o,zn]\gamma_{n}:=[g_{1}^{-1}o,z_{n}]. Since Buseman cocyles extend continuously to the horofunction boundary, we obtain

|Bξ(g1o,g2o)d(g1o,g2o))|2r.|B_{\xi}(g_{1}o,g_{2}o)-d(g_{1}o,g_{2}o))|\leq 2r.

The upper bound is given as follows:

μg1o(Πg1oF(g2o,r))Πg1oF(g2o,r)eωBξ(g1o,g2o)𝑑μg2o(ξ)λe2ωrμg2oeωd(g1o,g2o)\begin{array}[]{rl}\mu_{g_{1}o}(\Pi_{g_{1}o}^{F}(g_{2}o,r))&\leq\displaystyle\int_{\Pi_{g_{1}o}^{F}(g_{2}o,r)}e^{-\omega B_{\xi}(g_{1}o,g_{2}o)}d\mu_{g_{2}o}(\xi)\\ \\ &\leq\lambda e^{2\omega r}\|\mu_{g_{2}o}\|\cdot e^{-\omega d(g_{1}o,g_{2}o)}\end{array}

The proof of lemma is complete. ∎

7.2. Shadow Principle for conformal density supported inside

Fix a basepoint oXo\in\mathrm{X} and ω>ωH\omega>\omega_{H}. Given xXx\in\mathrm{X}, define μx=1𝒫H(ω,o,o)hHeωd(x,ho)\mu_{x}=\frac{1}{\mathcal{P}_{H}(\omega,o,o)}\sum_{h\in H}\mathrm{e}^{-\omega d(x,ho)} supported on HoHo, we have

zHo:dμxdμy(z)=eω[d(x,z)d(y,z)]\forall z\in Ho:\quad\frac{d\mu_{x}}{d\mu_{y}}(z)=\mathrm{e}^{-\omega[d(x,z)-d(y,z)]}
Lemma 7.2.

There are constants θ,D>0\theta,D>0 such that, given any Δ>0\Delta>0 and, g1,g2Gg_{1},g_{2}\in G we have

#HAΓ(g2o,n,Δ)θ#HAΓ(g2o,n,Δ+D)Ωg1oF(g2o,D).\#H\cap A_{\Gamma}(g_{2}o,n,\Delta)\leq\theta\#H\cap A_{\Gamma}(g_{2}o,n,\Delta+D)\cap\Omega_{g_{1}o}^{F}(g_{2}o,D).
Proof.

Let hHh\in H and g:=g11g2g:=g_{1}^{-1}g_{2}. According to Lemma 5.7, for given gg and g21hg_{2}^{-1}h, we pick fhFf_{h}\in F such that for any pPp\in P, the word (g,fh,p,fh1,g21h)(g,f_{h},p,f_{h}^{-1},g_{2}^{-1}h) labels an (L,τ)(L,\tau)–admissible path, where gogo is an (r,F)(r,F)–barrier for fhFf_{h}\in F. This implies gfhpfh1g21hΩoF(go,r)gf_{h}pf_{h}^{-1}g_{2}^{-1}h\in\Omega_{o}^{F}(go,r), or equivalently 𝐡:=g2fhpfh1g21hΩg1oF(g2o,r)\mathbf{h}:=g_{2}f_{h}pf_{h}^{-1}g_{2}^{-1}h\in\Omega_{g_{1}o}^{F}(g_{2}o,r). As pPp\in P can be arbitrarily chosen, the definition of confined subgroup HH allows us to choose phPp_{h}\in P for the element g21hGg_{2}^{-1}h\in G so that h1g2fhph(fh1g21h)h^{-1}g_{2}f_{h}\cdot p_{h}\cdot(f_{h}^{-1}g_{2}^{-1}h) lies in HH and so 𝐡=g2fhphfh1g21h\mathbf{h}=g_{2}f_{h}p_{h}f_{h}^{-1}g_{2}^{-1}h lies in HH.

Set D=max{d(o,fo)+d(o,po):fF,pP}D=\max\{d(o,fo)+d(o,po):f\in F,p\in P\}. If d(ho,g2o)[nΔ,n+Δ]d(ho,g_{2}o)\in[n-\Delta,n+\Delta] we have d(𝐡o,g2o)[nΔD,n+Δ+D]d(\mathbf{h}o,g_{2}o)\in[n-\Delta-D,n+\Delta+D]. We claim that this assignment

π:HAΓ(g2o,n,Δ)\displaystyle\pi:H\cap A_{\Gamma}(g_{2}o,n,\Delta) HAΓ(g2o,n,Δ+D)Ωg1oF(g2o,r)\displaystyle\longrightarrow H\cap A_{\Gamma}(g_{2}o,n,\Delta+D)\cap\Omega_{g_{1}o}^{F}(g_{2}o,r)
h\displaystyle h 𝐡\displaystyle\longmapsto\mathbf{h}

is uniformly finite to one. Indeed, assume 𝐡=𝐡\mathbf{h}=\mathbf{h^{\prime}} for some hhh\neq h^{\prime}. Applying multiplication by g11g_{1}^{-1} on the left for 𝐡=𝐡\mathbf{h}=\mathbf{h^{\prime}} gives gfhphfh1g21h=gfhphfh1g21hgf_{h}p_{h}f_{h}^{-1}g_{2}^{-1}h=gf_{h^{\prime}}p_{h^{\prime}}f_{h^{\prime}}^{-1}g_{2}^{-1}h^{\prime}. As mentioned above, these two words label two (L,τ)(L,\tau)–admissible paths with the same endpoints. Noting that g1h,g1hAΓ(o,n,Δ)g^{-1}h,g^{-1}h^{\prime}\in A_{\Gamma}(o,n,\Delta), Lemma 2.12 implies d(g1ho,g1ho)=d(ho,ho)Rd(g^{-1}ho,g^{-1}h^{\prime}o)=d(ho,h^{\prime}o)\leq R. Consequently, at most N:={hH:d(o,ho)R}N:=\sharp\{h\in H:d(o,ho)\leq R\} elements have the same image under the map π\pi. Setting θ=1/N\theta=1/N thus completes the proof. ∎

The following is the Shadow Principle we want for all s>wHs>w_{H}.

Lemma 7.3.

Fix any ω>ωH\omega>\omega_{H}. Let {μx}xX\{\mu_{x}\}_{x\in\mathrm{X}} be a ω\omega–dimensional HH–quasi-equivariant quasi-conformal density supported on HoXHo\subseteq\mathrm{X}. Then there exists r0>0r_{0}>0 such that

μyeωd(x,y)λμx(ΩxF(y,r))λ,ϵ,rμyeωd(x,y)\begin{array}[]{rl}\|\mu_{y}\|e^{-\omega\cdot d(x,y)}\quad\prec_{\lambda}\quad\mu_{x}(\Omega_{x}^{F}(y,r))\quad\prec_{\lambda,\epsilon,r}\quad\|\mu_{y}\|e^{-\omega\cdot d(x,y)}\\ \end{array}

for any x,yGox,y\in Go and rr0r\geq r_{0}.

Proof.

If zΩxF(y,r)z\in\Omega_{x}^{F}(y,r) we have d(x,z)+2rd(x,y)+d(y,z)d(x,z)+2r\geq d(x,y)+d(y,z), so μx(ΩxF(y,r))e2rωμyeωd(x,y)\mu_{x}(\Omega_{x}^{F}(y,r))\leq\mathrm{e}^{2r\omega}\|\mu_{y}\|e^{-\omega\cdot d(x,y)}. Thus, it remains to prove the lower bound. As the triangle inequality implies μx(ΩxF(y,r))eωd(x,y)μy(ΩxF(y,r))\mu_{x}(\Omega_{x}^{F}(y,r))\geq e^{-\omega d(x,y)}\mu_{y}(\Omega_{x}^{F}(y,r)), it suffices to show μy(ΠxF(y,r))Θμy\mu_{y}(\Pi_{x}^{F}(y,r))\geq\Theta\|\mu_{y}\| for some Θ>0\Theta>0.

By definition,

𝒫H(ω,o,o)μy=hHeωd(y,ho)<\mathcal{P}_{H}(\omega,o,o)\|\mu_{y}\|=\sum_{h\in H}\mathrm{e}^{-\omega d(y,ho)}<\infty

Assume that x=g1o,y=g2ox=g_{1}o,y=g_{2}o, and set g=g11g2g=g_{1}^{-1}g_{2} for simplicity. Note that ΩxF(y,r)=g1ΩoF(go,r)\Omega_{x}^{F}(y,r)=g_{1}\Omega_{o}^{F}(go,r). We remark, however, that μy(ΩxF(y,r))=μgo(ΩoF(go,r))\mu_{y}(\Omega_{x}^{F}(y,r))=\mu_{go}(\Omega_{o}^{F}(go,r)) may not hold, as μx\mu_{x} is only HH–conformal density but g1g_{1} may not be in HH.

By Lemma 7.2, we have

hHesd(g2o,ho)\displaystyle\sum_{h\in H}e^{-sd(g_{2}o,ho)} C1(s)n=0esn#HA(g2o,n,Δ)\displaystyle\leq C_{1}(s)\sum_{n=0}^{\infty}e^{-sn}\#H\cap A(g_{2}o,n,\Delta)
θC1(s)n=0esn#(HA(g2o,n,Δ+D)Ωg1oF(g2o,r))\displaystyle\leq\theta C_{1}(s)\sum_{n=0}^{\infty}e^{-sn}\cdot\#\left(H\cap A(g_{2}o,n,\Delta+D)\cap\Omega_{g_{1}o}^{F}(g_{2}o,r)\right)
θC1(s)C2(s)hHΩg1oF(g2o,r)esd(g2o,ho).\displaystyle\leq\theta C_{1}(s)C_{2}(s)\sum_{h\in H\cap\Omega_{g_{1}o}^{F}(g_{2}o,r)}e^{-sd(g_{2}o,ho)}.

Here, C1(s)=ΔesΔ,C2(s)=(Δ+D)es(Δ+D)C_{1}(s)=\Delta\cdot e^{-s\Delta},C_{2}(s)=(\Delta+D)\cdot e^{-s(\Delta+D)}. Setting Θ=θC1(s)C2(s)\Theta=\theta C_{1}(s)C_{2}(s), this is equivalent to the following

𝒫H(ω,o,o)μy(ΩxF(y,r))=hΩxF(y,r)eωd(y,ho)ΘhΩxF(y,r)eωd(y,ho)=𝒫H(ω,o,o)μy\mathcal{P}_{H}(\omega,o,o)\mu_{y}(\Omega_{x}^{F}(y,r))=\sum_{h\in\Omega_{x}^{F}(y,r)}\mathrm{e}^{-\omega d(y,ho)}\geq\Theta\sum_{h\in\Omega_{x}^{F}(y,r)}\mathrm{e}^{-\omega d(y,ho)}=\mathcal{P}_{H}(\omega,o,o)\|\mu_{y}\|

Hence, the shadow lemma is proved. ∎

8. Cogrowth tightness of confined subgroups

We continue the setup of §7, but with the first item replaced with a stronger assumption

  • The proper action ΓX\Gamma\curvearrowright\mathrm{X} has purely exponential growth.

We first give a uniform upper bound for all conjugates of a confined subgroup HH.

Lemma 8.1.

There exist constants C,Δ>0C,\Delta>0 so that the following holds

(gHg1AΓ(o,n,Δ))CenωH\sharp(gHg^{-1}\cap A_{\Gamma}(o,n,\Delta))\leq C\mathrm{e}^{n\omega_{H}}

for any gGg\in G.

Proof.

By Lemma 5.9, there exists a finite set of contracting elements F~\tilde{F} such that any gHg1gHg^{-1} contains a set, denoted by FgF_{g}, of three pairwise independent contracting elements from F~\tilde{F}. Applying Lemma 2.10 with FgF_{g}, the conclusion follows by the same argument as in [71, Proposition 5.2], where CC does not depend on gg since there are only finitely many choices of FgF~F_{g}\subseteq\tilde{F} independent of gg. We include the sketch below and refer to [71, Proposition 5.2] for full details.

Set Sn:=gHg1AΓ(o,n,Δ)S_{n}:=gHg^{-1}\cap A_{\Gamma}(o,n,\Delta). The idea is to prove the following for any n,m1n,m\geq 1

SnSmθkjkS(n+m+j)\sharp S_{n}\sharp S_{m}\leq\theta\sum_{-k\leq j\leq k}\sharp S(n+m+j)

where θ,k\theta,k do not depend on gg. This is proved by considering the map

π:Sn×Sm\displaystyle\pi:\quad S_{n}\times S_{m} gHg1\displaystyle\quad\longrightarrow\quad gHg^{-1}
(a,b)\displaystyle(a,b) afb\displaystyle\quad\longmapsto\quad afb

where fFgf\in F_{g} is chosen by Lemma 2.10. In particular, afbafb labels admissible path so |d(o,afb)d(o,ao)d(o,bo)k|d(o,afb)-d(o,ao)-d(o,bo)\leq k where kk depends on d(o,fo)d(o,fo). We proved there (also see Lemma 2.12) that π\pi fails to be injective, only if d(ao,bo)d(ao,bo) is greater than a constant depending only on FgF_{g}. This constant determines the value of θ\theta, so the proof of the above inequality is proved. ∎

Lemma 8.2.

There exist constants C,Δ>0C,\Delta>0 with the following property. For any gΓg\in\Gamma and any n1n\geq 1, we have

(gHg1AΓ(o,n,Δ))CenωΓ/2\sharp(gHg^{-1}\cap A_{\Gamma}(o,n,\Delta))\geq C\mathrm{e}^{n\omega_{\Gamma}/2}
Proof.

Let us consider the conjugate g0Hg01g_{0}Hg_{0}^{-1} for given g0Γg_{0}\in\Gamma. Define a map as follows

πg0:AΓ(g01o,n,Δ)\displaystyle\pi_{g_{0}}:\quad A_{\Gamma}(g_{0}^{-1}o,n,\Delta) g0Hg01\displaystyle\quad\longrightarrow\quad g_{0}Hg_{0}^{-1}
g\displaystyle g g0gfp(g0gf)1\displaystyle\quad\longmapsto\quad g_{0}gfp(g_{0}gf)^{-1}

where ff and pp are chosen for g0gg_{0}g according to Lemma 5.8. Hence, |d(o,πg0(g)o)2d(o,g0go)|D|d(o,\pi_{g_{0}}(g)o)-2d(o,g_{0}go)|\leq D so we obtain πg0(g)AΓ(o,2n,Δ)\pi_{g_{0}}(g)\in A_{\Gamma}(o,2n,\Delta^{\prime}) where Δ:=2Δ+D\Delta^{\prime}:=2\Delta+D.

We shall prove that πg0\pi_{g_{0}} is uniformly finite to one. That is to say, there exists a constant N>1N>1 independent of g0g_{0} and gHg\in H so that πg01(g)\pi_{g_{0}}^{-1}(g) contains at most NN elements.

Indeed, if πg0(g)=πg0(g)\pi_{g_{0}}(g)=\pi_{g_{0}}(g^{\prime}) then g0gfp(g0gf)1=g0gfp(g0gf)1g_{0}gfp(g_{0}gf)^{-1}=g_{0}g^{\prime}fp(g_{0}g^{\prime}f)^{-1}. According to Lemma 5.8, the words (g0g,f,p,f1,(g0g)1)(g_{0}g,f,p,f^{-1},(g_{0}g)^{-1}) and (g0g,f,p,f1,(g0g)1)(g_{0}g^{\prime},f,p,f^{-1},(g_{0}g^{\prime})^{-1}) label respectively two (L,τ)(L,\tau)–admissible paths with the same endpoints. Thus, any geodesic α\alpha with the same endpoints rr–fellow travels them, so g0gog_{0}go and g0gog_{0}g^{\prime}o have a distance at most rr to α\alpha, where rr is given by Proposition 2.9. For g,gAΓ(g01o,n,Δ)g,g^{\prime}\in A_{\Gamma}(g_{0}^{-1}o,n,\Delta), we have |d(o,g0go)d(o,g0go)|2Δ|d(o,g_{0}go)-d(o,g_{0}g^{\prime}o)|\leq 2\Delta. Hence, the rr–closeness then implies that d(g0go,g0go)4r+2Δd(g_{0}go,g_{0}g^{\prime}o)\leq 4r+2\Delta. Thus, if d(go,go)d(go,g^{\prime}o) is larger than a constant R=R(r,Δ)R=R(r,\Delta) given by Lemma 2.12, we would obtain a contradiction. The finite number N:={go:d(o,go)R}N:=\sharp\{go:d(o,go)\leq R\} is hence the desired upper bound on the preimage of πg0\pi_{g_{0}}, which is clearly uniform independent of g0Γg_{0}\in\Gamma.

To conclude the proof, since ΓX\Gamma\curvearrowright\mathrm{X} is assumed to have purely exponential growth, we have AΓ(o,n,Δ)C0enωΓ\sharp A_{\Gamma}(o,n,\Delta)\geq C_{0}\mathrm{e}^{n\omega_{\Gamma}} for some C0,ΔC_{0},\Delta. Setting C=C0/NC=C_{0}/N, we have (gHg1AΓ(o,n,Δ))CenωΓ/2\sharp(gHg^{-1}\cap A_{\Gamma}(o,n,\Delta^{\prime}))\geq C\mathrm{e}^{n\omega_{\Gamma}/2}. The lemma is proved. ∎

An immediate corollary is as follows. If ΓX\Gamma\curvearrowright\mathrm{X} is SCC, this could be also derived from Theorems 1.11 and 1.10, for more general confined subgroups with compact confining subsets. See Corollary 1.12.

Corollary 8.3.

The Poincaré series associated to HH diverges at s=ωΓ/2:s=\omega_{\Gamma}/2:

𝒫H(ωΓ/2,o,o)=\mathcal{P}_{H}(\omega_{\Gamma}/2,o,o)=\infty

In particular, ωHωΓ/2\omega_{H}\geq\omega_{\Gamma}/2.

Here is another corollary from the proof. For any gGg\in G, there exists pgPp_{g}\in P such that gpgg1Hgp_{g}g^{-1}\in H. This defines a map as follows:

π1:gGgpgg1\pi_{1}:g\in G\longmapsto gp_{g}g^{-1}
Lemma 8.4.

There exists a subset KΓK\subseteq\Gamma, on which the above map π1\pi_{1} is injective, so that gKeωΓd(o,go)=\sum_{g\in K}\mathrm{e}^{-\omega_{\Gamma}d(o,go)}=\infty.

Proof.

The above defined map π1\pi_{1} is exactly the one πg0\pi_{g_{0}} for g0=1g_{0}=1 in the proof of Lemma 8.2. As πg0\pi_{g_{0}} is uniformly finite to one and ΓX\Gamma\curvearrowright\mathrm{X} is of divergent type, we can then find KK with the desired divergence property. ∎

Lemma 8.5.

Suppose that ωH=ωΓ/2\omega_{H}=\omega_{\Gamma}/2. Then for any s>ωHs>\omega_{H} and gΓg\in\Gamma, we have

hHesd(o,ghg1o)hHesd(o,ho)\sum_{h\in H}\mathrm{e}^{-sd(o,ghg^{-1}o)}\asymp\sum_{h\in H}\mathrm{e}^{-sd(o,ho)}

where the implicit constant does not depend on gg.

Proof.

We can write for any gΓg\in\Gamma and any s>ωHs>\omega_{H},

hHesd(o,ghg1o)n1(gHg1AΓ(o,n,Δ))esn\sum_{h\in H}\mathrm{e}^{-sd(o,ghg^{-1}o)}\asymp\sum_{n\geq 1}\sharp(gHg^{-1}\cap A_{\Gamma}(o,n,\Delta))\mathrm{e}^{-sn}

where the implicit constant depends on the width Δ\Delta of the annulus, but is independent of gg.

As HH contains contracting elements by Lemma 5.10, we have the upper bound by Lemma 8.1

CenωH(gHg1AΓ(o,n,Δ))CenωHC\mathrm{e}^{n\omega_{H}}\leq\sharp(gHg^{-1}\cap A_{\Gamma}(o,n,\Delta))\leq C^{\prime}\mathrm{e}^{n\omega_{H}}

where the lower bound is given by Lemma 8.2. The conclusion follows by substituting the growth estimates into the above series. ∎

8.1. Confined subgroups of divergence type

We say that a subset intersects trivially a subgroup if their intersection is contained in the trivial subgroup.

Proposition 8.6.

If a discrete group H<Isom(X)H<\mathrm{Isom}(\mathrm{X}) of divergence type is confined by Γ\Gamma with a finite confining subset PP that intersects trivially with E(Γ)E(\Gamma), then ωH>ωΓ/2\omega_{H}>\omega_{\Gamma}/2.

The proof presented here is similar to the proof of [72, Theorem 9.1(Case 1)], where HH is assumed to be normal in Γ\Gamma. We emphasize that HH is not necessarily to be contained in Γ\Gamma. We need some preparation before starting the proof.

Fix goΓogo\in\Gamma o. Let {μxgo}xX\{\mu_{x}^{go}\}_{x\in\mathrm{X}} be a PS measure on the horofunction boundary hX\partial_{h}{\mathrm{X}}, which is an accumulation point of {μxs,o}\{\mu_{x}^{s,o}\} in (18) supported on HgoHgo for a sequence sωHs\searrow\omega_{H}. This is a ωH\omega_{H}–dimensional HH–equivariant conformal density.

As HXH\curvearrowright\mathrm{X} is of divergence type, Lemma 2.37 implies that push forwarding {μxgo}xX\{\mu_{x}^{go}\}_{x\in\mathrm{X}} for different goΓogo\in\Gamma o almost gives the unique quasi-conformal density on the reduced horofunction boundary. That is to say, there are absolutely continuous with respect to each other, with uniformly bounded derivatives. Now, the Patterson’s construction gives

xX:μxs,go=limsωH+𝒫H(s,x,go)𝒫H(s,o,go),μxs,o=limsωH+𝒫H(s,x,o)𝒫H(s,o,o).\forall x\in\mathrm{X}:\quad\|\mu_{x}^{s,go}\|=\lim_{s\to\omega_{H}+}\frac{\mathcal{P}_{H}(s,x,go)}{\mathcal{P}_{H}(s,o,go)},\quad\|\mu_{x}^{s,o}\|=\lim_{s\to\omega_{H}+}\frac{\mathcal{P}_{H}(s,x,o)}{\mathcal{P}_{H}(s,o,o)}.

Recall that Hg=g1HgH^{g}=g^{-1}Hg. From the definition of Poincaré series (15), we have

s>ωH:𝒫H(s,go,go)𝒫H(s,o,go)=[𝒫H(s,go,o)𝒫Hg(s,o,o)]1\forall s>\omega_{H}:\quad\frac{\mathcal{P}_{H}(s,go,go)}{\mathcal{P}_{H}(s,o,go)}=\Bigg{[}\frac{\mathcal{P}_{H}(s,go,o)}{\mathcal{P}_{H^{g}}(s,o,o)}\Bigg{]}^{-1}

On account of Lemma 2.37, we fix in the following statement, a PS measure {μx}xX\{\mu_{x}\}_{x\in\mathrm{X}} on the horofunction boundary hX\partial_{h}{\mathrm{X}}. This is a limit point of {μxs,o}\{\mu_{x}^{s,o}\} in (18) supported on HoHo for a sequence sωHs\searrow\omega_{H}.

Lemma 8.7.

If ωH=ωΓ/2\omega_{H}=\omega_{\Gamma}/2, then there exists a constant M>0M>0 such that

(33) xΓo:μxM\forall x\in\Gamma o:\quad\|\mu_{x}\|\leq M
Proof.

By Lemma 8.5, if ωH=ωΓ/2\omega_{H}=\omega_{\Gamma}/2, then 𝒫Hg(s,o,o)𝒫H(s,o,o)\mathcal{P}_{H^{g}}(s,o,o)\asymp\mathcal{P}_{H}(s,o,o) for s>ωHs>\omega_{H}. Hence, we have

s>ωH:𝒫H(s,go,go)𝒫H(s,o,go)[𝒫H(s,go,o)𝒫H(s,o,o)]1\forall s>\omega_{H}:\quad\frac{\mathcal{P}_{H}(s,go,go)}{\mathcal{P}_{H}(s,o,go)}\asymp\Bigg{[}\frac{\mathcal{P}_{H}(s,go,o)}{\mathcal{P}_{H}(s,o,o)}\Bigg{]}^{-1}

where the implicit constant does not depend on gg.

For given goΓogo\in\Gamma o, let us take the same sequence of ss (depending on gogo) so that {μxs,go}xX\{\mu_{x}^{s,go}\}_{x\in\mathrm{X}} and {μxs,o}xX\{\mu_{x}^{s,o}\}_{x\in\mathrm{X}} converge to an HH–conformal density {μxgo}xX\{\mu_{x}^{go}\}_{x\in\mathrm{X}} and HH–conformal density {μxo}xX\{\mu_{x}^{o}\}_{x\in\mathrm{X}} respectively. The above relation then gives

(34) μgogoμgoo1\|\mu_{go}^{go}\|\asymp\|\mu_{go}^{o}\|^{-1}

Push forward the limiting measures {μxo}xX\{\mu_{x}^{o}\}_{x\in\mathrm{X}} and {μxgo}xX\{\mu_{x}^{go}\}_{x\in\mathrm{X}} to the quasi-conformal densities on the quotient [mH][\partial_{m}H], which we keep by the same notation. The constant λ\lambda in Definition 2.24 is universal for any {μxgo}xX\{\mu_{x}^{go}\}_{x\in\mathrm{X}} with goΓogo\in\Gamma o, as the difference of two horofunctions in the same locus of a Myrberg point is universal. Thus, Lemma 2.37 gives a constant M0=M(λ)M_{0}=M(\lambda) independent of goΓogo\in\Gamma o:

M1μgogoμgooMμgogoM^{-1}\|\mu_{go}^{go}\|\leq\|\mu_{go}^{o}\|\leq M\|\mu_{go}^{go}\|

so the relation (34) implies μgoo2M\|\mu_{go}^{o}\|^{2}\leq M.

Recall that {μx}\{\mu_{x}\} and {μxo}\{\mu_{x}^{o}\} may be different limit points of {μxs,o}xX\{\mu_{x}^{s,o}\}_{x\in\mathrm{X}}, where {μx}\{\mu_{x}\} is the PS measure fixed at the beginning of the proof. However, applying Lemma 2.37 again gives μxλμxo\|\mu_{x}\|\asymp_{\lambda}\|\mu_{x}^{o}\| for any xXx\in\mathrm{X} and thus for x=gox=go. Combinning with the above bound μgoo2M\|\mu_{go}^{o}\|^{2}\leq M, we obtain the desired upper bound on μgo\|\mu_{go}\| in (33) depending only on λ,M0\lambda,M_{0}. The proof is completed. ∎

We now recall the following result, which says that a boundary point lies in a uniformly finite many shadows from a fixed annulus.

Lemma 8.8.

[72, Lemma. 6.8] For given Δ>0\Delta>0, there exists N=N(Δ)N=N(\Delta) such that for every n1n\geq 1, any ξX\xi\in\partial{\mathrm{X}} is contained in at most NN shadows ΠoF(v,r)\Pi_{o}^{F}(v,r) where vAΓ(o,n,Δ)v\in A_{\Gamma}(o,n,\Delta).

We are ready to complete the proof of Proposition 8.6.

Proof of Proposition 8.6.

We assume ωH=ωΓ/2\omega_{H}=\omega_{\Gamma}/2 by way of contradiction, so Lemma 8.7 could apply. With the Shadow Principal 7.1, we obtain

μo(ΠoF(go,r))eωHd(o,go)\mu_{o}(\Pi_{o}^{F}(go,r))\asymp\mathrm{e}^{-\omega_{H}d(o,go)}

for any goΓogo\in\Gamma o. The same argument as in [72, Prop. 6.6] proves ωHωΓ\omega_{H}\geq\omega_{\Gamma}. Indeed, by Lemma 8.8, that every point ξhX\xi\in\partial_{h}{\mathrm{X}} are contained in at most N0N_{0} shadows for some uniform N0>0N_{0}>0. Hence,

vAΓ(o,n,Δ)μo(ΠoF(v,r))N0μo(hX)N0\sum_{v\in A_{\Gamma}(o,n,\Delta)}\mu_{o}(\Pi_{o}^{F}(v,r))\leq N_{0}\mu_{o}(\partial_{h}{\mathrm{X}})\leq N_{0}

which by the Shadow Lemma 2.31 gives AΓ(o,n,Δ)enωH\sharp A_{\Gamma}(o,n,\Delta)\prec\mathrm{e}^{n\omega_{H}}, and thus ωΓωH\omega_{\Gamma}\leq\omega_{H} gives a contradiction. ∎

8.2. Completion of proof of Theorem 1.13

First of all, the non-strict inequality ωHωG2\omega_{H}\geq\frac{\omega_{G}}{2} follows from Corollary 8.3. The difficulty is the strict part of this inequality.

If HXH\curvearrowright\mathrm{X} is of divergence type, then the conclusion follows by Proposition 8.6. If HXH\curvearrowright\mathrm{X} is of convergence type, then

hHeωGd(o,ho)/2<\sum_{h\in H}\mathrm{e}^{-\omega_{G}d(o,ho)/2}<\infty

and assume for the contradiction that ωH=ωG2\omega_{H}=\frac{\omega_{G}}{2}. Let KK be given by Lemma 8.4, so we obtain that

hHeωGd(o,ho)/2gKeωGd(o,gpgg1o)/2gKeωGd(o,go)\sum_{h\in H}\mathrm{e}^{-\omega_{G}d(o,ho)/2}\geq\sum_{g\in K}\mathrm{e}^{-\omega_{G}d(o,gp_{g}g^{-1}o)/2}\succ\sum_{g\in K}\mathrm{e}^{-\omega_{G}d(o,go)}

This contradicts the divergence action of GXG\curvearrowright\mathrm{X}. This completes the proof of Theorem 1.13.

9. Growth and co-growth Inequality

Under the setup of §7, we now prove Theorem 1.15 following closely Coulon’s argument in [19].

Again, HH is only assumed to be confined by Γ\Gamma, but not necessarily contained in Γ\Gamma. Denote Γ/H={Hg:gΓ}\Gamma/H=\{Hg:g\in\Gamma\} the collection of right HH–cosets with representatives in Γ\Gamma. Note that HH and HgHg may be not necessarily contained entirely in Γ\Gamma. Consider the Hilbert space =2(Γ/H)\mathcal{H}=\ell^{2}(\Gamma/H) with over the coset space Γ/H\Gamma/H.

For each t>0t>0, define

φt:Γ/H\displaystyle\varphi_{t}:\quad\Gamma/H\quad\longrightarrow \displaystyle\quad\mathbb{R}
Hg\displaystyle Hg\quad\longmapsto etd(Ho,Hgo)\displaystyle\quad\mathrm{e}^{-td(Ho,Hgo)}

Recall that ωΓ/H\omega_{\Gamma/H} is the convergence radius of the series

gΓetd(Ho,Hgo)\sum_{g\in\Gamma}\mathrm{e}^{-td(Ho,Hgo)}

Thus, if t>ωΓ/H/2t>\omega_{\Gamma/H}/2 then φt\varphi_{t}\in\mathcal{H}; if t<ωΓ/H/2t<\omega_{\Gamma/H}/2 then φt\varphi_{t}\notin\mathcal{H}. For each s>0s>0, define

ϕs:Γ/H\displaystyle\phi_{s}:\quad\Gamma/H\quad\longrightarrow \displaystyle\quad\mathbb{R}
Hg\displaystyle Hg\quad\longmapsto hHesd(o,hgo)\displaystyle\quad\sum_{h\in H}\mathrm{e}^{-sd(o,hgo)}

We now prove the following.

Lemma 9.1.

For any s>ωGs>\omega_{G}, we have ϕs\phi_{s}\in\mathcal{H}.

Proof.

Consider the PS-measure μxs\mu_{x}^{s} for s>ωHs>\omega_{H} supported on HoHo:

μxs=1𝒫H(s,o,o)hHesd(x,ho)Dirac(ho).\mu^{s}_{x}=\frac{1}{\mathcal{P}_{H}(s,o,o)}\sum\limits_{h\in H}e^{-sd(x,ho)}\cdot{\mbox{Dirac}}{(ho)}.

We compute the norm

(35) ϕs2\displaystyle\|\phi_{s}\|^{2} =Hg1=Hg2g1,g2Γes(d(o,g1o)+d(o,g2o))\displaystyle=\sum_{\underset{g_{1},g_{2}\in\Gamma}{Hg_{1}=Hg_{2}}}\mathrm{e}^{-s(d(o,g_{1}o)+d(o,g_{2}o))}
=gΓesd(o,go)(hHesd(go,ho))\displaystyle=\sum_{g\in\Gamma}\mathrm{e}^{-sd(o,go)}\left(\sum_{h\in H}\mathrm{e}^{-sd(go,ho)}\right)
=𝒫G(s,o,o)gΓesd(o,go)μgos\displaystyle=\mathcal{P}_{G}(s,o,o)\sum_{g\in\Gamma}\mathrm{e}^{-sd(o,go)}\|\mu_{go}^{s}\|

By Shadow Principle 7.3, we have

μgosesd(o,go)μos(ΩoF(go,r))\|\mu_{go}^{s}\|\mathrm{e}^{-sd(o,go)}\prec\mu_{o}^{s}(\Omega_{o}^{F}(go,r))

Given n1n\geq 1, any element in the following set

Ω(o,n)={hG:d(o,ho)>n}\Omega(o,n)=\{h\in G:d(o,ho)>n\}

is contained at most NN members from the family of cones {ΩoF(go,r):gAΓ(o,n,Δ)}\{\Omega_{o}^{F}(go,r):g\in A_{\Gamma}(o,n,\Delta)\}, where NN is a uniform number depending on r,Δr,\Delta.

It follows that

gA(o,n,Δ)μgosesd(o,go)gA(o,n,Δ)μos(ΩoF(go,r))μos(Ω(o,n))\sum_{g\in A(o,n,\Delta)}\|\mu_{go}^{s}\|\mathrm{e}^{-sd(o,go)}\prec\sum_{g\in A(o,n,\Delta)}\mu_{o}^{s}(\Omega_{o}^{F}(go,r))\prec\mu_{o}^{s}(\Omega(o,n))

From (35), we thus obtain

ϕs2\displaystyle\|\phi_{s}\|^{2} n1(hΩ(o,n)esd(o,ho))\displaystyle\prec\sum_{n\geq 1}\left(\sum_{h\in\Omega(o,n)}\mathrm{e}^{-sd(o,ho)}\right)
hΩ(o,n)d(o,ho)esd(o,ho).\displaystyle\prec\sum_{h\in\Omega(o,n)}d(o,ho)\mathrm{e}^{-sd(o,ho)}.

The last term is the derivative of the Poincaré series 𝒫H(s,o,o)\mathcal{P}_{H}(s,o,o), so is finite for s>ωHs>\omega_{H}. ∎

Let s>ωHs>\omega_{H} and t>ωΓ/H/2t>\omega_{\Gamma/H}/2. Recall that 𝒫Γ(s+t,o,o)=gΓe(s+t)d(o,go)\mathcal{P}_{\Gamma}(s+t,o,o)=\sum_{g\in\Gamma}\mathrm{e}^{-(s+t)d(o,go)}. Note that Γ\Gamma may be properly contained in the union of HgHg over gΓg\in\Gamma. Summing up the elements of Γ\Gamma in the same HgHg and noting d(Ho,Hgo)d(o,Hgo)d(Ho,Hgo)\leq d(o,Hgo), we have

𝒫Γ(s+t,o,o)HgΓ/H(esd(Ho,Hgo)hHetd(o,hgo))\mathcal{P}_{\Gamma}(s+t,o,o)\leq\sum_{Hg\in\Gamma/H}\left(\mathrm{e}^{-sd(Ho,Hgo)}\sum_{h\in H}\mathrm{e}^{-td(o,hgo)}\right)

The Cauchy-Schwartz inequality gives the finiteness on the scalar product of ϕs\phi_{s} and φt\varphi_{t}:

(ϕs,φr)\displaystyle(\phi_{s},\varphi_{r}) =HgΓ/H(etd(Ho,Hgo)hGetd(o,hgo))\displaystyle=\sum_{Hg\in\Gamma/H}\left(\mathrm{e}^{-td(Ho,Hgo)}\sum_{h\in G}\mathrm{e}^{-td(o,hgo)}\right)
ϕs2φt2<\displaystyle\leq\|\phi_{s}\|^{2}\|\varphi_{t}\|^{2}<\infty

This implies that s+tωHs+t\geq\omega_{H} and thus ωH+ωΓ/H/2ωΓ\omega_{H}+\omega_{\Gamma/H}/2\geq\omega_{\Gamma}. Theorem 1.15 is proved.

Part III More about Hopf decomposition, and a relation with quotient growth

10. Characterization of maximal quotient growth

In this section, we relate the conservative/dissipative boundary actions to the growth of quotient spaces. The main result is a dichotomy of quotient growth for any subgroup in a hyperbolic group, where the slower growth is equivalent to the conservative action on the Gromov boundary.

We start with a general development on the Dirichlet domain and its relation with small horospheric limit set.

10.1. Dirichlet domain and small horospheric limit set

The construction of Dirichlet domain tessellates the real hyperbolic spaces into convex polyhedrons (also known as Voronoi tessellation), so provides an important tool to study Kleinian groups with Poincaré polyhedron theorem. In general, the Dirichlet domain can fail to be convex even in other symmetric spaces. The construction of Dirichlet domain is general and relies purely on the metric; in particular, it can be discussed in the setting of coarse metric geometry. Notably, the Nielsen spanning tree is a Dirichlet domain in disguise in the work on free groups of [35]. We first give some variant of the Dirichlet domain in our setting, which relates to the small horospheric limit set (Definition 2.46). This is a key tool for analyzing the quotient growth.

Assume that HH acts properly on a proper geodesic space X\mathrm{X} with a contracting element. Fix a basepoint oXo\in\mathrm{X}. Without loss of generality, we may assume that the stabilizer of oo in HH is trivial: as we can always embed isometrically X\mathrm{X} into a larger one (say, attaching a cone at a point with nontrivial stabilizer to make the finite group action free on the base).

Given a (possibly negative) real number RR, let 𝐃R(o)\mathbf{D}_{R}(o) be the set of points xXx\in\mathrm{X} that are RR–closer to oo than any point in HoHo. Namely, x𝐃R(o)x\in\mathbf{D}_{R}(o) if and only if d(x,o)d(x,Ho)+Rd(x,o)\leq d(x,Ho)+R. Equivalently, 𝐃R(o)=1hH𝐇R(o,ho)\mathbf{D}_{R}(o)=\cap_{1\neq h\in H}\overleftarrow{\mathbf{H}}_{R}(o,ho) is a countable intersection of RR–half spaces defined as follows

𝐇R(x,y):={zX:d(z,x)d(z,y)+R}\overleftarrow{\mathbf{H}}_{R}(x,y):=\{z\in\mathrm{X}:d(z,x)\leq d(z,y)+R\}

This forms a locally finite family of closed sets (via the same proof of [59, Theorem 6.6.13]), so 𝐃R(o)\mathbf{D}_{R}(o) is a closed subset. It is obvious that 𝐃R(o)𝐃R(o)\mathbf{D}_{R}(o)\subseteq\mathbf{D}_{R^{\prime}}(o) for RRR\leq R^{\prime}, and 𝐃R(o)𝐃0(o)𝐃R(o)\mathbf{D}_{-R}(o)\subseteq\mathbf{D}_{0}(o)\subseteq\mathbf{D}_{R}(o) for R>0R>0. See Fig. 8 for an illustration of these notions.

We remark that the introduction of 𝐃R(o)\mathbf{D}_{R}(o) with negative RR is an essential novelty here, which is crucial in the Claim 2 in proof of Theorem 10.13.

Dirichlet domain

For R=0R=0, 𝐃(o):=𝐃0(o)\mathbf{D}(o):=\mathbf{D}_{0}(o) is the so-called Dirichlet domain centered at oo for the action HXH\curvearrowright\mathrm{X}. It is a fundamental domain in the following sense: H𝐃(o)=XH\cdot\mathbf{D}(o)=\mathrm{X} and hInt(𝐃(o))Int(𝐃(o))=h\cdot\textrm{Int}(\mathbf{D}(o))\cap\textrm{Int}(\mathbf{D}(o))=\emptyset for any h1h\neq 1. Denote by π:XX/H\pi:\mathrm{X}\to\mathrm{X}/H the quotient map, whose quotient topology is induced by the metric d¯\bar{d}. By [59, Theorem 6.6.7], X/H\mathrm{X}/H is homeomorphic to π(𝐃(o))\pi(\mathbf{D}(o)), where the latter is equipped with quotient topology via the restriction π:𝐃(o)π(𝐃(o))\pi:\mathbf{D}(o)\to\pi(\mathbf{D}(o)). See [59, Ch. 6] for relevant discussion.

In a real hyperbolic space, the notion of a bisector is useful in analyzing the Dirichlet domain, as it intersects the convex polyhedron 𝐃(o)\mathbf{D}(o) in faces. We here adopt a more general version of bisectors. For R0R\geq 0, set Bis(x,y;R):={zX:|d(z,x)d(z,y)|R}\mathrm{Bis}(x,y;R):=\{z\in\mathrm{X}:|d(z,x)-d(z,y)|\leq R\}, and then 𝐇R(x,y)=𝐇0(x,y)Bis(x,y;R)\overleftarrow{\mathbf{H}}_{R}(x,y)=\overleftarrow{\mathbf{H}}_{0}(x,y)\cup\mathrm{Bis}(x,y;R).

Examples 10.1.

Here are two examples to clarify the notion of bisectors.

  1. (1)

    In trees, it is readily checked that 𝐇R(x,y)\overleftarrow{\mathbf{H}}_{R}(x,y) is not contained in NT(𝐇0(x,y))N_{T}(\overleftarrow{\mathbf{H}}_{0}(x,y)) for any T>0T>0. Indeed, the limit set of the former set properly contains that of the latter set in the end boundary.

  2. (2)

    In a simply connected negatively pinched Riemannian manifold, it can be shown that Bis(x,y;R)\mathrm{Bis}(x,y;R) has finite Hausdorff distance (depending on RR) to Bis(x,y;0)\mathrm{Bis}(x,y;0) (using crucially the lower bound of curvature, i.e. the fatness of the comparison triangle). It follows that Bis(x,y;R)\mathrm{Bis}(x,y;R) is quasiconvex and its limit set remains the same for any R0R\geq 0.

ooh1oh^{-1}ohoho𝐃R(o)\mathbf{D}_{R}(o)𝐃R(o)\mathbf{D}_{-R}(o)𝐃0(o)\mathbf{D}_{0}(o)𝐇R(o,h1o)\overleftarrow{\mathbf{H}}_{-R}\left(o,h^{-1}o\right)𝐇R(o,ho)\overleftarrow{\mathbf{H}}_{R}(o,ho)Bis(o,ho;R)\mathrm{Bis}(o,ho;R)
Figure 8. Illustrating RR–half spaces for possible RR and their resulted Dirichlet domains

Let S+RS^{+R} denote the closed RR–neighborhood of a subset SS in X\mathrm{X}.

Lemma 10.2.

The following holds for any real number RR\in\mathbb{R}.

  1. (1)

    𝐃R(o)\mathbf{D}_{R}(o) is star-shaped at oo: any geodesic [o,x][o,x] for x𝐃R(o)x\in\mathbf{D}_{R}(o) is contained in 𝐃R(o)\mathbf{D}_{R}(o).

  2. (2)

    If R0R\geq 0, then 𝐃+R(o))𝐃2R(o)\mathbf{D}^{+R}(o))\subseteq\mathbf{D}_{2R}(o).

Proof.

(1). For any x𝐃R(o)x\in\mathbf{D}_{R}(o), we wish to prove that any geodesic [x,o][x,o] is contained in 𝐃R(o)\mathbf{D}_{R}(o). Indeed, if y(x,o)y\in(x,o) is not in 𝐃R(o)\mathbf{D}_{R}(o), there exists hHh\in H so that d(y,ho)<d(y,o)+Rd(y,ho)<d(y,o)+R. This implies d(x,ho)d(x,y)+d(y,ho)<d(x,o)+Rd(x,ho)\leq d(x,y)+d(y,ho)<d(x,o)+R, a contradiction with x𝐃R(o)x\in\mathbf{D}_{R}(o).

(2). If x𝐃+R(o)x\in\mathbf{D}^{+R}(o), then for some y𝐃(o)y\in\mathbf{D}(o), we have d(x,y)Rd(x,y)\leq R and d(y,o)d(y,Ho)d(y,o)\leq d(y,Ho). So d(x,o)d(y,Ho)+Rd(x,Ho)+2Rd(x,o)\leq d(y,Ho)+R\leq d(x,Ho)+2R, i.e.: x𝐃2R(o)x\in\mathbf{D}_{2R}(o). ∎

We now fix a convergence boundary X\partial{\mathrm{X}} for X\mathrm{X}, denote by Λ(S)\Lambda(S) the set of accumulations points of SS in X\partial{\mathrm{X}}. By Definition 2.13(B), we have [Λ(S+R)]=[Λ(S)][\Lambda(S^{+R})]=[\Lambda(S)].

The main result of this subsection is the following.

Theorem 10.3.

For any large R>0R>0,

  1. (1)

    [Λr,Fcon(Γo)]ΛHor(Ho)=H([Λ𝐃R(o)][Λr,Fcon(Γo)]).[\Lambda^{\textrm{con}}_{\textrm{r,F}}{(\Gamma o)}]\setminus\Lambda^{\mathrm{Hor}}{(Ho)}=H\cdot\left([\Lambda\mathbf{D}_{R}(o)]\cap[\Lambda^{\textrm{con}}_{\textrm{r,F}}{(\Gamma o)}]\right).

  2. (2)

    Λhor(Ho)[Λ𝐃R(o)]=\Lambda^{\mathrm{hor}}{(Ho)}\cap[\Lambda\mathbf{D}_{R}(o)]=\emptyset.

It is worth noting the following much simplified statement in hyperbolic or CAT(0) spaces. We remark that the proof in this case could be quite short and straightforward (cf. [51, Cor. 2.14]). We recommend the readers to figure out the proof themselves, instead of reading the following one which deals mainly with general metric spaces.

Theorem 10.4.

Let X\mathrm{X} be Gromov hyperbolic with Gromov boundary X\partial{\mathrm{X}} or CAT(0) with visual boundary X\partial{\mathrm{X}}. Then for any large R>0R>0.

  1. (1)

    XΛHor(Ho)=HΛ𝐃R(o).\partial{\mathrm{X}}\setminus\Lambda^{\mathrm{Hor}}{(Ho)}=H\cdot\Lambda\mathbf{D}_{R}(o).

  2. (2)

    Λhor(Ho)Λ𝐃R(o)=\Lambda^{\mathrm{hor}}{(Ho)}\cap\Lambda\mathbf{D}_{R}(o)=\emptyset.

The proof is achieved by a series of elementary lemmas.

Recall that a point ξ\xi is visual if a geodesic ray γ\gamma originating at oo terminates at [ξ][\xi], i.e. all the accumulation points of γ\gamma are contained in [ξ][\xi]. We emphasize that ξ\xi may not be an accumulation point of γ\gamma, so it is necessary to make statements on [][\cdot]–classes rather than boundary points. Any (r,F)(r,F)–conical point is visual by Lemma 2.35.

Remark 10.5.

The assumption ξΛr,Fcon(Γo)\xi\in\Lambda^{\textrm{con}}_{\textrm{r,F}}{(\Gamma o)} in the next three results is used to guarantee the following. If xnξx_{n}\to\xi, then any limiting geodesic ray of (a subsequence of) [o,xn][o,x_{n}] accumulates into [ξ][\xi]. If X\mathrm{X} is hyperbolic or CAT(0), this fact holds for any ξX\xi\in\partial{\mathrm{X}} with X\partial{\mathrm{X}} being Gromov boundary or visual boundary. Therefore, Λr,Fcon(Γo)\Lambda^{\textrm{con}}_{\textrm{r,F}}{(\Gamma o)} could be replaced with X\partial{\mathrm{X}} in these setups.

Recall the definition of horoballs in (2.22). The assertion (1) in Theorem 10.3 is proved by the following.

Lemma 10.6.

There exists a constant R=R(F)>0R=R(F)>0 with the following property. Let ξ[Λr,Fcon(Γo)]ΛHor(Ho)\xi\in[\Lambda^{\textrm{con}}_{\textrm{r,F}}{(\Gamma o)}]\setminus\Lambda^{\mathrm{Hor}}{(Ho)} be not a big horospheric limit point. Then there exist some hHh\in H and a geodesic [ho,ξ][ho,\xi] which is contained in 𝐃R(ho)\mathbf{D}_{R}(ho).

Proof.

According to definition, if ξΛHor(Ho)\xi\notin\Lambda^{\mathrm{Hor}}{(Ho)}, some horoball centered at [ξ][\xi] contains only finitely many elements from HoHo, so L=min{B[ξ](ho,o):hH}L=\min\{B_{[\xi]}(ho,o):h\in H\} is a finite value. Moreover, the horoball ([ξ],L)\mathcal{HB}([\xi],L) contains some hoHoho\in Ho, but [ξ](ho,ho)0\mathcal{B}_{[\xi]}(ho,h^{\prime}o)\leq 0 for any hoHoh^{\prime}o\in Ho. Let CC be the common contracting constant for elements in FF. Note that [Λr,Fcon(Γo)][\Lambda^{\textrm{con}}_{\textrm{r,F}}{(\Gamma o)}] is a subset of 𝒞20Chor\mathcal{C}^{\mathrm{hor}}_{20C} by Lemma 2.35. For any sequence of xnξΛr,Fcon(Γo)x_{n}\to\xi\in\Lambda^{\textrm{con}}_{\textrm{r,F}}{(\Gamma o)}, we obtain from (8) for n0n\gg 0:

Bxn(hno,hno)=d(xn,ho)d(xn,ho)21CB_{x_{n}}(h_{n}o,h_{n}^{\prime}o)=d(x_{n},ho)-d(x_{n},h^{\prime}o)\leq 21C

That is, setting R=21CR=21C, xn𝐇R(ho,ho)x_{n}\in\overleftarrow{\mathbf{H}}_{R}(ho,h^{\prime}o) for any hHh^{\prime}\in H, so we obtain xn𝐃R(o)x_{n}\in\mathbf{D}_{R}(o) by definition. See Fig. 9. Now, take a geodesic ray γ\gamma starting at hoho and ending at [ξ][\xi] by Lemma 2.35 (cf. Remark 10.5). If xnx_{n} is taken on γ\gamma tending to [ξ][\xi], the star-sharpedness implies γ𝐃R(o)\gamma\subseteq\mathbf{D}_{R}(o) by Lemma 10.2. Thus ξ[Λ𝐃R(ho)]\xi\in[\Lambda\mathbf{D}_{R}(ho)]. ∎

ξ\xi(ξ,L)\mathcal{HB}(\xi,L)oohohohoHoh^{\prime}o\in Hoh1ξh^{-1}\xi𝐃R(o)\mathbf{D}_{R}(o)hhxnx_{n}Λ𝐃R(o)\Lambda\mathbf{D}_{R}(o)
Figure 9. Proof of Lemma 10.6

The following preparatory result is required in proving the assertion (2) in Theorem 10.3. This follows easily in hyperbolic or CAT(0) spaces, as explained in Remark 10.5.

Lemma 10.7.

Let ξ[Λ𝐃R(o)][Λr,Fcon(Γo)]\xi\in[\Lambda\mathbf{D}_{R}(o)]\cap[\Lambda^{\textrm{con}}_{\textrm{r,F}}{(\Gamma o)}] for some R>0R>0. Then 𝐃R(o)\mathbf{D}_{R}(o) contains a geodesic ray ending at [ξ][\xi] starting at oo.

Proof.

By [72, Lemma 4.4], such a geodesic ray exists for any ξΛr,Fcon(Γo)\xi\in\Lambda^{\textrm{con}}_{\textrm{r,F}}{(\Gamma o)}. We briefly recall the proof, and then explain γ\gamma can be taken into 𝐃R(o)\mathbf{D}_{R}(o) by using the star-shaped 𝐃R(o)\mathbf{D}_{R}(o). First, a conical point ξΛr,Fcon(Γo)\xi\in\Lambda^{\textrm{con}}_{\textrm{r,F}}{(\Gamma o)} by definition lies in infinitely many partial shadows ΠoF(gno,r)\Pi_{o}^{F}(g_{n}o,r) for gnΓg_{n}\in\Gamma. By the proof of [72, Lemma 4.4], we can find a sequence of ynΠoF(gno,r)y_{n}\in\Pi_{o}^{F}(g_{n}o,r) tending to ξ\xi, so that [o,yn][o,y_{n}] intersects NC(gnAx(f))N_{C}(g_{n}\mathrm{Ax}(f)) for a common fFf\in F in a diameter comparable with d(o,fo)d(o,fo) (as F=3\sharp F=3). By Ascoli-Arzela Lemma, the limiting geodesic ray of (a subsequence of) [o,yn][o,y_{n}] tends to [ξ][\xi] by Definition 2.13(B).

Now take xn𝐃R(o)x_{n}\in\mathbf{D}_{R}(o) tending to [ξ][\xi]. If xnΠoF(gno,r)x_{n}\in\Pi_{o}^{F}(g_{n}o,r), then [o,xn]𝐃R(o)[o,x_{n}]\subseteq\mathbf{D}_{R}(o) converges locally uniformly to a geodesic ray γ𝐃R(o)\gamma\subseteq\mathbf{D}_{R}(o) by Lemma 10.2. In the general case, if d(o,fo)d(o,fo) is sufficiently large, the CC–contracting property of gnAx(f)g_{n}\mathrm{Ax}(f) implies that [o,xn][o,x_{n}] intersects NC(gnAx(f))N_{C}(g_{n}\mathrm{Ax}(f)) in a diameter comparable with d(o,fo)d(o,fo). Hence, a subsequence of [o,xn][o,x_{n}] converges locally uniformly to a geodesic ray α\alpha ending at [ξ][\xi]. As we wanted, α\alpha is contained in the star-shaped DR(o)D_{R}(o). ∎

We now prove the assertion (2) in Theorem 10.3

Lemma 10.8.

Let ξ[Λr,Fcon(Γo)][Λ𝐃R(o)]\xi\in[\Lambda^{\textrm{con}}_{\textrm{r,F}}{(\Gamma o)}]\cap[\Lambda\mathbf{D}_{R}(o)] for some R>0R>0. Then ξΛhor(Ho)\xi\notin\Lambda^{\mathrm{hor}}{(Ho)}.

Proof.

By Lemma 10.7, take a geodesic ray γ𝐃R(o)\gamma\subseteq\mathbf{D}_{R}(o) starting at oo and terminating at [ξ][\xi].

By way of contradiction, assume that ξΛhor(Ho)\xi\in\Lambda^{\mathrm{hor}}{(Ho)} is a small horospheric limit point. Hence, any horoball ([ξ])\mathcal{HB}([\xi]) centered at [ξ][\xi] contains a sequence of hnoGoh_{n}o\in Go tending to ξ\xi. That is, Bξ(hno,o)B_{\xi}(h_{n}o,o)\to-\infty as nn\to\infty. Note that the Busemann cocycle associated to γ\gamma differs from B[ξ](hno,o)B_{[\xi]}(h_{n}o,o) up to a uniform error. This implies that d(x,hno)d(x,o)d(x,h_{n}o)-d(x,o)\to-\infty as xγx\in\gamma tends to [ξ][\xi]. Consequently, the definition of 𝐃R(o)\mathbf{D}_{R}(o) shows that x𝐃R(o)x\notin\mathbf{D}_{R}(o) for any fixed RR. This contradicts the choice of γ\gamma in 𝐃R(o)\mathbf{D}_{R}(o). ∎

Finally, the proof of Theorem 10.4 follows from Lemma 10.6 and Lemma 10.8, where [Λr,Fcon(Γo)][\Lambda^{\textrm{con}}_{\textrm{r,F}}{(\Gamma o)}] could be replaced with X\partial{\mathrm{X}} by Remark 10.5.

10.2. Big and small horospheric limit set

In this and next subsections, assume that

  • The proper action ΓX\Gamma\curvearrowright\mathrm{X} is cocompact on a proper hyperbolic space (X,d)(\mathrm{X},d), which is compactified with Gromov boundary X\partial{\mathrm{X}} endowed with maximal partition (i.e. each [][\cdot]-class is singleton, so [][] is omitted in what follows).

  • Let {μx:xX}\{\mu_{x}:x\in\mathrm{X}\} be the family of PS measures on Λ(Γo)=X\Lambda(\Gamma o)=\partial{\mathrm{X}} constructed from ΓX\Gamma\curvearrowright\mathrm{X}. Then μo\mu_{o} is a doubling measure without atoms; indeed it is the Hausdorff measure with respect to visual metric at the correct dimension (see [18]). Any subgroup H<Isom(X)H<\mathrm{Isom}(\mathrm{X}) preserves the measure class of μo\mu_{o} by Lemma 4.1.

Let a countable group HH (possibly not contained in Γ\Gamma) act properly on X\mathrm{X}. We shall prove the following by elaborating on the proof of Sullivan [65] in Kleinian groups (see also [51, Lemma 5.7]).

Theorem 10.9.

μo(ΛHor(Ho)Λhor(Ho))=0\mu_{o}(\Lambda^{\mathrm{Hor}}{(Ho)}\setminus\Lambda^{\mathrm{hor}}{(Ho)})=0.

The key part is to prove that the set of Garnett points is μo\mu_{o}–null. Recall that a point ξΛHor(Ho)\xi\in\Lambda^{\mathrm{Hor}}{(Ho)} is called Garnett point if there exists a unique horoball (ξ,L)\mathcal{HB}(\xi,L) (Definition 2.22) at ξ\xi so that

  • The interior of (ξ,L)\mathcal{HB}(\xi,L) contains no point from HoHo: Bξ(ho,o)LB_{\xi}(ho,o)\geq L for any hoHoho\in Ho;

  • any larger horoball contains infinitely many points of HoHo: Bξ(ho,o)[L,L+ϵ]B_{\xi}(ho,o)\in[L,L+\epsilon] for infinitely many hoHoho\in Ho and any ϵ<0\epsilon<0;

where Bξ(ho,o)=lim supzξ[d(z,ho)d(z,o)]B_{\xi}(ho,o)=\limsup_{z\to\xi}[d(z,ho)-d(z,o)]. Denote by 𝐆𝐚𝐫\mathbf{Gar} the set of Garnett points for HH.

We first recall a useful observation of Tukia [67, Appendix].

Lemma 10.10.

The set A=ΛHor(Ho)(Λhor(Ho)𝐆𝐚𝐫)A=\Lambda^{\mathrm{Hor}}{(Ho)}\setminus(\Lambda^{\mathrm{hor}}{(Ho)}\cup\mathbf{Gar}) is μo\mu_{o}–null.

Proof.

We include a short proof for completeness. Indeed, let ξΛHor(Ho)Λhor(Ho)\xi\in\Lambda^{\mathrm{Hor}}{(Ho)}\setminus\Lambda^{\mathrm{hor}}{(Ho)} be a non-Garnett point, so there exists a horoball (ξ)\mathcal{HB}(\xi) whose boundary contains a non-empty finite set of elements in HoHo but whose interior contains no point in HoHo. Consequently, AA is a disjoint union of measurable subsets indexed over finite subsets in HoHo, which are invariant under HH, so we could choose a measurable fundamental domain for HAH\curvearrowright A. By Lemma 2.48, ΛHor(Ho)\Lambda^{\mathrm{Hor}}{(Ho)} is a subset of the conservative part 𝐂𝐨𝐧𝐬\mathbf{Cons}, so μo(A)=0\mu_{o}(A)=0. ∎

To complete the proof of Theorem 10.9, it suffices to prove that 𝐆𝐚𝐫\mathbf{Gar} is a μo\mu_{o}–null set.

Lemma 10.11.

μo(𝐆𝐚𝐫)=0\mu_{o}(\mathbf{Gar})=0.

Proof.

By way of contradiction, assume that μo(𝐆𝐚𝐫)>0\mu_{o}(\mathbf{Gar})>0.

We define a height function 𝔥:𝐆𝐚𝐫0\mathfrak{h}:\mathbf{Gar}\to\mathbb{R}_{\geq 0}. Set 𝔥(ξ)=d(o,(ξ))\mathfrak{h}(\xi)=d(o,\mathcal{H}(\xi)), where (ξ)\mathcal{H}(\xi) is the unique horoball centered at ξ\xi whose interior does not contain any point of HoHo and any horoball centered at ξ\xi larger than that contains infinitely many points of hoho. We verify that 𝔥\mathfrak{h} is a measurable function. Indeed, if ξΛBis(x,y;ϵ)\xi\in\Lambda\mathrm{Bis}(x,y;\epsilon), then |Bξ(x,y)|ϵ|B_{\xi}(x,y)|\leq\epsilon; if |Bξ(x,y)|ϵ|B_{\xi}(x,y)|\leq\epsilon, then ξΛBis(x,y;ϵ+1/n)\xi\in\Lambda\mathrm{Bis}(x,y;\epsilon+1/n) for any n>1n>1. This implies that the set ξX\xi\in\partial{\mathrm{X}} with |Bξ(x,y)|ϵ|B_{\xi}(x,y)|\leq\epsilon is a countable intersection of closed subsets, so is Borel measurable. By definition, 𝐆𝐚𝐫\mathbf{Gar} is the limit supremum of such measurable sets. Moreover, the set of ξ𝐆𝐚𝐫\xi\in\mathbf{Gar} with 𝔥(ξ)(r1,r2)\mathfrak{h}(\xi)\in(r_{1},r_{2}) is measurable.

As μo\mu_{o} is a doubling measure, the Lebesgue density theorem holds: Let 𝐆\mathbf{G} be the μo\mu_{o}–full set of density points ξ𝐆𝐚𝐫\xi\in\mathbf{Gar}, so

  1. (1)

    𝔥:𝐆\mathfrak{h}:\mathbf{G}\to\mathbb{R} is approximately continuous: ξn𝐆ξ𝐆\xi_{n}\in\mathbf{G}\to\xi\in\mathbf{G} implies 𝔥(ξn)𝔥(ξ)\mathfrak{h}(\xi_{n})\to\mathfrak{h}(\xi);

  2. (2)

    for any sequence of metric balls BnXB_{n}\subseteq\partial{\mathrm{X}} centered at ξ𝐆\xi\in\mathbf{G} with radius tending to 0, we have

    μo(𝐆𝐚𝐫Bn)μo(Bn)1.\frac{\mu_{o}(\mathbf{Gar}\cap B_{n})}{\mu_{o}(B_{n})}\to 1.

As in [65], we are going to find a “forbidden” region CnBnC_{n}\subseteq B_{n}, consisting of non-Garnett points, with a definite μo\mu_{o}–measure. This contradicts the item (2). The co-compact action is crucial to apply Shadow Lemma 2.31: for any xXx\in\mathrm{X} and any fixed large constant rr, we have

(36) μo(Πo(x,r))reωΓd(o,x).\mu_{o}(\Pi_{o}(x,r))\asymp_{r}\mathrm{e}^{-\omega_{\Gamma}d(o,x)}.
hno(ζ,L)h_{n}o\in\mathcal{HB}(\zeta,L)oopnp_{n}qnq_{n}DDξ\xiζ\zetaCnC_{n}Bn:=Πo(pn,r)B_{n}:=\Pi_{o}(p_{n},r)Cn:=Πo(qn,r)C_{n}:=\Pi_{o}(q_{n},r)xx(ξ,L)\mathcal{HB}(\xi,L)Lξ-L_{\xi}Ln/2L_{n}/2
Figure 10. The proof of Lemma 10.11 where X\mathrm{X} is a tree and r=0r=0. The forbidden region CnC_{n} contains no Garnett point: a horoball at ζCn\zeta\in C_{n} containing xx must contain hnoh_{n}o

For ξ𝐆𝐚𝐫\xi\in\mathbf{Gar}, there exists a unique horoball (ξ,Lξ)\mathcal{HB}(\xi,L_{\xi}) for some real number LξL_{\xi}, so that Bξ(o,ho)>LξB_{\xi}(o,ho)>L_{\xi} for all hHh\in H, but Lξ+1>Bξ(o,hno)>LξL_{\xi}+1>B_{\xi}(o,h_{n}o)>L_{\xi} for infinitely many distinct hnoHoh_{n}o\in Ho. Let x(ξ,Lξ)x\in\mathcal{HB}(\xi,L_{\xi}) be a projection point of oo onto \mathcal{HB}, so 𝔥(ξ)=d(o,x)\mathfrak{h}(\xi)=d(o,x) by definition. As horoballs in hyperbolic spaces are quasi-convex, we have |Lξ𝔥(ξ)|100δ|L_{\xi}-\mathfrak{h}(\xi)|\leq 100\delta, and d(x,[o,hno])100δd(x,[o,h_{n}o])\leq 100\delta.

Denote Ln:=d(x,hno)L_{n}:=d(x,h_{n}o). Choose pn,qn[x,hno]p_{n},q_{n}\in[x,h_{n}o] with d(x,pn)=Ln/2d(x,p_{n})=L_{n}/2 and d(x,qn)=Ln/2+Dd(x,q_{n})=L_{n}/2+D for a large constant DrD\gg r. Set Bn:=Πo(pn,r)B_{n}:=\Pi_{o}(p_{n},r) and Cn:=Πo(qn,r)C_{n}:=\Pi_{o}(q_{n},r). See Fig. 10 for approximate tree configuration of these points.

As d(x,[o,hno])100δd(x,[o,h_{n}o])\leq 100\delta and |d(x,pn)d(x,qn)|D|d(x,p_{n})-d(x,q_{n})|\leq D, we see that |d(o,pn)d(o,qn)|D+200δ|d(o,p_{n})-d(o,q_{n})|\leq D+200\delta. Shadow estimates (36) implies the existence of a positive constant cc depending only on DD:

μo(Cn)cμo(Bn){\mu_{o}(C_{n})}\geq c{\mu_{o}(B_{n})}

We claim that for all n0n\gg 0, CnC_{n} consists of non-Garnett points up to a μo\mu_{o}-null set. If not, take a density point ζCn𝐆\zeta\in C_{n}\cap\mathbf{G}. Let (ζ,Lζ)\mathcal{HB}(\zeta,L_{\zeta}) be the unique horoball associated with ζ\zeta, so |Lζ𝔥(ζ)|100δ|L_{\zeta}-\mathfrak{h}(\zeta)|\leq 100\delta follows as above. Let y(ζ,Lζ)y\in\mathcal{HB}(\zeta,L_{\zeta}) so that d(o,y)=𝔥(ζ)d(o,y)=\mathfrak{h}(\zeta). Our goal is to prove (ζ,Lζ)\mathcal{HB}(\zeta,L_{\zeta}) must contain hnoh_{n}o, contradicting the definition of a Garnett point.

Indeed, the approximate continuity implies |𝔥(ξ)𝔥(ζ)|=|d(o,x)d(o,y)|1|\mathfrak{h}(\xi)-\mathfrak{h}(\zeta)|=|d(o,x)-d(o,y)|\leq 1 for ζCnξ\zeta\in C_{n}\to\xi. As ζΠo(qn,r)\zeta\in\Pi_{o}(q_{n},r), we have d(qn,[o,ζ))rd(q_{n},[o,\zeta))\leq r. If DrD\gg r, the thin-triangle inequality shows that d(x,[o,ζ])10δd(x,[o,\zeta])\leq 10\delta and thus d(x,y)100δd(x,y)\leq 100\delta. Now, by the choice of d(x,pn)=Ln/2d(x,p_{n})=L_{n}/2 and d(hno,qn)=Ln/2Dd(h_{n}o,q_{n})=L_{n}/2-D, we have d(y,qn)d(x,qn)d(x,y)Ln/2+D100δd(y,q_{n})\geq d(x,q_{n})-d(x,y)\geq L_{n}/2+D-100\delta, implying

d(hno,qn)Ln/2Dd(y,qn)2D+100δd(h_{n}o,q_{n})\leq L_{n}/2-D\leq d(y,q_{n})-2D+100\delta

Noting d(qn,z)rd(q_{n},z)\leq r for some z[o,ζ)z\in[o,\zeta), we choose DrD\gg r so that d(hno,z)d(y,z)100δd(h_{n}o,z)\leq d(y,z)-100\delta. Lemma 6.1 implies hno(ζ,Lζ)h_{n}o\in\mathcal{HB}(\zeta,L_{\zeta}). Hence, we conclude that Cn𝒢=C_{n}\cap\mathcal{G}=\emptyset. Recall that μo(Cn)cμo(Bn){\mu_{o}(C_{n})}\geq c{\mu_{o}(B_{n})}, this contradicts the property that ξ\xi is a density point as described in (2). The proof of the lemma is complete. ∎

We conclude this subsection with a remark on possible extensions of Theorem 10.9 to mapping class groups.

Remark 10.12.

Consider the (non-cocompact) action of the mapping class group Γ=𝐌𝐨𝐝(Σg)\Gamma=\mathbf{Mod}(\Sigma_{g}) on Teichmüller space X\mathrm{X} for g2g\geq 2. The proof of Lemma 10.11 could be adapted to its subgroup H<ΓH<\Gamma, provided that (36) holds for all xXx\in\mathrm{X}. (Hyperbolicity could be circumvented by using partial shadows). By the Shadow Lemma 2.31, it is known that (36) holds for a cocompact subset of points. By work of [7], the conformal density μx\mu_{x} gives the same mass to Thurston boundary 𝒫\mathscr{PMF}, and the volume growth in Teichmüller space does not depend on the basepoint. This provides positive evidence that (36) holds for all xXx\in\mathrm{X}.

10.3. Characterizing maximal quotient growth

Let B(o,n)XB(o,n)\subseteq\mathrm{X} and B¯(π(o),n)X/H\bar{B}(\pi(o),n)\subseteq\mathrm{X}/H denote respectively the dd–ball and d¯\bar{d}–ball at oo and π(o)\pi(o) of radius n0n\geq 0. Note that π(B(o,n))=B¯(π(o),n)\pi(B(o,n))=\bar{B}(\pi(o),n): HgoB(π(o),n)Hgo\in B(\pi(o),n) if and only if d(Ho,Hgo)nd(Ho,Hgo)\leq n. We say that X/H\mathrm{X}/H has slower growth than X\mathrm{X} if given any maximal RR–separated subset ZZ (resp. WW) of X\mathrm{X} (resp. X/H\mathrm{X}/H) for some R>0R>0, we have

WB(πo,n)ZB(o,n)0,as n.\frac{\sharp W\cap B(\pi o,n)}{\sharp Z\cap B(o,n)}\to 0,\;\text{as }n\to\infty.

This definition is independent of the choice of Z,WZ,W and RR, as different choice yields comparable ball growth function up to multiplicative constants. By the Shadow Lemma, the notion of slower growth in orbit points could be re-formulated using volume. For example, if X\mathrm{X} is a simply connected rank-1 manifold with a co-compact action, ZB(o,n)\sharp Z\cap B(o,n) is comparable to the volume of B(o,n)B(o,n). In a non-cocompact action, an important potential example could be given by mapping class groups, if the full shadow principle holds as stated in Remark 10.12.

The following result in Kleinian groups was first obtained in [65, Theorem VI] and later in free groups in [35, Theorem 4.1] and [8, Theorem 7]. Recall from the previous subsection, ΓX\Gamma\curvearrowright\mathrm{X} is assumed to be geometric action on a hyperbolic space, and μo\mu_{o} is the corresponding Hausdorff measure on the Gromov boundary X\partial{\mathrm{X}}.

Theorem 10.13.

Assume that HH acts properly on a proper hyperbolic space X\mathrm{X}. Then the small/big horospheric limit set is μo\mu_{o}–full, if and only if X/H\mathrm{X}/H grows slower than X\mathrm{X}. Moreover, μo(Λhor(Ho))<1\mu_{o}(\Lambda^{\mathrm{hor}}{(Ho)})<1 is equivalent to the following

{Hgo:d¯(Ho,Hgo)n,gΓ}eωΓn.\sharp\{Hgo:\bar{d}(Ho,Hgo)\leq n,g\in\Gamma\}\asymp\mathrm{e}^{\omega_{\Gamma}n}.

As an immediate corollary, we obtain a characterization of conservative actions.

Corollary 10.14.

Assume that HH is torsion-free. Then the action H(X,μo)H\curvearrowright(\partial{\mathrm{X}},\mu_{o}) is conservative if and only if the action HXH\curvearrowright\mathrm{X} grows slower than ΓX\Gamma\curvearrowright\mathrm{X}.

Proof of Theorem 10.13.

By Theorems 10.9 and 10.3, μo(Λhor(Ho))=μo(ΛHor(Ho))=1\mu_{o}(\Lambda^{\mathrm{hor}}{(Ho)})=\mu_{o}(\Lambda^{\mathrm{Hor}}{(Ho)})=1 if and only if μo(Λ𝐃R(o))=0\mu_{o}(\Lambda\mathbf{D}_{R}(o))=0 for any sufficiently large R>0R>0.

Let π:XX/H\pi:\mathrm{X}\to\mathrm{X}/H be the projection. Fix a basepoint oXo\in\mathrm{X} and consider the orbit A=ΓoA=\Gamma o. Then NR(A)=XN_{R}(A)=\mathrm{X} for some R>0R>0 as the action ΓX\Gamma\curvearrowright\mathrm{X} is co-compact by assumption. The image A¯=π(A)\bar{A}=\pi(A) is equipped with the metric d¯(Hx,Hy)=d(Hx,Hy)\bar{d}(Hx,Hy)=d(Hx,Hy) for x,yAx,y\in A, and write

AΓ/H(o,n,Δ)={Hgo:|d(go,Ho)n|Δ}A_{\Gamma/H}(o,n,\Delta)=\{Hgo:|d(go,Ho)-n|\leq\Delta\}

According to (3), we have

ωΓ/H=lim supnlogAΓ/H(o,n,Δ)n\omega_{\Gamma/H}=\limsup_{n\to\infty}\frac{\log A_{\Gamma/H}(o,n,\Delta)}{n}

Recall AΓ(o,n,Δ)={goΓo:|d(o,go)n|Δ}A_{\Gamma}(o,n,\Delta)=\{go\in\Gamma o:|d(o,go)-n|\leq\Delta\}. For each HgoAΓ/H(o,n,Δ)Hgo\in A_{\Gamma/H}(o,n,\Delta), choose goΓogo\in\Gamma o so that d(go,o)=d(Hgo,Ho)d(go,o)=d(Hgo,Ho). That is, gogo is a closest point in HgoHgo to oo, so gogo is an element in 𝐃(o)\mathbf{D}(o). Thus, every element in AΓ/H(o,n,Δ)A_{\Gamma/H}(o,n,\Delta) admits a (possibly non-unique) lift in the following intersection:

Bn:=AΓ(o,n,Δ)𝐃(o)B_{n}:=A_{\Gamma}(o,n,\Delta)\cap\mathbf{D}(o)

Thus, AΓ/H(o,n,Δ)Bn\sharp A_{\Gamma/H}(o,n,\Delta)\leq\sharp B_{n}.

Claim.

If μo(Λ𝐃R(o))=0\mu_{o}(\Lambda\mathbf{D}_{R}(o))=0, then Bn=o(eωΓn)\sharp B_{n}=o(\mathrm{e}^{\omega_{\Gamma}n}) and AΓ/H(o,n,Δ)=o(eωΓn)\sharp A_{\Gamma/H}(o,n,\Delta)=o(\mathrm{e}^{\omega_{\Gamma}n})

Proof of the Claim.

Define

Λn:=goBnΠoF(go,r).\Lambda_{n}:=\bigcup_{go\in B_{n}}\Pi_{o}^{F}(go,r).

Let Λ\Lambda be the limit supremum of Λn\Lambda_{n}. By construction, Λ𝐃R(o)\Lambda\mathbf{D}_{R}(o) coincides with Λ\Lambda. Indeed, any point ξΛ𝐃R(o)\xi\in\Lambda\mathbf{D}_{R}(o), there exists a geodesic ray in 𝐃R(o)\mathbf{D}_{R}(o) ending at ξ\xi. Then we can choose gnBng_{n}\in B_{n} so that d(gno,γ)Rd(g_{n}o,\gamma)\leq R. This implies Λ𝐃R(o)Λ\Lambda\mathbf{D}_{R}(o)\subseteq\Lambda. The converse inclusion follows from the definition of BnB_{n}. Hence, μo(Λn)0\mu_{o}(\Lambda_{n})\to 0 as nn\to\infty.

As the family of shadows ΠoF(x,r)\Pi_{o}^{F}(x,r) for xBnx\in B_{n} has uniformly bounded multiplicity, we have by the Shadow Lemma

μo(Λn)xBnμo(ΠoF(x,r))eωΓnBn\mu_{o}(\Lambda_{n})\asymp\sum_{x\in B_{n}}\mu_{o}(\Pi_{o}^{F}(x,r))\asymp\mathrm{e}^{-\omega_{\Gamma}n}\sharp B_{n}

where the implicit constant depends on rr. Hence, the claim follows. ∎

Claim.

If μo(Λ𝐃R(o))>0\mu_{o}(\Lambda\mathbf{D}_{R}(o))>0, then AΓ/H(o,n,Δ)eωΓn\sharp A_{\Gamma/H}(o,n,\Delta)\asymp\mathrm{e}^{\omega_{\Gamma}n}.

Proof of the Claim.

Recall that 𝐃(o)\mathbf{D}(o) is a fundamental domain for HXH\curvearrowright\mathrm{X}, so any point xXx\in\mathrm{X} admits at least one translate hxhx into 𝐃(o)\mathbf{D}(o). If h1xh2x𝐃(o)h_{1}x\neq h_{2}x\in\mathbf{D}(o), we have d(h1x,o)=d(h2x,o)d(h_{1}x,o)=d(h_{2}x,o) and then xx lies on the bisector Bis(h11o,h21o;0)\mathrm{Bis}(h_{1}^{-1}o,h_{2}^{-1}o;0).

Consider the facet-like set

F(h):=Bis(o,ho;R)𝐃R(o)F(h):=\mathrm{Bis}(o,ho;R)\cap\mathbf{D}_{R}(o)

and the set Λ(h)\Lambda(h) of accumulation points of F(h)F(h) in X\partial{\mathrm{X}}. By Theorem 10.3, the intersection Λ(h)ΛHor(Ho)\Lambda(h)\cap\Lambda^{\mathrm{Hor}}{(Ho)} has no small horospheric limit point, so is μo\mu_{o}–null by Theorem 10.9.

Note that 𝐃R(o)\mathbf{D}_{-R}(o) for R>0R>0 is contained in the interior of 𝐃(o)\mathbf{D}(o), where π:XX/H\pi:\mathrm{X}\to\mathrm{X}/H restricts as an injective map. See Fig. 8. By definition, Λ𝐃R(o)\Lambda\mathbf{D}_{R}(o) is the union of Λ𝐃R(o)\Lambda\mathbf{D}_{-R}(o) with countably many facet-like sets F(h)F(h), which are μo\mu_{o}-null, so we obtain μo(Λ𝐃R(o))=μo(Λ𝐃R(o))>0\mu_{o}(\Lambda\mathbf{D}_{-R}(o))=\mu_{o}(\Lambda\mathbf{D}_{R}(o))>0.

By injectivity of π:𝐃R(o)X/H\pi:\mathbf{D}_{-R}(o)\to\mathrm{X}/H restricted as map, B^n:=AΓ(o,n,Δ)𝐃R(o)\hat{B}_{n}:=A_{\Gamma}(o,n,\Delta)\cap\mathbf{D}_{-R}(o) injects into X/H\mathrm{X}/H. Consider similarly

Λ^n:=goB^nΠoF(go,r).\hat{\Lambda}_{n}:=\bigcup_{go\in\hat{B}_{n}}\Pi_{o}^{F}(go,r).

Let Λ~\tilde{\Lambda} be the limit of Λ~m:=mnΛ^m\tilde{\Lambda}_{m}:=\cup_{m\geq n}\hat{\Lambda}_{m}. As the above claim, the star-shapedness of 𝐃R(o)\mathbf{D}_{-R}(o) shows that Λ~=Λ𝐃R(o)\tilde{\Lambda}=\Lambda\mathbf{D}_{-R}(o), and Λ~m\tilde{\Lambda}_{m} is covered with a uniform multiplicity by the family of ΠoF(go,r)\Pi_{o}^{F}(go,r) for goB^n{go\in\hat{B}_{n}}.

Consequently, μo(B^n)>c\mu_{o}(\hat{B}_{n})>c for some uniform c>0c>0 independent of nn, and

μo(B^n)xB^nμo(Πo(x,r))eωΓnB^n\mu_{o}(\hat{B}_{n})\asymp\sum_{x\in\hat{B}_{n}}\mu_{o}(\Pi_{o}(x,r))\asymp\mathrm{e}^{-\omega_{\Gamma}n}\sharp\hat{B}_{n}

We thus obtain B^neωΓn\sharp\hat{B}_{n}\asymp\mathrm{e}^{\omega_{\Gamma}n}, and then the conclusion follows by B^nAΓ(o,n,Δ)\sharp\hat{B}_{n}\leq\sharp A_{\Gamma}(o,n,\Delta). ∎

The proof of the theorem is completed by the above two claims. ∎

The first claim actually does not make use of the assumption that ΓX\Gamma\curvearrowright\mathrm{X} is cocompact and μo\mu_{o} is doubling. These two assumptions are only required in Lemma 10.11. So we have.

Corollary 10.15.

Assume that ΓX\Gamma\curvearrowright\mathrm{X} is a proper action on a hyperbolic space and μx\mu_{x} be the ωΓ\omega_{\Gamma}–dimensional Γ\Gamma–equivariant quasi-conformal density on X\partial{\mathrm{X}}. Let HH be a subgroup in Γ\Gamma. If μo(Λhor(Ho))=0\mu_{o}(\Lambda^{\mathrm{hor}}{(Ho)})=0, then AΓ/H(o,n,Δ)=o(eωΓn)\sharp A_{\Gamma/H}(o,n,\Delta)=o(\mathrm{e}^{\omega_{\Gamma}n}).

11. Subgroups with nontrivial conservative and dissipative components

To complement the part 1, we construct in this section examples with non-trivial Hopf decomposition. Through this section, suppose that

  • Γ\Gamma admits a proper SCC action on X\mathrm{X} with contracting elements.

  • Let X\partial{\mathrm{X}} be a convergence boundary for X\mathrm{X} and {μx:xX}\{\mu_{x}:x\in\mathrm{X}\} be a quasi-conformal, Γ\Gamma–equivariant density of dimension ωΓ\omega_{\Gamma} on X\partial{\mathrm{X}}.

We are going to construct a free subgroup of infinite rank with nontrivial conservative and dissipative components. The construction is motivated by examples in free groups given in [35].

11.1. Preparation

Pick up any non-pinched contracting element fΓf\in\Gamma and form the contracting system ={gAx(f):gΓ}\mathscr{F}=\{g\mathrm{Ax}(f):g\in\Gamma\}, whose stabilizers are accordingly gE(f)g1gE(f)g^{-1}.

Projection complex

For a sufficiently large K>0K>0, the projection complex 𝒫K()\mathcal{P}_{K}(\mathscr{F}) defined by Bestvina-Bromberg-Fujiwara [9] is a quasi-tree of infinite diameter ([9]):

  • The vertex set consists of all elements in \mathscr{F}.

  • Two vertices uvu\neq v\in\mathscr{F} are connected by one edge if for any ww\in\mathscr{F} with uwvu\neq w\neq v, we have dw(u,v)>K\textbf{d}_{w}(u,v)>K. Here the projection is understood between w,u,vw,u,v\in\mathscr{F} as axes in X\mathrm{X}.

Remark 11.1.

Actually, the function “dw\textbf{d}_{w}” is a slight modification of dw(u,v)=πw(u)πw(v)\textbf{d}_{w}(u,v)=\|\pi_{w}(u)\cup\pi_{w}(v)\|, up to an additive amount. See [9, Def. 3.1] and [10, Sec. 4] for relevant discussions. We shall use the later one, which does not affect the estimates in practice.

Any two vertices uvu\neq v\in\mathscr{F} are connected by a standard path γ\gamma consisting of consecutive vertices

w0:=u,w1,,wn:=vw_{0}:=u,w_{1},\cdots,w_{n}:=v

where dwi(u,v)>K\textbf{d}_{w_{i}}(u,v)>K. See [9, Theorem 3.3]. This does not certifies the connectedness of 𝒫K()\mathcal{P}_{K}(\mathscr{F}), but also plays a crucial role in the whole theory. Among the useful facts, we mention

  1. (1)

    Standard paths are 22–quasi-geodesics ([10, Corollary 3.7]).

  2. (2)

    A triangle formed by three standard paths α,β,γ\alpha,\beta,\gamma is almost a tripod: γ\gamma is contained in the union αβ\alpha\cup\beta, with an exception of at most two consecutive vertices ([10, Lemma 3.6]).

By the rotating family theory, Dahmani-Guirardel-Osin [24] proved that

Theorem 11.2.

There exists a (not possibly every) large n>0n>0 so that the normal closure G:=\llanglefn\rrangleG:=\llangle f^{n}\rrangle is a free group of infinite rank.

We recall some ingredients in the proof using projection complex recently given in [11] inspired by Clay-Mangahas-Margalit [17].

Fix a basepoint v0:=Ax(f)v_{0}:=\mathrm{Ax}(f) at 𝒫K()\mathcal{P}_{K}(\mathscr{F}), so v0v_{0} is fixed by E(f)<ΓE(f)<\Gamma. By definition, GG is generated by all conjugates of fn\langle f^{n}\rangle in Γ\Gamma. If nn is chosen so fn\langle f^{n}\rangle is normal in E(f)E(f), then GG acts on 𝒫K()\mathcal{P}_{K}(\mathscr{F}) with the stabilizers exactly from those conjugates of fn\langle f^{n}\rangle. In particular, the stabilizer of v0v_{0} is now fn\langle f^{n}\rangle.

If such n0n\gg 0 is large enough, the action G𝒫K()G\curvearrowright\mathcal{P}_{K}(\mathscr{F}) is proved in [11, Thm 3.2] to be an LL–spinning family in sense of [17]:

(37) dv0(w,fnw)>L,v0w\textbf{d}_{v_{0}}(w,f^{n}w)>L,\;\forall v_{0}\neq w\in\mathscr{F}

We follow closely the account of [11] to explain.

Construction of free base

We do it inductively and start with G0=fnG_{0}=\langle f^{n}\rangle and S0={v0}S_{0}=\{v_{0}\}. Let N1N_{1} be the 11–neighborhood of W0:=S0W_{0}:=S_{0} and S1S_{1} be the set of vertices up to G0G_{0}–conjugacy in N1W0N_{1}\setminus W_{0}. Set G1=Gv:vS1S0G_{1}=\langle G_{v}:v\in S_{1}\cup S_{0}\rangle and W1=G1N1W_{1}=G_{1}N_{1}. Then S0S1S_{0}\cup S_{1} is a fundamental set for the action G1W10G_{1}\curvearrowright W_{1}^{0}, i.e. containing exactly one point from each orbit. Here W10W_{1}^{0} denotes the vertex set of W1W_{1}.

For i2i\geq 2, let Wi=GiNi1W_{i}=G_{i}\cdot N_{i-1} and define Ni+1N_{i+1} to be the 11–neighborhood of WiW_{i}. Let Si+1S_{i+1} be the set of vertices up to GiG_{i}–conjugacy in Ni+1WiN_{i+1}\setminus W_{i}, so that each vertex in Si+1S_{i+1} is adjacent to some in SiS_{i}. This can be done, as Wi0=Gi(0j<iSj)W_{i}^{0}=G_{i}\cdot(\cup_{0\leq j<i}S_{j}). Then Gi+1:=Gv:v0jiSjG_{i+1}:=\langle G_{v}:v\in\cup_{0\leq j\leq i}S_{j}\rangle is generated by the stabilizers of vertices in Ni+1N_{i+1}.

Canoeing path

Recall that KK is a large constant chosen for 𝒫K()\mathcal{P}_{K}(\mathscr{F}). Any two vertices u,vu,v in WiW_{i} can be connected by an LL–canoeing path γ\gamma for some L=L(K)>0L=L(K)>0 (see [11, Def. 4.3 & Lemma 4.9]):

  1. (1)

    γ\gamma is a concatenation γ1γ2γn\gamma_{1}\cdot\gamma_{2}\cdot\cdots\cdot\gamma_{n} where γi\gamma_{i} is a geodesic or the union of two geodesics.

  2. (2)

    The common endpoint vv of γi\gamma_{i} and γi+1\gamma_{i+1} has LL–large angle: if vγi,v+γi+1v_{-}\in\gamma_{i},v_{+}\in\gamma_{i+1} are previous and next vertices, then dv(v,v+)>L\textbf{d}_{v}(v_{-},v_{+})>L, where v,v,v+v,v_{-},v_{+}\in\mathscr{F} are understood as axes in X\mathrm{X}.

By [11, Prop. 4.4], LL–large angle points are contained in the standard path from uu to vv.

In [11], through the canoe path, it is inductively proven that Gi+1G_{i+1} is generated by GiG_{i} with a free group generated by SiS_{i}. Exhausting 𝒫K()\mathcal{P}_{K}(\mathscr{F}) by WiW_{i}, GG is obtained as the direct limit of GiG_{i}, which is a free group generated by stabilizers of vertices in S:=0j<SjS:=\cup_{0\leq j<\infty}S_{j}. Recall that G0=fnG_{0}=\langle f^{n}\rangle and S0={v0}S_{0}=\{v_{0}\}, and consider the splitting

G=G0HG=G_{0}\star H

where H:=Gv:vSS0H:=\langle G_{v}:v\in S\setminus S_{0}\rangle is a free group. We also need the associated Bass-Serre tree TT. More precisely, as the quotient graph T/GT/G is an interval, we partition T0=V1V2T^{0}=V_{1}\cup V_{2} with the stabilizers of vertices in V1V_{1} being conjugate to G0G_{0} and the stabilizers of vertices in V2V_{2} to HH. Moreover,

  • the set of G0G_{0}–vertices (resp. HH–vertices) in V1V_{1} (resp. V2V_{2}) are labeled by left G0G_{0}–cosets (resp. HH–cosets) in GG;

  • the set of edges adjacent to uV1u\in V_{1} (resp. uV2u\in V_{2}) are bijective to the elements in G0G_{0} (resp. HH).

11.2. Subgroups with nontrivial conservative part

We now state the main theorem in this section.

Theorem 11.3.

The subgroup HH has proper limit set [Λ(Ho)][Λ(Γo)][\Lambda(Ho)]\subsetneq[\Lambda(\Gamma o)]. Moreover, If ωG<ωΓ\omega_{G}<\omega_{\Gamma}, then μo(ΛHor(Ho))μo(Λhor(Ho))>0\mu_{o}(\Lambda^{\mathrm{Hor}}{(Ho)})\geq\mu_{o}(\Lambda^{\mathrm{hor}}{(Ho)})>0.

In particular, since HH has a proper limit set, it has maximal quotient growth by [36]. By Corollary 1.20, which will be proven in Section 10, Theorem 11.3 implies that the action of HH has non-trivial Hopf decomposition.

Remark 11.4.

In the construction, we have that Γ/G\Gamma/G is non-elementary acylindrically hyperbolic, in particular it is non-amenable. The amenability conjecture states that for a normal subgroup GG, ωG=ωΓ\omega_{G}=\omega_{\Gamma} if and only if the quotient group Γ/G\Gamma/G is amenable. If the amenability conjecture is verified for any SCC action, then the assumption that ωG<ωΓ\omega_{G}<\omega_{\Gamma} could be removed.

The remainder of this subsection is devoted to the proof of Theorem 11.3, which relies on the following decomposition of the limit set of GG.

The end boundary T\partial T of the Bass-Serre tree TT associated to G0HG_{0}\star H consists of all geodesic rays from a fixed point. As TT is locally infinite, it is non-compact. Note that [Λ(G0o)]=[f][f+][\Lambda(G_{0}o)]=[f^{-}]\cup[f^{+}] for G0=fnG_{0}=\langle f^{n}\rangle. In what follows, we understand [Λ(Γo)][\Lambda(\Gamma o)] as the quotient space by identifying each [][\cdot]–class as one point. Recall that 𝒞\mathcal{C} is the set of non-pinched points in X\partial{\mathrm{X}} (Definition 2.13(C)), and the points outside 𝒞\mathcal{C} are pinched.

Proposition 11.5.

We have the following covering

(38) [Λ(Go)]=[gGg[Λ(Ho)]][gGg[f±]]ι(T)[\Lambda(Go)]=\left[\bigcup_{g\in G}g[\Lambda(Ho)]\right]\bigcup\left[\bigcup_{g\in G}g[f^{\pm}]\right]\bigcup\iota(\partial T)

where a map ι:T[Λ(Γo)]\iota:\partial T\to[\Lambda(\Gamma o)] is defined below in Lemma 11.7. Moreover, this covering becomes a disjoint union if every term on RHS and LHS is replaced by its intersection with the set of non-pinched points 𝒞\mathcal{C}.

In a variety of examples, it is known that X=𝒞\partial{\mathrm{X}}=\mathcal{C} consists of non-pinched points, so the union in (38) becomes a disjoint union. We record this for further reference.

Corollary 11.6.

Suppose that (X,X)(\mathrm{X},\partial{\mathrm{X}}) is one of the items (1),(2),(3) and (5) in Examples 2.15. Then the following holds:

(39) [Λ(Go)]=[gGg[Λ(Ho)]][gGg[f±]]ι(T)[\Lambda(Go)]=\left[\bigsqcup_{g\in G}g[\Lambda(Ho)]\right]\bigsqcup\left[\bigsqcup_{g\in G}g[f^{\pm}]\right]\bigcup\iota(\partial T)

where a map ι:T[Λ(Γo)]\iota:\partial T\to[\Lambda(\Gamma o)] is defined below in Lemma 11.7.

As shown in (39), the first points of any contracting elements other than ff must be contained in some GG–translate of Λ(Ho)\Lambda(Ho).

In the next three steps, we shall describe in detail how a geodesic on the Bass-Serre tree TT produces a canoeing path in the projection complex 𝒫K()\mathcal{P}_{K}(\mathscr{F}), which is then lifted to an admissible path in the space X\mathrm{X}. This is the crux of the proof of Proposition 11.5.

Let UU be a non-empty reduced word over the alphabet (HG0)1(H\cup G_{0})\setminus 1. We decompose UU into nonempty subwords UiU_{i} separated by letters in G0=fnG_{0}=\langle f^{n}\rangle:

(40) U=(h1,fn1,h2,fn2,,hk,fnk)\displaystyle U=(h_{1},f^{n_{1}},h_{2},f^{n_{2}},\cdots,h_{k},f^{n_{k}})

where hiHh_{i}\in H and n|ni0n|n_{i}\neq 0 unless the last one i=ki=k. For each hih_{i}, we have a minimal j=j(Ui)j=j(U_{i}) so that hih_{i} represents an element in GjG0HG_{j}\setminus G_{0}\subseteq H. Then the element g=h1fn1hkfnkg=h_{1}f^{n_{1}}\cdots h_{k}f^{n_{k}} is represented by UU. If gg is not in HH, then k2k\geq 2 and n10n_{1}\neq 0.

Fix an oriented edge e=[𝐯0,𝐯1]e=[\mathbf{v}_{0},\mathbf{v}_{1}] in TT with endpoint stabilizers G0G_{0} and HH respectively. Denote e=[𝐯1,𝐯0]\overleftarrow{e}=[\mathbf{v}_{1},\mathbf{v}_{0}] the reversed edge. We consider the basepoint 𝐯0\mathbf{v}_{0} corresponding to the coset G0G_{0} in TT, and the basepoint v0=Ax(f)v_{0}=\mathrm{Ax}(f) in 𝒫K()\mathcal{P}_{K}(\mathscr{F}) stabilized by G0G_{0}.

eeh1Hh_{1}\in Hh1eh_{1}eG0G_{0}fn1f^{n_{1}}fn1ef^{n_{1}}eh2h_{2}h2eh_{2}efnk1ef^{n_{k-1}}ehkh_{k}hkeh_{k}efnkf^{n_{k}}𝐯0\mathbf{v}_{0}𝐯1\mathbf{v}_{1}g𝐯0g\mathbf{v}_{0}
Figure 11. Unfolding the word UU as a geodesic α\alpha in Bass-Serre tree TT. The reader should be cautioned that the labels indicate the action of stabilizers on vertices understood up to appropriate conjugation.
  1. Step 1

    The word UU gets unfolded as a geodesic α\alpha in TT starting with the edge ee, at which h1Hh_{1}\in H rotates the edge ee at 𝐯1\mathbf{v}_{1} to h1eh_{1}e. Rotating edges consecutively in this way, we write down

    α=e(h1e)(h1fn1e)(h1fn1h2e)(h1fn1hke)\alpha=e\cdot(h_{1}\overleftarrow{e})\cdot(h_{1}f^{n_{1}}e)\cdot(h_{1}f^{n_{1}}h_{2}\overleftarrow{e})\cdot\cdots\cdot(h_{1}f^{n_{1}}\cdots h_{k}\overleftarrow{e})

    where the last edge terminates at the endpoint fixed by corresponding conjugate of fnkf^{n_{k}}. See Fig. 11.

  2. Step 2

    Recall Wj(i)=Gj(i)Nj(i)1W_{j(i)}=G_{j(i)}\cdot N_{j(i)-1} and hiGj(i)h_{i}\in G_{j(i)}. As mentioned above, there exists an LL–canoeing path γi\gamma_{i} in Wj(i)W_{j(i)} from v0v_{0} to hiv0h_{i}v_{0}. These hi1γih_{i}^{-1}\gamma_{i} and γi+1\gamma_{i+1} are adjacent at v0v_{0}, around which fnif^{n_{i}} rotates γi+1\gamma_{i+1}. By (37), dv0(v,fniv)>Ld_{v_{0}}(v,f^{n_{i}}v)>L for any vv0v\neq v_{0}, so hiv0h_{i}v_{0} is a LL–large angle point of the concatenation γi(hifniγi+1)\gamma_{i}\cdot(h_{i}f^{n_{i}}\gamma_{i+1}). We then obtain an LL–canoeing path γ\gamma from v0v_{0} to gv0gv_{0} as the concatenation of γi\gamma_{i} (1in)(1\leq i\leq n) up to appropriate translation,

    γ=γ1(h1fn1γ2)(h1fn1h2fn2γ3)(h1fn1hk1fnk1γk)\gamma=\gamma_{1}\cdot(h_{1}f^{n_{1}}\gamma_{2})\cdot(h_{1}f^{n_{1}}h_{2}f^{n_{2}}\gamma_{3})\cdot\cdots\cdot(h_{1}f^{n_{1}}\cdots h_{k-1}f^{n_{k-1}}\gamma_{k})

    where the common endpoints of two consecutive paths are LL–large angle points fixed by the corresponding conjugates of fnif^{n_{i}}. See Fig. 12.

    γ1\gamma_{1}γ2\gamma_{2}fn1f^{n_{1}}v0v_{0}h1v0h_{1}v_{0}h1fn1h2v0h_{1}f^{n_{1}}h_{2}v_{0}γk\gamma_{k}gv0gv_{0}h1fn1hk1v0h_{1}f^{n_{1}}\cdots h_{k-1}v_{0}fnk1f^{n_{k-1}}
    Figure 12. Canoeing path γ\gamma in 𝒫K()\mathcal{P}_{K}(\mathscr{F}): join canoeing paths γi\gamma_{i} at large angle points h1fn1hifniv0h_{1}f^{n_{1}}\cdots h_{i}f^{n_{i}}v_{0}. Those large points are contained in the standard path β\beta (possibly βγ\beta\neq\gamma) from v0v_{0} to gv0gv_{0}.
  3. Step 3

    By [11, Prop. 4.4], those LL–large angle points lie on the standard path β\beta from v0v_{0} to gv0gv_{0}. We lift the standard path β\beta to get an (L^,τ)(\hat{L},\tau)–admissible path β~\tilde{\beta} in X\mathrm{X} as follows, with contracting sets given by the axis wi:=h1fn1hiAx(f)w_{i}:=h_{1}f^{n_{1}}\cdots h_{i}\mathrm{Ax}(f) where w0=Ax(f),wk+1=gAx(f)w_{0}=\mathrm{Ax}(f),w_{k+1}=g\mathrm{Ax}(f):

    Choose two points ow0o\in w_{0} and gowk+1go\in w_{k+1} in X\mathrm{X}. Connect o,goo,go by concatenating geodesics in consecutive axis wiw_{i} and then geodesics from πwi1(wi)\pi_{w_{i-1}}(w_{i}) to πwi(wi1)\pi_{w_{i}}(w_{i-1}). See a schematic illustration (13) of a lifted path.

    By [72, Lemma 2.22], this gives an (L^,τ)(\hat{L},\tau)–admissible path from oo to gogo with associated contracting axes wiw_{i}’s. The constant L^\hat{L} is comparable with LL and τ\tau depends only on \mathscr{F}. See [72] for details.

w0=v0w_{0}=v_{0}wiw_{i}wi+1w_{i+1}wk+1=gv0w_{k+1}=gv_{0}πwi(wi+1)\pi_{w_{i}}(w_{i+1})πwi+1(wi)\pi_{w_{i+1}}(w_{i})wiw_{i}wi+1w_{i+1}wk+1=gv0w_{k+1}=gv_{0}w0=v0w_{0}=v_{0}ow0o\in w_{0}gowk+1go\in w_{k+1}β:\beta:β~:\tilde{\beta}:
Figure 13. Lift a standard path β\beta (at bottom) in 𝒫K()\mathcal{P}_{K}(\mathscr{F}) to admissible path (at top) in X\mathrm{X}

We first establish the desired map mentioned in Proposition 11.5.

Lemma 11.7.

There exists a map ι:T[Λ(Γo)]\iota:\partial T\to[\Lambda(\Gamma o)] so that the image ι(T)\iota(\partial T) is contained in the conical limit set [Λcon(Go)][\Lambda^{\mathrm{con}}(Go)]. Moreover, if ι(p)=ι(q)\iota(p)=\iota(q) for pqp\neq q, then ι(p)\iota(p) is outside the set 𝒞\mathcal{C} of non-pinched points.

Proof.

By construction, T0={gH,gG0:gG}T^{0}=\{gH,gG_{0}:g\in G\}, so any geodesic ray α\alpha in TT from v0=G0v_{0}=G_{0} has the list of all vertices AnA_{n}, which are alternating left HH and G0G_{0} cosets. This determines a unique infinite word of form (40). By Step (3), the lifted (L~,τ)(\tilde{L},\tau)–admissible path β~\tilde{\beta} in X\mathrm{X} intersects infinitely many translates XnX_{n} of Nr(Ax(f))N_{r}(\mathrm{Ax}(f)) under GG, as AnA_{n} contains infinitely many G0G_{0}–cosets. By Definition 2.13(B), XnX_{n} accumulates into a [][\cdot]–class [ξ][\xi]. If we choose a sequence gnAng_{n}\in A_{n}, we see d(gno,[o,ξ])rd(g_{n}o,[o,\xi])\leq r so ξ\xi lies in [Λcon(Go)][\Lambda^{\mathrm{con}}(Go)]. Setting ι(α+)=[ξ]\iota(\alpha^{+})=[\xi] defines the desired map ι:T[Λcon(Go)]\iota:\partial T\to[\Lambda^{\mathrm{con}}(Go)].

To justify the claim of injectivity, consider a distinct geodesic αβ\alpha\neq\beta originating at v0v_{0} with the set of vertices BnB_{n}, so that [ξ]=ι(α+)=ι(β+)[\xi]=\iota(\alpha^{+})=\iota(\beta^{+}). We shall show that [ξ][\xi] is pinched. Let γ\gamma be the bi-infinite geodesic in the tree TT from α+\alpha^{+} to β+\beta^{+}, i.e.: γ=(αβ)[v0,w0)\gamma=(\alpha\cup\beta)\setminus[v_{0},w_{0}) as a set, where w0w_{0} is the vertex that β\beta departs from α\alpha. As shown in the first paragraph, there are two sequences gnoAnog_{n}o\in A_{n}o and gnoBnog^{\prime}_{n}o\in B_{n}o tending to [ξ][\xi].

We read off from γ\gamma a bi-infinite alternating word UU in the form (40). As described in Step (3), we obtain from γn\gamma_{n} an (L~,τ)(\tilde{L},\tau)–admissible path γ^n\hat{\gamma}_{n} from gnog_{n}o to gnog_{n}^{\prime}o, where the contracting subsets correspond to the G0G_{0}–vertices on γ\gamma. By Proposition 2.9, [gn,gno][g_{n},g_{n}^{\prime}o] has at most rr–distance to the entry and exit points of the contracting subsets. See Fig. 3. Recall that TT is a bipartite graph with G0G_{0}–vertices and HH–vertices. Look at the first G0G_{0}–vertex ww on β\beta after w0w_{0}, and its exit point xx of the contracting subset. By the above discussion, we have d(x,[gno,gno])rd(x,[g_{n}o,g_{n}^{\prime}o])\leq r, that is [gno,gno][g_{n}o,g^{\prime}_{n}o] is non-escaping. Hence, [ξ]𝒞[\xi]\notin\mathcal{C} is pinched. ∎

We continue to examine the possible intersection in (38).

Lemma 11.8.

The following hold.

  1. (1)

    For any gGHg\in G\setminus H, the intersection [Λ(Ho)]g[Λ(Ho)][\Lambda(Ho)]\cap g[\Lambda(Ho)] consist of only pinched points.

  2. (2)

    [Λ(Ho)][\Lambda(Ho)] intersects ι(T)\iota(\partial T) only in pinched points.

  3. (3)

    [Λ(Ho)]g[f±]=[\Lambda(Ho)]\cap g[f^{\pm}]=\emptyset for any 1gG1\neq g\in G.

Proof.

(1). Without loss of generality, we may represent gg as a nonempty alternating word U=𝔥1fn1𝔥kfnkU=\mathfrak{h}_{1}f^{n_{1}}\cdots\mathfrak{h}_{k}f^{n_{k}} as in (40), ending with letters in G0G_{0}. Let hnoHoh_{n}o\in Ho and gnogHog_{n}o\in gHo tend to the same [ξ][\xi]. We will show [ξ]𝒞[\xi]\notin\mathcal{C} ia pinched: that is, γn:=[hno,gno]\gamma_{n}:=[h_{n}o,g_{n}o] intersects a compact set for all n1n\geq 1.

To this end, consider the nonempty alternating word WW representing hn1gnh_{n}^{-1}g_{n}. As hnHh_{n}\in H and gngHg_{n}\in gH, the word WW has the form h~nfn1𝔥kfnk\tilde{h}_{n}f^{n_{1}}\cdots\mathfrak{h}_{k}f^{n_{k}}, where h~n=hn1𝔥1\tilde{h}_{n}=h_{n}^{-1}\mathfrak{h}_{1} gives one letter in HH. We obtain from WW, as described in the Step (3) above, an (L,τ)(L,\tau)–admissible path β\beta, which has the same endpoints o,hn1gnoo,h_{n}^{-1}g_{n}o as hn1γnh_{n}^{-1}\gamma_{n}. Note that 𝔥1fn1o\mathfrak{h}_{1}f^{n_{1}}o is the exit vertex of the admissible path hnβh_{n}\beta. Up to translation, we obtain that γn\gamma_{n} has rr–distance to the fixed point 𝔥1fn1o\mathfrak{h}_{1}f^{n_{1}}o by Proposition 2.9. This shows that [ξ][\xi] lies outside of the non-pinched points 𝒞\mathcal{C}.

(2). The same reasoning shows that Λ(Ho)\Lambda(Ho) intersects ι(T)\iota(\partial T) only in pinched points. Indeed, if [ξ][\xi] lies in ι(T)\iota(\partial T), then it is an accumulation point of gnog_{n}o, where gng_{n} lies in infinitely many G0G_{0}–cosets (see the proof of Lemma 11.7). The same argument as above implies that γn\gamma_{n} is non-escaping.

(2). The same argument proves that Λ(Ho)g[f±]\Lambda(Ho)\cap g[f^{\pm}] must be pinched. However, since [f±][f^{\pm}] is assumed to be non-pinched points, we obtain Λ(Ho)g[f±]=\Lambda(Ho)\cap g[f^{\pm}]=\emptyset. ∎

It is now fairly easy to derive from Lemma 11.7 and Lemma 11.8:

Proof of Proposition 11.5.

Note that GG is the union of all left G0G_{0}–cosets and HH–cosets. Consider a limit point ξ[Λ(Go)]\xi\in[\Lambda(Go)] in the following complement

[Λ(Go)][gGg[Λ(Ho)]][gGg[Λ(G0o)]][\Lambda(Go)]\setminus\left[\bigcup_{g\in G}g[\Lambda(Ho)]\right]\bigcup\left[\bigcup_{g\in G}g[\Lambda(G_{0}o)]\right]

so that gno[ξ]g_{n}o\to[\xi], for a sequence of gnGg_{n}\in G belonging to infinitely many distinct G0G_{0}–cosets and/or HH–cosets denoted by AnA_{n}. By the above construction of Bass-Serre tree, AnA_{n} corresponds to vertices on TT, so after passing to subsequence, we obtain a limiting geodesic ray α\alpha in TT starting at 𝐯𝟎\mathbf{v_{0}}: that is, the intersection of α\alpha with [𝐯𝟎,gn𝐯𝟎][\mathbf{v_{0}},g_{n}\mathbf{v_{0}}] becomes unbounded as nn\to\infty. By the proof of Lemma 11.7, this implies ξ\xi is contained in ι(T)\iota(\partial T).

The “moreover” statement follows from Lemma 11.8. The proof is now complete. ∎

Proof of Theorem 11.3.

First of all, as G<ΓG<\Gamma is an infinite normal subgroup, μo(ΛHor(Go))=μo(Λhor(Go))=1\mu_{o}(\Lambda^{\mathrm{Hor}}(Go))=\mu_{o}(\Lambda^{\mathrm{hor}}(Go))=1 by Theorem 1.11.

By Proposition 11.5, [Λ(Go)][\Lambda(Go)] is covered by a countable union G[Λ(Ho)]G\cdot[\Lambda(Ho)] with the remaining “exceptional” set EE:

E:=[gGg[Λ(G0o)]]ι(T).E:=\left[\bigcup_{g\in G}g[\Lambda(G_{0}o)]\right]\cup\iota(\partial T).

As ι(T)\iota(\partial T) is contained in the conical points, so is shadowed by the elements in GoGo. By the assumption that ωG<ωΓ\omega_{G}<\omega_{\Gamma}, we have ι(T)\iota(\partial T) is μo\mu_{o}–null by Lemma 2.34. This then proves μo(E)=0\mu_{o}(E)=0.

This concludes the proof modulo of the following claim.

Claim.

A small/big horospheric limit points of GG is either in EE or a small/big horospheric limit point in a GG–translate of Λ(Ho)\Lambda(Ho). That is,

[ΛHor(Go)]E[gGg[ΛHor(Ho)]][\Lambda^{\mathrm{Hor}}(Go)]\subseteq E\bigcup\left[\bigcup_{g\in G}g[\Lambda^{\mathrm{Hor}}(Ho)]\right]
Proof of the claim.

Indeed, let ξΛHor(Go)E\xi\in\Lambda^{\mathrm{Hor}}(Go)\setminus E. Then there exists a sequence gnoGog_{n}o\in Go tending to [ξ][\xi] in some (or any) horoball ([ξ])\mathcal{HB}([\xi]) based at [ξ][\xi]. As ξE\xi\notin E, gnGg_{n}\in G lies in finitely many HH–cosets, and after passing to subsequence, we may assume gnH=g0Hg_{n}H=g_{0}H for n1n\gg 1. By Lemma 11.8, g0Λ(Ho)Λ(Ho)=g_{0}\Lambda(Ho)\cap\Lambda(Ho)=\emptyset for g0Hg_{0}\notin H, so we conclude that g0H=Hg_{0}H=H and thus ξΛHor(Ho)\xi\in\Lambda^{\mathrm{Hor}}(Ho). ∎

As EE is μo\mu_{o}–null, we thus obtain from Proposition 11.5 that μo(ΛHor(Ho))μo(Λhor(Ho))>0\mu_{o}(\Lambda^{\mathrm{Hor}}{(Ho)})\geq\mu_{o}(\Lambda^{\mathrm{hor}}{(Ho)})>0. ∎

We conclude this section with a corollary which might be of independent interest. Define a topology on M:=T0TM:=T^{0}\cup\partial T as follows (cf. [23]). A sequence of points xnxx_{n}\to x if and only if the geodesic [xn,x][x_{n},x] eventually passes no element in any given finite set of edges (which is required adjacent to xx if xT0x\in T^{0}). This makes MM homeomorphic to a Cantor space, in particular, it is compact. If TT is the Bass-Serre tree of Γ=HK\Gamma=H\star K, then MM is exactly the Bowditch boundary of a hyperbolic group Γ\Gamma relative to {H,K}\{H,K\}. Moreover, if both H,KH,K are one-ended, it is also the end boundary of Γ\Gamma.

Corollary 11.9.

In the setting of Corollary 11.6, there exists a continuous surjective GG–map from Λ(Go)\Lambda(Go) to MM, where MM can be identified with the Bowditch boundary of GG relative to factors G0G_{0} and HH.

The reader may have noticed that item (4) in Examples 2.15 is not covered in the corollaries. It is not clear whether the decomposition holds on the Thurston boundary without taking the intersection with 𝒰\mathscr{UE}.

References

  • [1] M. Abért, N. Bergeron, I. Biringer, T. Gelander, N. Nikolov, J. Raimbault, and I. Samet, On the growth of L2L^{2}-invariants for sequences of lattices in Lie groups, Ann. of Math. (2) 185 (2017), no. 3, 711–790. MR 3664810
  • [2] M. Abért, Y. Glasner, and B. Virág, Kesten’s theorem for invariant random subgroups, Duke Math. J. 163 (2014), no. 3, 465–488. MR 3165420
  • [3] by same author, The measurable Kesten theorem, Ann. Probab. 44 (2016), no. 3, 1601–1646. MR 3502591
  • [4] G. Arzhantseva and C. Cashen, Cogrowth for group actions with strongly contracting elements, Ergodic Theory Dynam. Systems 40 (2020), no. 7, 1738–1754. MR 4108903
  • [5] G. Arzhantseva, C. Cashen, and J. Tao, Growth tight actions, Pacific Journal of Mathematics 278 (2015), 1–49.
  • [6] G. Arzhantseva and I. Lysenok, Growth tightness for word hyperbolic groups, Math. Z. 241 (2002), no. 3, 597–611.
  • [7] J. Athreya, A. Bufetov, A. Eskin, and M. Mirzakhani, Lattice point asymptotics and volume growth on Teichmüller space, Duke Math. J. 161 (2012), no. 6, 1055–1111.
  • [8] Y. Bahturin and A. Olshanskii, Actions of maximal growth, Proceedings of the London Mathematical Society 101 (2010), no. 1, 27–72.
  • [9] M. Bestvina, K. Bromberg, and K. Fujiwara, Constructing group actions on quasi-trees and applications to mapping class groups, Publications mathématiques de l’IHÉS 122 (2015), no. 1, 1–64, arXiv:1006.1939.
  • [10] M. Bestvina, K. Bromberg, K. Fujiwara, and A. Sisto, Acylindrical actions on projection complexes, Enseign. Math. 65 (2019), no. 1-2, 1–32. MR 4057354
  • [11] M. Bestvina, R. Dickmann, G. Domat, S. Kwak, P. Patel, and E. Stark, Free products from spinning and rotating families, Enseign. Math. 69 (2023), no. 3-4, 235–260. MR 4599247
  • [12] M. Bestvina and K. Fujiwara, A characterization of higher rank symmetric spaces via bounded cohomology, Geometric and Functional Analysis 19 (2009), no. 1, 11–40, arXiv:math/0702274.
  • [13] A. Le Boudec and N. Matte Bon, Subgroup dynamics and CC^{*}-simplicity of groups of homeomorphisms, Ann. Sci. Éc. Norm. Supér. (4) 51 (2018), no. 3, 557–602. MR 3831032
  • [14] by same author, Confined subgroups and high transitivity, Ann. H. Lebesgue 5 (2022), 491–522. MR 4443296
  • [15] E. Breuillard, M. Kalantar, M. Kennedy, and N. Ozawa, CC^{*}-simplicity and the unique trace property for discrete groups, Publ. Math. Inst. Hautes Études Sci. 126 (2017), 35–71. MR 3735864
  • [16] R. Brooks, The bottom of the spectrum of a Riemannian covering, J. Reine Angew. Math. 357 (1985), 101–114. MR 783536
  • [17] M. Clay, J. Mangahas, and D. Margalit, Right-angled Artin groups as normal subgroups of mapping class groups, Compos. Math. 157 (2021), no. 8, 1807–1852. MR 4292178
  • [18] M. Coornaert, Mesures de Patterson-Sullivan sure le bord d’un espace hyperbolique au sens de Gromov, Pac. J. Math. (1993), no. 2, 241–270.
  • [19] R. Coulon, Patterson-Sullivan theory for groups with a strongly contracting element, preprint, arXiv:2206.07361.
  • [20] R. Coulon, F. Dal’Bo, and A. Sambusetti, Growth gap in hyperbolic groups and amenability, Geom. Funct. Anal. 28 (2018), no. 5, 1260–1320. MR 3856793
  • [21] R. Coulon, R. Dougall, B. Schapira, and S. Tapie, Twisted Patterson-Sullivan measures and applications to amenability and coverings, appear in Memoirs of the AMS, arXiv: 1809.10881. 2020.
  • [22] Rémi Coulon, Ergodicity of the geodesic flow for groups with a contracting element, 2024.
  • [23] P. SOARDI D. CARTWRIGHT and W. WOESS, Martin and end compactifications for nonlocally finite graphs, Transactions of the American Mathematical Society 338 (1993), no. 2, 679–693.
  • [24] 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
  • [25] F. Dal’bo, P. Otal, and M. Peigné, Séries de Poincaré des groupes géométriquement finis, Israel Journal of math. 118 (2000), no. 3, 109–124.
  • [26] F. Dal’bo, M. Peigné, J.C. Picaud, and A. Sambusetti, On the growth of quotients of Kleinian groups, Ergodic Theory and Dynamical Systems 31 (2011), no. 3, 835–851.
  • [27] K. Falk and K. Matsuzaki, On horospheric limit sets of Kleinian groups, J. Fractal Geom. 7 (2020), no. 4, 329–350. MR 4184586
  • [28] M. Fraczyk, Kesten’s theorem for uniformly recurrent subgroups, Ergodic Theory Dynam. Systems 40 (2020), no. 10, 2778–2787. MR 4138911
  • [29] M. Fraczyk and T. Gelander, Infinite volume and infinite injectivity radius, Ann. of Math. (2) 197 (2023), no. 1, 389–421. MR 4515682
  • [30] K. Fujiwara, K. Nagano, and T. Shioya, Fixed point sets of parabolic isometries of cat(0)-spaces, Commentarii Mathematici Helvetici 81 (2006), 305–335.
  • [31] I. Gekhtman and A. Levit, Stationary random subgroups in negative curvature, 2023.
  • [32] I. Gekhtman, Y.L. Qing, and K. Rafi, Genericity of sublinearly morse directions in CAT(0) spaces and the Teichmüler space, preprint, arXiv:2208.04778.
  • [33] E. Ghys and P. de la Harpe, Sur les groupes hyperboliques d’après Mikhael Gromov, Progress in Math., Birkau¨\ddot{u}ser, 1990.
  • [34] E. Glasner and B. Weiss, Uniformly recurrent subgroups, Recent trends in ergodic theory and dynamical systems, Contemp. Math., vol. 631, Amer. Math. Soc., Providence, RI, 2015, pp. 63–75. MR 3330338
  • [35] R. Grigorchuk, V. Kaimanovich, and T. Nagnibeda, Ergodic properties of boundary actions and the Nielsen-Schreier theory, Adv. Math. 230 (2012), no. 3, 1340–1380. MR 2921182
  • [36] S.Z. Han, W.Y. Yang, and Y.Q. Zou, Counting double cosets with application to generic 3-manifolds, 2023.
  • [37] B. Hartley and A. E. Zalesskiĭ, Confined subgroups of simple locally finite groups and ideals of their group rings, J. London Math. Soc. (2) 55 (1997), no. 2, 210–230. MR 1438625
  • [38] P. Tukia J. Anderson, K. Falk, Conformal measures associated to ends of hyperbolic n-manifolds, The Quarterly Journal of Mathematics (2007), 1–15.
  • [39] J. Jaerisch and K. Matsuzaki, Weighted cogrowth formula for free groups, Groups Geom. Dyn. 14 (2020), no. 2, 349–368.
  • [40] V. Kaimanovich, Hopf decomposition and horospheric limit sets, Ann. Acad. Sci. Fenn. Math. 35 (2010), no. 2, 335–350. MR 2731695
  • [41] V. Kaimanovich and H. Masur, The poisson boundary of the mapping class group, Invent. math. 125 (1996), 221–264.
  • [42] V. A. Kaimanovich, Ergodic properties of the horocycle flow and classification of Fuchsian groups, J. Dynam. Control Systems 6 (2000), no. 1, 21–56. MR 1738739
  • [43] M. Kalantar and M. Kennedy, Boundaries of reduced CC^{*}-algebras of discrete groups, J. Reine Angew. Math. 727 (2017), 247–267. MR 3652252
  • [44] H. Kesten, Full Banach mean values on countable groups, Math. Scand. 7 (1959), 146–156. MR 112053
  • [45] G. Kneiper, The uniqueness of the measure of maximal entropy for geodesic flows on rank 1 manifolds, Annals of Mathematics 148 (1998), no. 1, 291–314.
  • [46] G. Link, Hopf-Tsuji-Sullivan dichotomy for quotients of Hadamard spaces with a rank one isometry, Discrete Contin. Dyn. Syst. 38 (2018), no. 11, 5577–5613. MR 3917781
  • [47] L.X. Liu and W.X. Su, The horofunction compactification of the Teichmüller metric, Handbook of Teichmüller theory. Vol. IV, IRMA Lect. Math. Theor. Phys., vol. 19, Eur. Math. Soc., Zürich, 2014, pp. 355–374. MR 3289706
  • [48] K. Matsuzaki, Conservative action of Kleinian groups with respect to the Patterson-Sullivan measure, Comput. Methods Funct. Theory 2 (2002), no. 2, 469–479. MR 2038133
  • [49] by same author, Isoperimetric constants for conservative Fuchsian groups, Kodai Math. J. 28 (2005), no. 2, 292–300. MR 2153917
  • [50] K. Matsuzaki, Y. Yabuki, and J. Jaerisch, Normalizer, divergence type, and Patterson measure for discrete groups of the Gromov hyperbolic space, Groups Geom. Dyn. 14 (2020), no. 2, 369–411. MR 4118622
  • [51] Katsuhiko Matsuzaki and Masahiko Taniguchi, Hyperbolic manifolds and Kleinian groups, Oxford Mathematical Monographs, The Clarendon Press, Oxford University Press, New York, 1998, Oxford Science Publications. MR 1638795
  • [52] Curtis T. McMullen, Renormalization and 3-manifolds which fiber over the circle, Annals of Mathematics Studies, vol. 142, Princeton University Press, Princeton, NJ, 1996. MR 1401347
  • [53] A. Yu. Olshanskii, Subnormal subgroups in free groups, their growth and cogrowth, Math. Proc. Cambridge Philos. Soc. 163 (2017), no. 3, 499–531. MR 3708520
  • [54] S. Patterson, The limit set of a Fuchsian group, Acta Mathematica (1976), no. 1, 241–273.
  • [55] by same author, Spectral theory and Fuchsian groups, Math. Proc. Cambridge Philos. Soc. 81 (1977), no. 1, 59–75. MR 447120
  • [56] V. Pit and B. Schapira, Finiteness of Gibbs measures on noncompact manifolds with pinched negative curvature, Ann. Inst. Fourier (Grenoble) 68 (2018), no. 2, 457–510. MR 3803108
  • [57] Y.L. Qing and W. Y. Yang, Genericity of sublinear Morse directions in general metric spaces, arXiv:2404.18762, 2023.
  • [58] Jean-François Quint, An overview of patterson-sullivan theory, available at https://www.math.u-bordeaux.fr/ jquint/publications/courszurich.pdf.
  • [59] John G. Ratcliffe, Foundations of hyperbolic manifolds, second ed., Graduate Texts in Mathematics, vol. 149, Springer, New York, 2006. MR 2249478
  • [60] T. Roblin, Sur la fonction orbitale des groupes discrets en courbure négative, Ann. Inst. Fourier 52 (2002), 145–151.
  • [61] by same author, Ergodicité et équidistribution en courbure négative, no. 95, Mémoires de la SMF, 2003.
  • [62] A. Sambusetti, Growth tightness of free and amalgamated products, Ann. Sci. École Norm. Sup. série 35 (2002), no. 4, 477 – 488.
  • [63] B. Schapira and S. Tapie, Regularity of entropy, geodesic currents and entropy at infinity, Ann. Sci. Éc. Norm. Supér. (4) 54 (2021), no. 1, 1–68. MR 4245867
  • [64] D. Sullivan, The density at infinity of a discrete group of hyperbolic motions, Publ. Math. IHES (1979), 171–202.
  • [65] by same author, On the ergodic theory at infinity of an arbitrary discrete group of hyperbolic motions, Riemann surfaces and related topics: Proceedings of the 1978 Stony Brook Conference (State Univ. New York, Stony Brook, N.Y., 1978), Ann. of Math. Stud., vol. No. 97, Princeton Univ. Press, Princeton, NJ, 1981, pp. 465–496. MR 624833
  • [66] by same author, Entropy, hausdorff measures old and new, and limit sets of geometrically finite Kleinian groups, Acta Mathematica (1984), 259–277.
  • [67] P. Tukia, Conservative action and the horospheric limit set, Ann. Acad. Sci. Fenn. Math. 22 (1997), no. 2, 387–394. MR 1469798
  • [68] C. Walsh, The asymptotic geometry of the Teichmüller metric, Geom. Dedicata 200 (2019), 115–152. MR 3956189
  • [69] Wenyuan Yang, Growth tightness for groups with contracting elements, Math. Proc. Cambridge Philos. Soc 157 (2014), 297 – 319.
  • [70] by same author, Purely exponential growth of cusp-uniform actions, Ergodic Theory Dynam. Systems 39 (2019), no. 3, 795–831. MR 3904188
  • [71] by same author, Statistically convex-cocompact actions of groups with contracting elements, Int. Math. Res. Not. IMRN (2019), no. 23, 7259–7323. MR 4039013
  • [72] by same author, Conformal dynamics at infinity for groups with contracting elements, arXiv: 2208.04861, 2023.