Confined subgroups in groups with contracting elements
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 rate2000 Mathematics Subject Classification:
Primary 20F65, 20F67, 37D401. Introduction
Let be a locally compact, second countable topological group. The space of all closed subgroups in , equipped with the Chabauty topology, is a compact metrizable space on which acts continuously by conjugation. Measurable and topological dynamics of the action 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 , have attracted a lot of interest since the terminology was coined in [2]. Their topological counterpart, namely –minimal systems in , were introduced as uniformly recurrent subgroups (URS) in [34]. URS play an important role in the connections between reduced –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 be two subgroups of , where is not necessarily contained in . We say that is confined by in if its –orbit (i.e. all conjugates of under ) does not accumulate in to the trivial subgroup. Equivalently, there exists a compact confining subset such that for all . When , we simply say that is a confined subgroup of . Every subgroup in a non-trivial URS of is confined by . Conversely, if is confined by , then the –orbit closure of in contains a non-trivial –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 , which acts on a geodesic metric space 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 is called contracting if the orbital map
is a quasi-isometric embedding, so that the image has contracting property: any ball disjoint with has -bounded projection to for a constant independent of .
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 acting on a suitable boundary of the space ; and the growth rates of and its quotient . Our investigation focuses mainly on the questions related to these aspects for a discrete confined subgroup .
Fix a basepoint . We define the growth rate (or critical exponent) of the orbit as
(1) |
where . When is discrete, we also consider the growth rate , defined below in (3) analogously to (1), of the image under the projection equipped with the quotient metric. 222In the literature, if is a subgroup of , is usually called relative growth; the relative growth of a normal subgroup is referred as cogrowth relative to the growth of (particularly on , see [33]). Occasionally, the terminology is reversed: is called cogrowth relative to the growth of (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 of Laplace-Beltrami spectrum on as follows:
where is the Hausdorff dimension of visual boundary. If is a lattice, then .
In the discrete setting, by the cogrowth formula due to Grigorchuk, for a subgroup in the free group on generators, the spectral radius of simple random walks on is given by
where is the –regular tree (the standard Cayley graph of ), and is the Hausdorff dimension of the space of ends of .
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 of some natural classes of groups, a normal subgroup attains the maximal relative growth rate if and only if the quotient 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 holds for normal subgroups 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 remains largely unexplored in the literature. Furthermore, the relations between and still remain mysterious, despite recent works [39, 19].
In this paper, for a confined subgroup of , we establish the conservativity of the boundary action of , a strict lower bound (such an inequality is sometimes referred to as cogrowth tightness), and an inequality relating to and . 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 space , equipped with the Gromov or visual boundary in the first and second cases respectively. Let be a non-elementary discrete subgroup containing a loxodromic or rank-1 element accordingly. If the action has purely exponential growth (or more generally, of divergence type), there exists a unique –dimensional Patterson-Sullivan measure class on the Gromov or visual boundary constructed from the action . Denote by the set of isometries in which fix pointwise the limit set (Definition 5.5).
Theorem 1.3.
Assume the action has purely exponential growth. Let be a nontrivial torsion-free discrete subgroup confined by with a compact confining subset. Then . Furthermore,
-
(1)
If preserves the measure class of , then the action is conservative.
-
(2)
If admits a finite confining subset intersecting trivially , then and .
In many circumstances, the group of isometries fixing boundary pointwise is trivial (or finite), e.g. if 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 , is finite and the assumption in the item (2) is always fulfilled.
An important subclass of (uniquely) geodesically complete CAT spaces are provided by Hadamard manifolds (i.e. a complete, simply connected –dimensional Riemannian manifold with non-positive sectional curvature). Its visual boundary , which is defined as the set of equivalence classes of geodesic rays, is homeomorphic to . Let be a torsion-free discrete group. The quotient manifold is called rank-1 if it admits a closed geodesic without a perpendicular parallel Jacobi field. In other words, is rank-1 if and only if contains a contracting element [12, Theorem 5.4].
Theorem 1.4.
Let be a rank-1 manifold, and assume that the geodesic flow on the unit tangent bundle has finite measure of maximal entropy. Let be the unique –dimensional Patterson-Sullivan measure class on constructed from the action . Let be a nontrivial torsion-free discrete subgroup confined by with a compact confining subset. Then . Furthermore,
-
(1)
If preserves the measure class of , then the action is conservative.
-
(2)
If admits a finite confining subset intersecting trivially , then and .
As noted above, if has full limit set , then 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 is not necessarily a lattice in . The maximal entropy measure lies in the measure class modulo –action on (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 is a uniform lattice and is a rank-1 symmetric space of noncompact type, then is the Lebesgue measure on invariant under (see [58]). In this restricted setting, if is confined by (equivalently by any uniform lattice, see Lemma 5.3), the strict inequality was proved by Gekhtman-Levit [31] (without the finite confining subset assumption).
- •
-
•
If is contained in , the assumptions after “Furthermore” are redundant, and the corresponding conclusions hold without them. A confined subgroup in a non-uniform lattice is not necessarily confined in , so the inequality in item (2) does not follow from [31].
Another application is given for confined subgroups in the mapping class group of a closed orientable surface (). The finitely generated group is actually the orientation preserving isometry group of the Teichmüller space with Teichmüller metric, on which it acts properly with growth rate . Investigating the similarities between and lattices in semi-simple Lie groups has been an active research theme. The following result fits into the rank-1 phenomenon of .
Theorem 1.6.
Consider the measure class preserving action of on the Thurston boundary of the Teichmüller space with Thurston measure (cf. [7] recalled in Example 2.26).
Let be a nontrivial confined subgroup. If , assume is infinite. Then the following holds
-
(1)
is conservative.
-
(2)
.
-
(3)
.
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 acts cocompactly on the –thick part of for any fixed . It follows that in the setting of Theorem 1.6, is a confined subgroup of if and only if has bounded injectivity radius from above over any fixed thick part of .
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 is a proper action on CAT(-1) space. Let be a -dimensional -conformal density on the Gromov boundary for some . Then we have the dichotomy: either the following equivalent statements hold:
-
(I.1)
is divergent at .
-
(I.2)
is supported on the set of conical points .
-
(I.3)
is ergodic. In particular, is ergodic.
or the following equivalent statements hold:
-
(II.1)
is convergent at .
-
(II.2)
is null on the set of conical points .
-
(II.3)
is completely dissipative.
Under the set (I) of conditions, we must have and 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 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 (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- symmetric space of non-compact type, with a family of –equivariant conformal measures in the Lebesgue measure class on the visual boundary of . This class of conformal measures can be realized as Patterson-Sullivan measures for any uniform lattice . See [58].
- •
-
•
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.
- •
-
•
In recent works [19, 72], the horofunction boundary 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 admits a convergence boundary (or keep one of these examples in mind). Motivated by rank-1 symmetric space, we assume that there is an auxiliary proper action , in order to endow a –dimensional –equivariant conformal density on . Sometimes, we assume that 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) |
Most of results actually only assume a PEG action , 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 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 on .
In contrast to the HTS dichotomy of geodesic flow invariant measures, which can be applicable to the –action on , 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 in a free group with respect to the uniform measure on . These works serve as a source of inspiration in our considerations in a more general framework.
The limit set denotes the set of accumulation points of in a convergence boundary . In the last two cases in Example 1.9, may depend on . To avoid this difficulty, a nontrivial partition on was introduced so that the set of –classes on 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.
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 act properly with contracting elements, and be the PS measure constructed from a proper SCC action . Assume that . Then , where denotes the big horospheric limit set of defined in (2.46).
In addition, assume that is torsion-free and preserves the measure class of on (in particular, if is a subgroup of ). Then the action of on is completely dissipative.
Note that, if is hyperbolic and is co-compact, then any preserves the measure class of (see Lemma 4.1). We expect Theorem 1.10 to hold under the weaker assumption that 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 the set of isometries in which fix every –class in the limit set (Definition 5.5).
Theorem 1.11.
Suppose that a discrete subgroup is confined by with a compact confining subset . Assume that is of divergence type, and furthermore:
-
(i)
Either is torsion-free and is a Gromov hyperbolic or geodesically complete CAT(0) space, or
-
(ii)
the confining subset is finite, and intersects trivially .
Then . In particular, is conservative.
Note that the action of on 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 is SCC. In the setting of Theorem 1.11, we have .
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 strictly larger than the SCC action considered in §1.2 (particularly in Theorem 4.2).
Theorem 1.13.
Assume that the proper action has purely exponential growth with a contracting element. If is a discrete subgroup confined by with a finite confining subset that intersects trivially with , then .
If is normal, Coulon [19] and independently the third named author [72] proved this inequality using boundary measures in greater generality (only assuming is of divergence type). In addition, if is a normal subgroup of divergence type, then . 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 .
Corollary 1.14.
Assume that is a PEG action with contracting elements. Let be a sequence of subgroups so that is confined in for each . Then .
This in particular applies to an –subnormal subgroup for some integer , for which there exists a strictly increasing subnormal series, , of length . This statement was known in free groups by Olshanskii [53], who also proved that the lower bound is sharp: for any and , there exists some –subnormal subgroup for which .
Next, consider the space equipped with the quotient metric. That is, given , define . Denote by the natural projection.
Fixing a point , consider the ball-like set in the image of in :
and define its growth rate as follows:
(3) |
As and have a finite Hausdorff distance, the growth rate does not depend on . If is the Cayley graph of and is a subgroup of , then is the Schreier graph associated with . If in addition, is a normal subgroup of , then is the usual growth rate of 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 is of divergence type with a contracting element. If a discrete group is confined by with a finite confining subset that intersects trivially with , then
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 is not required to be contained in .
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 .
Note that it follows from Theorem 1.15 that if has sub-exponential growth (i.e. ), then . By [19][Cor. 4.27], for a normal subgroup , we have if is amenable. We do not know what condition on might characterize the equality 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 (i.e. fixing pointwise) intersects in a finite subgroup. If is a torsion-free subgroup, then the confining subset of intersects trivially , so the assumption for (2) holds automatically.
For mapping class groups, if , then is trivial, so Theorem 1.6 follows exactly as above. If , contains hyperelliptic involution, then an additional argument is needed to conclude the proof. Indeed, since is residually finite, we may pass to finite index torsion-free subgroups of , where 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 orbit in the quotient . A proper action is called growth tight if holds for any infinite normal subgroup . In the setting of Theorem 1.11, the action of a torsion-free on the boundary is conservative. The inequality in particular implies that the quotient growth of is slower than in the following sense:
Following Bahturin-Olshanski [8], we say that has maximal quotient growth 444In terms of [8], is the growth function of the right (usually non-isometric) action of on . The right action is not relevant here, so we take the term of quotient growth instead. if the following holds:
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 acts properly and co-compactly on a proper hyperbolic space and is the unique Patterson-Sullivan measure class on the Gromov boundary . Let be a subgroup. Then the small/big horospheric limit set is –full, if and only if grows slower than . Moreover, is equivalent to the following
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 be a conformal measure on a boundary without atoms (cf. Example 2.49). Under which conditions, do we have ?
This question has been raised in Sullivan’s work [65], where he showed for the real hyperbolic space that a discrete subgroup is conservative with respect to if and only if the volume of grows slower than . The key part of the proof is that the big and small horospheric limit sets differ in a –null set. We elaborate this proof to answer the question positively for any subgroup 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 has purely exponential growth with a contracting element. Let be an infinite normal subgroup of . Then .
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 is torsion-free. Then the action is conservative if and only if the quotient growth of is slower than .
Another corollary is characterizing the purely exponential growth of right cosets as in the “moreover” statement.
Let be two subgroups with proper limit sets (we say these are of second kind following the terminology in Kleinian groups). In [36], it is shown that the double cosets over have purely exponential growth:
If , this amounts to counting the right coset as above. In hyperbolic groups, this case is thus covered in Theorem 1.17, which furthermore gives a characterization of purely exponential growth of by . If , a characterization remains unknown to us for purely exponential growth of left cosets .
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 admits a proper SCC action on a proper hyperbolic space . Let be a loxodromic element. Then there exist a large integer so that the normal closure contains subgroups of second kind with nontrivial conservative part and dissipative part.
We expect this result extends to any SCC action. Indeed, denoting , if is proved, could be replaced with a general metric space. Amenability criterion in [21] for hyperbolic spaces says that is equivalent to the amenability of . If is sufficiently large, is non-amenable, so . If the amenability criterion holds for any SCC action, then follows and so does Theorem 1.21.
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 for simplicity. Our standing assumption is that the space admits a convergence boundary , and acts properly on . Readers who are not familiar with the background on convergence boundary might keep in mind the example where is hyperbolic space with Gromov boundary : 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 , as presented above in §1.2, makes extensive use of the conformal measure theory developed in [72] on the convergence boundary, provided that is of divergence type. The case of horofunction boundary was independently obtained by Coulon [19]. The key facts we use include the following:
-
(1)
the Shadow Lemma 2.31 holds for , and
- (2)
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 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 is a confined subgroup, we push a generic set of limit points into the horospheric limit set of to prove the conservativity of the action of , 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 of contracting elements in with the following property: for any , there exist and such that lies in and
where depends only on and .
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 be a –dimensional –quasi-equivariant quasi-conformal density supported on . Then there exists such that
for any and .
The usual Shadow Lemma asserts only the above inequalities on a smaller region of , where . Hence, the Shadow Principle provides useful information in the case that is a much larger set. If is cocompact, the above result holds over the whole space .
Cogrowth tightness of confined subgroups
The nonstrict part of inequality 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 on the mass of over :
(4) |
Once this is proved, the coefficient in the Shadow Principle disappears, so we would obtain from a standard covering argument. That is, the number of shadows , which contain a given point of is bounded by a uniform constant , and therefore
which yields and thus follows for normal subgroups of divergence type.
To make life easier, let us assume is a CAT(-1) space, so the set (I) of conditions in Theorem 1.8 holds. As is of divergence type, there exists a unique PS measure class of dimension , up to scaling. In particular, the PS measure is the unique limit point of the following one
for any , where is the Poincaré series associated to . Note that
(5) |
The proof would be finished at this point, if is a normal subgroup: the RHS is the inverse of the mass . The uniqueness of the limit concludes that , 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 , we make a crucial use of Lemma 1.22 to obtain the following estimates on growth function of each conjugate
in Lemma 8.5, which yields the coarse equality of the associated Poincaré series:
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 as above, resulting in the equality . This contradicts our assumption, concluding the proof of the strict inequality.
The argument by contradiction leaves open whether equality 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 of a contracting element for some sufficiently large . It is well known that such a is a free group of infinite rank ([24]). Removing one generator gives a subgroup 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 by the Amenability Theorem [21] for SCC action on hyperbolic spaces. We mention that the equivalence of with non-amenability of 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 be a proper geodesic metric space. Let be the isometry group endowed with compact open topology. It is well known that a subgroup is discrete if and only if acts properly on ([59, Theorem 5.3.5]).
Let be a path parametrized by arc-length, from the initial point to the terminal point . If , the restriction of to for is referred to as a positive ray, and its complement a negative ray. By abuse of language, we often denote them by and (in particular, when they represent boundary points to which the half rays converge as in Definition 2.13).
Given two parametrized points , denotes the parametrized subpath of going from to , while is a choice of a geodesic between .
A path is called a –quasi-geodesic for if for any rectifiable subpath ,
where denotes the length of .
Denote by (or simply ) the concatenation of two paths provided that .
Let be real-valued functions. Then means that there is a constant depending on parameters such that . The symbol is defined similarly, and means both and are true. The constant will be omitted if it is a universal constant.
2.1. Contracting geodesics
Let be a closed subset of and be a point in . By we mean the set-distance between and , i.e.
Let
be the set of closet point projections from to . Since is a proper metric space, is non empty. We refer to as the projection set of to . Define , where denotes the diameter.
Definition 2.1.
We say a closed subset is –contracting for a constant if, for all pairs of points , we have
Any such is called a contracting constant for . A collection of –contracting subsets shall be referred to as a –contracting system.
An element is called contracting if it acts co-compactly on a contracting bi-infinite quasi-geodesic. Equivalently, the map is a quasi-geodesic with a contracting image.
Unless explicitly stated, let us assume from now on that is a discrete group, so is a proper action (i.e. with discrete orbits and finite point stabilizers).
A group is called elementary if it is virtually 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 , we have
We shall suppress and write if is clear in context.
Keeping in mind the basepoint , the axis of is defined as the following quasi-geodesic
(6) |
Notice that and for any contracting element .
An element preserves the orientation of a bi-infinite quasi-geodesic if and has finite Hausdorff distance for any half ray of . Let be the subgroup of with possibly index 2 which elements preserve the orientation of their axis. Then we have
and contains all contracting elements in , and consists of torsion elements.
Definition 2.3.
Two contracting elements in a discrete group are called independent if the collection is a contracting system with bounded intersection: for any , there exists so that
This is equivalent to the bounded projection: for some independent of .
In a possibly nondiscrete group , we say that are weakly independent if and 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 and a set in . A geodesic contains an –barrier for if there exists an element so that
(7) |
By abuse of language, the point or the axis is called –barrier on .
2.2. Extension Lemma
We fix a finite set of independent contracting elements and let . The following notion of an admissible path allows to construct a quasi-geodesic by concatenating geodesics via .
Definition 2.6 (Admissible Path).
Given , a path is called -admissible in , if is a concatenation of geodesics , where the two endpoints of each lie in some , and the following Long Local and Bounded Projection properties hold:
-
(LL)
Each for has length bigger than , and could be trivial;
-
(BP)
For each , we have and , where and by convention.
The collection is referred to as contracting subsets associated with the admissible path.
Remark 2.7.
-
(1)
The path could be allowed to be trivial, so by the (BP) condition, it suffices to check . It will be useful to note that admissible paths could be concatenated as follows: Let and be –admissible. If has length bigger than , then the concatenation has a natural –admissible structure.
-
(2)
In many situations, could be chosen as the bounded projection constant of . In fact, if is large relative to , an –admissible path could be always truncated near contracting subsets so that it becomes an –admissible path. See [72, Lemma 2.14].
We frequently construct a path labeled by a word , which by convention means the following concatenation
where the basepoint is understood in context. With this convention, the paths labeled by and respectively differ, depending on whether is a geodesic or not.
A sequence of points in a path is called linearly ordered if for each .
Definition 2.8 (Fellow travel).
Let be an admissible path. We say has –fellow travel property for some if for any geodesic with the same endpoints as , there exists a sequence of linearly ordered points () on such that
In particular, for each .
The following result says that a local long admissible path enjoys the fellow travel property.
Proposition 2.9.
[69] For any , there exist depending only on such that any –admissible path has –fellow travel property. In particular, it is a –quasi-geodesic.
The next lemma gives a way to build admissible paths.
Lemma 2.10 (Extension Lemma).
For any independent contracting elements , there exist constants depending only on with the following property.
Choose any element for each to form the set satisfying . Let be any two elements.
-
(1)
There exists an element such that the path
is an –admissible path relative to .
-
(2)
The point is an –barrier for any geodesic .
Remark 2.11.
Since admissible paths are local conditions, we can connect via any number of elements 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 and label two –admissible paths with the same endpoints, where satisfy Proposition 2.9. For any , there exists with the following property. If and , then .
2.3. Horofunction boundary
We recall the notion of horofunction boundary.
Fix a basepoint . For each , we define a Lipschitz map by
This family of –Lipschitz functions sits in the set of continuous functions on vanishing at . Endowed with the compact-open topology, the Arzela-Ascoli Lemma implies that the closure of gives a compactification of . The complement of in this compactification is called the horofunction boundary of and is denoted by .
A Buseman cocycle (independent of ) is given by
The topological type of horofunction boundary is independent of the choice of a basepoint. Every isometry of induces a homeomorphism on :
Depending on the context, we may use both and to denote a point in the horofunction boundary.
Finite difference relation.
Two horofunctions have –finite difference for if the –norm of their difference is –bounded:
The locus of consists of horofunctions so that have –finite difference for some . The loci of horofunctions form a finite difference equivalence relation on . The locus of a subset is the union of loci of all points in .
If and are sequences with , then .
2.4. Convergence boundary
Let be a proper metric space admitting an isometric action of a non-elementary countable group with a contracting element. Consider a metrizable compactification , so that is open and dense in . We also assume that the action of extends by homeomorphism to .
We equip with a –invariant partition : implies for any . We say that is minimal if , and a subset is –saturated if .
We say that restricts to be a closed partition on a –saturated subset if and are two sequences with , then . (The points are not necessarily in .) If , this is equivalent to saying that the relation is a closed subset in , so the quotient space is Hausdorff. In general, may not be closed over the whole (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 tends (resp. accumulates) to if the limit point (resp. any accumulate point) is contained in the subset . This implies that tends or accumulates to in the quotient space . So, an infinite ray terminates at if any sequence of points in accumulates in .
Recall that is the cone of a subset with light source at . A sequence of subsets is escaping if for some (or any) .
Definition 2.13.
We say that is a convergence compactification of if the following assumptions hold.
-
(A)
Any contracting geodesic ray accumulates into a closed subset for some ; and any sequence with escaping projections tends to .
-
(B)
Let be an escaping sequence of –contracting quasi-geodesics for some . Then for any given , there exists a subsequence of (still denoted by ) and such that accumulates to : any convergent sequence of points tends to a point in .
-
(C)
The set of non-pinched points is non-empty. We say is non-pinched if are two sequences of points converging to , then is an escaping sequence of geodesic segments. Moreover, for any , the partition restricted to forms a closed relation.
Remark 2.14.
Assumption (C) is a non-triviality condition: any compactification with the coarsest partition asserting as one –class satisfy (A) and (B). We require 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 is maximal in the sense that every –class is singleton, then the “moreover” assumption is always true.
By Assumption (A), a contracting quasi-geodesic ray determines a unique –class of a boundary point denoted by . Similarly, the positive and negative rays of a bi-infinite contracting quasi-geodesic determine respectively two boundary –classes denoted by and .
Examples 2.15.
The first three convergence boundaries below are equipped with a maximal partition (that is, –classes are singletons).
-
(1)
Hyperbolic space with Gromov boundary , where all boundary points are non-pinched.
-
(2)
CAT(0) space with visual boundary (homeomorphic to horofunction boundary), where all boundary points are non-pinched.
-
(3)
The Cayley graph of a relatively hyperbolic group with Bowditch or Floyd boundary , where conical limit points are non-pinched.
If is infinitely ended, we could also take as the end boundary with the same statement.
-
(4)
Teichmüller space with Thurston boundary , where is given by Kaimanovich-Masur partition [41] and uniquely ergodic points are non-pinched.
-
(5)
Any proper metric space with horofunction boundary , where is given by finite difference partition and all boundary points are non-pinched. If is the cubical CAT(0) space, the horofunction boundary is exactly the Roller boundary. If 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 , where all boundary points are non-pinched.
Proof.
Only the “moreover” statement in Assumption (C) requires clarification. Under the finite difference relation , if is –bounded for some , then any is –bounded as well. According to the topology of the horofunction compactification, if and , then . This shows (however, may not be –bounded). ∎
If is a contracting element, the positive and negative rays of a contracting quasi-geodesic determine two boundary points denoted by and respectively. Write . We say that is non-pinched if 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 is a discrete group, which we assume in the next results.
Lemma 2.17.
If is a non-pinched contracting element, then coincides with the set stabilizer of the two fixed points .
Proof.
By Assumption A, the two half ray of the axis accumulates either to or to . The same holds for , which accumulates to . By definition, is the set of elements which preserves up to finite Hausdorff distance. Let preserve but . Choose two sequence of points so that , and let and . We claim that both and are unbounded sequences. If not, assume for concreteness that is bounded. Choose tending to (, if is bounded). The contracting property implies , where is the contracting constant of . As is bounded and tend to , this contradicts by Definition 2.13(C). ∎
Lemma 2.18.
[72, Lemma 3.12] If are two non-pinched independent contracting elements, then either or .
Lemma 2.19.
[72, Lemma 3.10] Let be two independent contracting elements. Then for all , is a contracting element so that tends to and tends to .
The following lemma is often used to choose an independent contracting element.
Lemma 2.20.
Let be a non-elementary discrete group. Assume that is a contracting element. Then there exists a contracting element that is independent with .
Proof.
The set of translated axis has bounded intersection. Pick two distinct conjugates denoted by of . By Lemma 2.19, is a contracting element whose fixed points tend to and respectively for any large . Note that the axis of has an overlap with and of a large length for . Taking into account every pair of elements in has bounded intersection, we conclude that has bounded intersection with every element in . That is, is independent with by definition. ∎
Let be a subset of . The limit set of , denoted by , is the set of all accumulation points of in . If is a subgroup, may depend on the choice of the basepoint , but the –loci does not by Definition 2.13(B). For this reason, we shall refer as the limit set of a subgroup . Moreover, the limit set of a discrete group enjoys the following desirable property as in the general theory of convergence group action.
Lemma 2.21.
[72, Lemma 3.9] Assume that is non-elementary. Let be a contracting element. Then for any .
2.5. Quasi-conformal density and Patterson’s construction
Let be a convergence compactification, with a nonempty set of non-pinched points in Definition 2.13. We need to restrict to a smaller subset of , on which some weak continuity of Busemann cocycles can be guaranteed.
Assumption D.
There exists a subset with a family of Buseman quasi-cocycles
so that for any , we have
(8) |
where may depend on but not on .
For a given , let denote the set of points for which (8) holds. Thus, .
In practice, the assumption (8) could allow to take the following concrete definition:
The following facts shall be used implicitly.
(9) |
Horofunctions in convergence boundary
We now define Buseman quasi-cocycles at a –class . Namely, given , define a Busemann quasi-cocycle at (resp. ) as follows
where the convergence (resp. ) takes place in . The equations in (9) hold for , and moreover, for all ,
(10) |
Horoballs at –classes
We give the following tentative definition of horoballs at a –class in a general convergence boundary. Given a real number , we define the horoball of (algebraic) depth . We take the convention that the center has depth .
Definition 2.22.
Let for . We define a horoball centered at of depth as follows
We omit if or it does not matter in context.
As the limit supper of continuous functions, is lower semi-continuous, so is a closed subset. By abuse of language, we say the level set is a horosphere or the boundary of .
Lemma 2.23.
For , we have
-
(1)
for .
-
(2)
has distance at least to the complement .
Proof.
(1). Observe that, for any and , there exists a geodesic ray starting at so that for any . ( is called a gradient line in some literature). By an Ascoli-Arzela argument, is obtained as a limiting ray of when . Thus, any point has a distance to some .
(2). If for some and , we obtain so . The conclusion follows. ∎
By abuse of language, we also call the set a horoball in any . In other words, all points in the same class share a common horoball.
Let be the set of finite positive Radon measures on , on which acts by push-forward: for any Borel set ,
Definition 2.24.
Let . A map
is a –dimensional, –quasi-equivariant, quasi-conformal density if for any and ,
(11) | ||||
(12) |
for a universal constant . We normalize to be a probability: its mass .
If is supported on non-pinched boundary points (i.e.: ), 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 is supported on the conical points. Take for instance the horofunction boundary , where is the whole boundary. Moreover, all Patterson-Sullivan measures are non-trivial quasi-conformal density in Examples 1.9.
Patterson’s construction of quasi-conformal density
Fix a basepoint . Consider the orbital points in the ball of radius :
(13) |
The critical exponent for a subset is independent of the choice of :
(14) |
which is intimately related to the (partial) Poincaré series
(15) |
as diverges for and converges for . The action is called of divergence type (resp. convergence type) if is divergent (resp. convergent) at .
Fix . The family of the annulus-like sets of radius centered at
(16) |
covers with multiplicity at most . It is useful to keep in mind that for
(17) |
Fix . We start by constructing a family of measures supported on for any given . Assume that is divergent at . Set
(18) |
where and . Note that is a probability measure supported on . If is convergent at , the Poincaré series in (18) needs to be replaced by a modified series as in [54].
Choose such that for each , is a convergent sequence in . The family of limit measures () are called Patterson-Sullivan measures.
By construction, Patterson-Sullivan measures are by no means unique a priori, depending on the choice of and . Note that is a normalized condition, where is a priori chosen basepoint. In what follows, we usually write for (i.e.: ).
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 of a closed oriented surface (). The space of measured foliations on , which is homeomorphic to , admits a –invariant ergodic measure, , called Thurston measure by the work of Masur and Veech. Positive reals act on by scaling, and let
be the natural quotient map.
is also called the Thurston boundary, and it gives a convergence boundary for with equipped with the Teichmüller metric. Here we take the Kaimanovich-Masur partition , whose precise description is not relevant here, but the set of uniquely ergodic points in is partitioned into singletons. It is a well-known fact that is supported on .
The induces a –dimensional, –equivariant, conformal density on as follows. Given , consider the extremal length function
which is square homogeneous, . Take the ball-like set For any , set . Direct computation gives
The family forms a conformal density of dimension : the term in the right-hand bracket coincides with for , and charges the full measure to . Note that this family is the same as the conformal density obtained from the action through the Patterson construction above (see [72]).
Theorem 2.27.
Suppose that acts properly on a proper geodesic space compactified with horofunction boundary . Then the family of Patterson-Sullivan measures is a -dimensional -equivariant conformal density supported on .
In the sequel, we write PS-measures as shorthand for Patterson-Sullivan measures.
Let be a nontrivial –dimensional –equivariant quasiconformal density on a convergence compactification : is supported on the set of non-pinched points in Definition 2.13(C).
2.6. Shadow Principle
Let be a set of three (mutually) independent contracting elements (), which form a -contracting system
(19) |
where the axis in (6) is -contracting with depending on the choice of the basepoint . We may often assume is large as possible, by taking sufficiently high power of . The contracting constant is not effected.
First of all, define the usual cone and shadow:
and be the topological closure in of .
The partial shadows and cones defined in the following depend on the choice of a contracting system in (19).
Definition 2.28 (Partial cone and shadow).
For , the –cone is the set of elements such that is a –barrier for some geodesic .
The –shadow is the topological closure in of the cone .
The follow terminology is from Roblin [60].
Definition 2.29.
We say that satisfies Shadow Principle over a subset if there is some large constant so that the following holds
for any and , where the implied constant depends on .
The most fundamental example is provided by the Sullivan’s Shadow Lemma, where could be taken to be any –cocompact subset (by –equivariance, thus remains uniformly bounded over ). Here are some other examples.
Examples 2.30.
-
(1)
Roblin realized that could be enlarged to , provided that the normalizer is a discrete subgroup.
-
(2)
Let be a rank-1 symmetric space. If is the Hausdorff dimension of the visual boundary , there exists a unique (up to scaling) –dimensional –equivariant conformal density on . It satisfies the Shadow Principle over the whole space , where for any .
-
(3)
We shall establish the Shadow Principle for confined subgroups in §7.
- (4)
We now recall the following shadow lemma on the convergence boundary.
Lemma 2.31 ([72, Lemma 6.3]).
Let be a nontrivial –dimensional –equivariant quasi-conformal density for some (i.e. supported on the set of non-pinched points). Then there exist with the following property.
Assume that for each . For given , there exist such that
for any .
Remark 2.32.
In what follows, when speaking of , we assume satisfy the Shadow Lemma 2.31.
2.7. Conical points
We now give the definition of a conical point. Recall that 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 is called –conical point if for some , the point lies in infinitely many –shadows for . We denote by the set of –conical points.
We also denote by the set of conical points for which there exists satisfying for some large . By definition, .
The following is the Borel-Cantelli Lemma stated in terms of conical points.
Lemma 2.34.
Let be a probability measure on that satisfies the shadow lemma as in Lemma 2.31, and let be a subset so that the partial Poincaré series . Then the following set
is a –null set, for any .
Proof.
By definition, we can write where
If then for all . On the other hand, the shadow lemma implies
so the tails of the series 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 :
-
(1)
is visual: for any basepoint there exists a geodesic ray starting at ending at .
-
(2)
If is of divergence type, charges full measure on and every –class is –null.
-
(3)
Let be two sequences tending to in . Then for any ,
In particular, is a subset of (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 . Consider a horoball centered at defined as Definition 2.22. Then any escaping sequence has the accumulation points in .
Proof.
According to definition, there exists a sequence of elements so that . By [72, Lemma 4.4], there exists a geodesic ray starting from terminating at so that
where and .
Fix any , and consider the exit point of at . Let as . Given , the following holds for by Lemma 2.35(3):
(20) |
We claim that the projection of all but finitely many to lies in a –neighborhood of . Indeed, if not, holds for infinitely many . The contracting property of implies that intersects for all , so . Combined with (20), we obtain . This is a contradiction, as is a unbounded sequence in a fixed ball around .
If is assumed, the above claim implies the corresponding projections of and to have distance at least , so the contracting property shows . That is, for infinitely many . As is chosen arbitrarily along , the assumption (B) in Definition 2.13 implies that and have the same limit, so . ∎
Uniqueness of quasi-conformal measures
Suppose that is of divergence type, where is compactified by the horofunction boundary . The quasi-conformal measures on are in general neither ergodic nor unique. To remedy this, we shall push the measures to the reduced horofunction boundary , modding out the finite difference relation. However, 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) |
where the intersection is taken over all possible choice of three independent contracting elements in . By [72, Lemma 2.27 & Lemma 5.6], the quotient map
is a closed map with compact fibers, thus is a completely metrizable and second countable topological space.
Let be a –dimensional –equivariant conformal density on . This is pushed forward to a –dimensional –quasi-equivariant quasi-conformal density, denoted by , on .
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 , let be the set of elements such that there exists some geodesic between and with the property that the interior of lies outside .
Definition 2.38 (SCC Action).
If there exist positive constants such that , then the proper action of on is called statistically convex-cocompact (SCC).
Remark 2.39.
The motivation for defining the set 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 , the set 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 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 is a non-elementary SCC action with contracting elements. Then has purely exponential growth: for any , .
For the remainder of this subsection, assume that is an SCC action with a contracting element. Fix a set of contracting elements in .
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 , a –interval of a geodesic segment means a connected subsegment of with length .
Definition 2.41.
Fix constants . We say that a geodesic is –contracting at –frequency if every –interval of contains a segment of length that is –close to a –contracting geodesic.
A geodesic ray is –contracting at –frequency if any sufficiently long initial segment of (i.e. for ) is –contracting at –frequency.
Furthermore, is frequently –contracting if it is –contracting at –frequency for any .
The above notion is not used in this paper, but motivates the following analogous notion involving a proper action .
Definition 2.42.
Fix and and . We say that a geodesic contains –barriers at –frequency if for every –segment of has –barriers.
An element has –barriers at –frequency if there exists a geodesic between and such that has –barriers at –frequency.
Let denote the set of elements in having no –barriers at –frequency for some . That is, an element of belongs to if and only if any geodesic between and contains a –interval that has no –barriers.
Lemma 2.43.
[57, Lemma 4.7] Fix . Then is growth tight.
Set . Let denote the set of limit points that is contained in infinitely many shadows at elements in . More precisely, is the limit superior of the following sequence, as ,
where is defined in (16). Hence,
Recall for an integer . Define the set of regularly contracting points as follows
We could take a further countable intersection over all possible 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 –regularly contracting rays lies in for certain . The following result thus implies [57, Theorem A], saying that –regularly contracting rays is –full.
Proposition 2.44.
[57] Assume that is an SCC action with contracting elements. Then the regularly contracting limit set is a –full subset of .
Proof.
Remark 2.45.
If 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 . In particular, is not necessarily a proper action with discrete orbits.
Assume admits a measure class preserving action on . We say that the action is (infinitely) conservative, if for any with , there are (infinitely many) so that It is called dissipative if there is a measurable wandering set : for each . If in addition, , then it is completely dissipative.
Let be the union of purely atomic ergodic components, and . Denote (resp. ) the subset of with trivial (resp. nontrivial) stabilizers, the subset of with infinite stabilizers. The partition and (mod 0) gives the Hopf decomposition as the disjoint union of conservative and completely dissipative components. If is proper on a hyperbolic space, consists of parabolic points with infinite stabilizer. If is torsion-free, parabolic points are countable.
By [40, Theorem 29], the infinite conservative part coincides (mod 0) with the following set
We now consider the –equivariant –dimensional conformal density on . We emphasize that is not necessarily contained in , but is assumed to preserve the measure class of , so that the following holds for :
Recall that is supported on .
We do not assume to be a discrete group, so may not be a discrete subset in . We still denote by the accumulation points of the orbit in the convergence boundary . It depends on the basepoint , but does not, because of Assumption (B) in Definition 2.13.
Let be defined in Assumption D as the set of points at which Busemann cocycles satisfy some weak form of continuity. According to the above discussion, is exactly the so-called big horospheric limit point defined as follows.
Definition 2.46.
A point is called big horospheric limit point if there exists such that for some . If, in addition, holds for any given , then is called small horospheric limit point. Denote by (resp. ) 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 –class. Thus, is –saturated. In terms of horoballs (2.22), is a big (resp. small) horospheric limit point if some (resp. any) horoball centered at contains infinitely many with .
Let (resp. ) denote the set of boundary points with nontrivial (resp. trivial) stabilizers in . Summarizing the above discussion, we have the Hopf decomposition for conformal measures.
Lemma 2.48.
Let act properly on and preserve the measure class of . Then is the infinite conservative part, and .
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 be a non-uniform lattice in with . We can put a conformal measure on the boundary at infinity, supported on the set of countably many parabolic fixed points. Indeed, fix any and for given . The value of will be adjusted below. The other points in is then determined by
where which does not depend on (for is invariant under the stabilizer ). Fix so that . Observe that can realize any value in , by adjusting with the following fact for given ,
which, in turn, follows from purely exponential growth of double cosets of the parabolic subgroup (e.g. [36])
As is finite, we are able to achieve that .
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 and .
We shall prove that for subgroups in a hyperbolic group, where is the Patterson-Sullivan measure for .
At last, we collect the notations of various limit sets studied in this paper:
-
•
is the limit set of , while is the –classes over .
-
•
is the set of –conical points, and denotes the set of usual conical points.
-
•
(resp. ) is the big/small horospheric limit set.
-
•
is the Myrberg limit set, and is the regularly contracting limit set.
They are related as
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 is a Gromov hyperbolic space and , , are subgroups of with and being discrete. Let be Patterson-Sullivan measures on the Gromov boundary constructed from . If is of divergence type, then 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 is Gromov hyperbolic, there exists such that any geodesic triangle admits a –center, that is, a point within a –distance to each side.
3.1. Preliminary
The notion of an admissible path (Def. 2.6) is a generalization of the –local quasi-geodesic paths:
where any two consecutive paths give a quasi-geodesic by (BP), of length at least (LL). It is well known that, in hyperbolic spaces, if , a local quasi-geodesic is a global quasi-geodesic.
Assume that 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 of three loxodromic elements of and constants that depend only on with the following property. For any , there exists an element such that the path
is a –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 is proper and acts on cocompactly. The goal is to prove the conclusion of Theorem 4.2 in this setting.
Lemma 3.2.
Let be a big horospheric point. Then for any , there exist infinitely many orbit points such that , where is a –center of for some .
Proof.
By the definition of a big horospheric point, there is an infinite sequence of points such that lies in some horoball for all and for some depending on . Recall the definition of a horoball: , so we obtain
for with sufficiently large. By hyperbolicity, if is a –center of for some , then . Combining these inequalities gives
yielding . Thus, . Note however that depends on . Given any , we can still obtain after dropping finitely many . The proof is complete. ∎
Theorem 3.3 (Theorem 1.10 in hyp. setup).
If , then .
Proof.
Fix a small , so . Let be the set of points that satisfy , where is a center of the triangle for some and some .
Choose a sufficiently large constant such that the Shadow Lemma 2.31 applies. If is a maximal –separated net in , then by Lemma 3.2, is contained in the limit superior of shadows. As the action is cocompact, we may assume upon increasing again.
Note that, for each , the corresponding in the first paragraph is lying on the –neighborhood of . This neighborhood contains at most points in the –separated set where is a universal constant.
By the Shadow Lemma 2.31 for the –conformal density, we compute
which is finite by definition of and . The Borel-Cantelli lemma implies that is –null. Hence, , 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 has full Patterson-Sullivan measure. One of the proofs will work whenever 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 ( Lemma 5.7).
Let be a finite set, so that no element in fixes pointwise the limit set . Then there exists a finite set of loxodromic elements and a constant with the following property.
For any , one can find so that for any , the word labels a quasi-geodesic:
Proof.
Let . By assumption, as all fixed points of loxodromic elements are dense in , moves the two fixed points of some . If moves for another , then must move the two fixed points of , which tends to and as . As is a finite set, a common could be chosen for each : . Equivalently, has bounded projection with for each .
With a bit more effort, we produce where moves the two endpoints of each , and the set of axis has bounded projection. This implies that each has bounded projection to at least two of . Consequently, for any , there exists a common so that
Set . The word labels an –local –quasi-geodesic path, denoted by , where depends on . This concludes the proof. ∎
The following is an immediate consequence of Lemma 3.4 applied to , and is chosen for according to the definition of confined subgroups.
Lemma 3.5 ( Lemma 5.8).
Under the assumption of Lemma 3.4, for any , there exist and such that lies in and
Recall that a point is a conical point if there exist a sequence of so that lies in a –neighborhood of a geodesic ray .
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 is a torsion-free discrete subgroup confined by with a compact confining subset . Then the big horospheric limit set contains all but countably many points of .
Proof.
We denote by the countable union of fixed points of all loxodromic and parabolic elements . We shall prove that any conical point is contained in . By definition, there exists a sequence of so that lies in a –neighborhood of a geodesic ray .
Set . The confined subgroup implies the existence of so that . Now, let so that . Thus, . Direct computation shows that lies in the horoball (or see Lemma 6.1 for this general fact).
A horoball only accumulates at the center . As the orbit is discrete, it suffices to prove that is an infinite subset. Arguing by contradiction, assume now that is a finite set. By taking a subsequence, we may assume that for any .
By assumption, is torsion-free, so must be of infinite order. According to the classification of isometries, is either hyperbolic or parabolic, so is .
As , the convergence implies that and then fixes . So gives a contradiction, which completes the proof. ∎
We now demonstrate how to use Lemma 3.5 to deal with finite confining subset .
Theorem 3.7 (Theorem 1.11(ii) in hyp. setup).
Assume that is a discrete subgroup confined by with a finite confining subset . Assume that no element in fixes pointwise . Then the big horospheric limit set contains .
Proof.
Given , there exists a sequence of such that . By Lemma 3.5, we can choose and so that labels a quasi-geodesic path.
If for any , then follows by Morse Lemma. Thus is an infinite subset. Setting
we argue exactly as in the proof of Theorem 6.5 and obtain that lies in the horoball . Hence, converges to , so 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 is assumed to be SCC, in particular it is of divergence type.
-
•
Let be a convergence boundary for and be the unique quasi-conformal, –equivariant density of dimension on .
-
•
Let be a discrete subgroup, which is confined by with a compact confining subset in .
We emphasize that is not necessarily contained in , but preserves the measure class of . This is motivated by the following example. If is the rank-1 symmetric space equipped with the Lebesgue measure on the visual boundary, any subgroup preserves the measure class of . More generally, we have.
Lemma 4.1.
Suppose that is a hyperbolic space on which acts properly and co-compactly. Let be the unique Patterson-Sullivan measure class of dimension on . Then preserves the measure class of .
Proof.
It is known that coincides with the Hausdorff measure of with respect to the visual metric. Namely, if is the visual metric for parameter , then where 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 contains no torsion, the set 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 , provided that .
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 is a subset in (Assumption D) on which the Busemann cocycles converge up to an additive error in (8).
Theorem 4.2.
If , then . In particular, if is torsion-free, the action of on is completely dissipative.
Proof.
By Proposition 2.44, has full measure on , and is a subset of . By taking the intersection , we may assume in addition that every point is a frequently contracting point. This is only the place in the proof where we use the SCC action.
For any sufficiently small , instead of Lemma 3.2, we now prove
Claim.
Let . There exist two sequences of elements such that
(22) | |||
(23) | |||
(24) |
Proof of the claim.
As is a big horospheric point, there exists tending to , where depends on . That is, where . Setting , we obtain from (8):
(25) |
Fix a big to be decided below, which also depends on .
Choose a point so that for . Consider the initial segment of and one –interval at the middle satisfying and . By definition of , any –interval of contains –barrier , so
where is a contracting element. Up to enlarge , we assume so (23) is satisfied.
Look at the triangle with vertices for . See Fig. (6). By the contracting property, intersects non-trivially either or . Indeed, if this is false, the projection of and to together has diameter at most . Meanwhile, the corresponding initial and terminal segments of entering and leaving both project to with diameter at most as well. Hence, we obtain . On the other hand, noting the above , we would get a contradiction if .
Moreover, as , the above argument further shows that intersects non-trivially either or .
We now claim that . Indeed, if not, we have for some and . Thus,
where the last inequality uses (25). On the other hand, as , we have
If is assumed, this would give a contradiction, so the above claim is proved.
By the claim, let us choose . Thus, we have , so (24) is satisfied.
Now we choose so that . Let be the set of points with the above property in the above Claim. Note that is contained in the limit superior of . Note again that is over used by at most times for a universal constant , as in (24) lies in a fixed neighborhood of . Shadow lemma hence allows to compute
Hence,
by Borel-Cantelli Lemma. Thus, implies , 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 be a locally compact, metrizable, topological group and . We say that is confined by if there exists a compact subset in such that intersects non-trivially with any conjugate for :
We refer as a confining subset for . If , we say that is a confined subgroup in .
In this paper, we are mainly interested in the case where or .
Note that any nontrivial normal subgroup is confined. Here are a few remarks in order.
Remark 5.2.
If is the confining subset, so is for . If is finite, we may further assume that is minimal: for any , there exists so that .
If is a confined subgroup in , then any subgroup containing is also confined. Note that being confined is inherited to finite index subgroups. Indeed, if is a finite index subgroup of , then any element admits a sufficiently large power in , so is a confined subgroup in .
Let be a discrete subgroup of . We define the injectivity radius of the action at a point as follows:
where is finite by the proper action. By definition, for , so the injectivity radius function descends to the quotient . When is torsion free, this coincides with the usual notion of injectivity radius, that is the maximal radius of the disk centered at in injecting into . It is clear that for any .
Lemma 5.3.
Assume that acts cocompactly (and not necessarily properly) on a –invariant subset . Denote . Then the subset of has injectivity radius bounded from above if and only if is confined by .
Proof.
(1). Suppose that is bounded from above by over the set . That is, for any , there exists such that . Fixing a basepoint , as acts cocompactly on , there exists a constant such that . Given , let so that , so we have . The set is compact, so is confined by .
(2). Suppose that is confined by with a confining subset , we need to bound . For any , the cocompact action of on implies the existence of such that . As is confined by , there exist and such that . Thus, we have . Setting completes the other direction of the proof: over . ∎
A direct corollary is the following.
Lemma 5.4.
Assume that a group acts cocompactly on . Then for a discrete subgroup , the following are equivalent:
-
(1)
is confined by ;
-
(2)
is confined in ;
-
(3)
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 of , where the action of extends to by homeomorphisms and preserves the partition (see Definition 2.13). Assume that is a non-elementary discrete subgroup of with non-pinched contracting elements.
Definition 5.5.
Let be a group. The elliptic radical consists of elements that induces an identity map on the limit set :
Note that may not fix pointwise . If we write .
Remark 5.6.
Recall that, given a contracting element , is the maximal elementary subgroup in that contains . By Lemma 2.17, if is non-pinched, then is the set stabilizer of in .
If , is the intersection of the maximal elementary subgroups (in Lemma 2.2) for all contracting elements . This is the unique maximal finite normal subgroup of .
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 be a finite set disjoint with . Then there exists a finite set of contracting elements and constants with the following property. Set .
For any , one can find so that for any , the word labels an –admissible path with associated contracting subsets and . Moreover,
Proof.
First of all, fixing a non-pinched contracting element , we claim that for any , there exists so that .
Indeed, if not, then for any and for any . By Lemma 2.21, the set of –translates of its fixed points is dense in : . To be more precise, for any , there exists such that for some . By assumption, . According to Definition 2.13(C), on forms a closed relation, this implies : that is, fixes every –class in , so . This gives a contradiction. Hence, we see that does not stabilize for any .
By Lemma 2.19, if are non-pinched contracting elements in , then for any is a contracting element so that the attracting fixed point tends to and the repelling fixed point tends to as . By the previous paragraphs, for any , we can find two non-pinched contracting elements , so that does not fix and does not fix . As and are disjoint closed subsets in by Definition 2.13(A), choose disjoint open neighbourhoods such that Lemma 2.19 shows that do not stabilize the fixed points of for any . We do not conclude here that , as is the subgroup in (not in ) stabilizing .
As is a finite set, a repeated application of the previous paragraph produces an infinite set of independent contracting elements such that for any and .
Now to conclude the proof, it suffices to invoke the same arguments in the proof of Lemma 2.10. Namely, for any finite set , there is a bounded intersection function so that for any and any ,
The function depends only on the configuration of the axis for . That is to say, remains the same if is chosen as a different set of representatives for . Recall that is the maximal elementary subgroup in defined in Lemma 2.2.
By the contracting property, for any and any given , the diameter is comparable with . Hence, with at most one exception ,
(26) |
and, since for any and a finite set of ,
(27) |
where is a constant depending on the axis for as above. Consequently, for any , there exists a common satisfying (26) so that
(28) |
Setting , the above equations (26)(27)(28) verify that for any , the word labels an –admissible path, denoted by , associated with the contracting sets and . Taking a high power of if necessary, the constant can be large enough to satisfy Lemma 2.10, so any geodesic with same endpoints as –fellow travels , so and have at most a distance to . Letting concludes the proof of the lemma. ∎
5.4. First consequences on confined subgroups
Until the end of this subsection, consider a subgroup , which is confined by with a finite confining subset . Assume that . The main situations we keep in mind are or .
The following is an immediate consequence of Lemma 5.7 applied to , and is chosen for according to the definition of confined subgroups.
Lemma 5.8.
There exists a finite subset of contracting elements in with the following property: for any , there exist and such that lies in and
where depends only on and .
This allows us to derive the following desirable property.
Lemma 5.9.
There exists a finite set of pairwise independent contracting elements in with the following property. Fix a conjugate for . For any there exists such that and their product is a contracting element. Moreover, any two such for distinct pairs are independent.
Proof.
For , applying Lemma 5.8 to with yields the choice of in so that lies in . It remains to note that is contracting. To this end, as and have bounded intersection, setting , we have that the power labels an –admissible path as follows
where the associated contracting subsets are given by the corresponding translated axis of and . Choosing sufficiently large, the path is contracting, so is contracting.
Finally, if and for , then we need to show that and are independent. That amounts to proving that their axes and have infinite Hausdorff distance. Indeed, if not, then lies in a finite neighborhood of , so they fellow travel a common bi-infinite geodesic (which could be obtained by applying Cantor argument with Ascoli-Arzela Lemma). For definiteness, say . The fellow travel property by Proposition 2.9 then forces a large intersection of some axis of with the axis of or the axis of . This would lead to a contraction with bounded intersection of . ∎
We get the following corollary on confined subgroups.
Lemma 5.10.
Assume that is a subgroup confined by with a finite confining subset that intersects trivially . Then must be a non-elementary subgroup with a contracting element. Moreover, .
Proof.
Let for . As above, we then choose and so that labels an –admissible path. Consider the sequence of contracting quasi-geodesic , so it follows that where is given by Proposition 2.9. Hence, we proved that , so the Assumption (B) in Definition 2.13 implies that also tends to . The proof is now complete. ∎
6. Conservativity of boundary actions of confined subgroups
6.1. Some preliminary geometric lemmas
We first prepare some geometric lemmas. Recall that is a subset in (Assumption D) on which the Busemann cocycles converge up to an additive error in (8).
Lemma 6.1.
Let be a geodesic ray ending at for some . If is a point that satisfies for some and a real number , then .
Proof.
We now arrive at a crucial observation in the proof of Theorem 6.5.
Lemma 6.2.
Given there exists with the following property. Assume that is –contracting. Consider a distinct axis for some . Let so that intersects in a diameter at least , and lies in a –neighborhood of the projection of to . Then is a contracting element.
Proof.
The –contracting property of implies that any geodesic outside has –bounded projection, so has –bounded projection to . According to the assumption, has –bounded projection to , so this verifies that the path
is an –admissible path with associated contracting subsets . If is sufficiently large, then is a contracting quasi-geodesic, concluding the proof that is contracting. ∎
As a corollary, we have.
Lemma 6.3.
Assume that is an –conical point. Let be a geodesic ray ending at with two distinct –barriers . Let be –close to the corresponding entry points of in and . Then 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 is a proper, geodesically complete, CAT(0) space. Let be a parabolic isometry. If fixes a conical point in for a proper action of on , then has the unique fixed point.
Proof.
By [30], under the assumption on , the fixed point set of a parabolic element has diameter at most in the Tits metric on the visual boundary . A conical point is visible from any other point ; that is, there is a bi-infinite geodesic between and . Thus, the angular metric between and is , so being the induced length metric, the Tits metric from to is at least . Hence, is singleton, completing the proof. ∎
6.2. Proof of Theorem 1.11
Assume that preserves the measure class of . By Lemma 2.35, is supported on the set of conical points . By Lemma 2.35, is a subset of (defined in Assumption D).
The conservativity of the action of on 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 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 without torsion; the second result crucially uses the Lemma 5.8 based on the finiteness of with torsion allowed.
Theorem 6.5.
Assume that is a torsion-free discrete confined subgroup. If is neither hyperbolic nor geodesically complete CAT(0), assume, in addition, that is finite. Then the big horospheric limit set contains a –full subset of .
Remark 6.6.
The proof shows that the conclusion of Theorem 6.5 remains valid for any atomless measure supported on . If is hyperbolic, could be replaced with . In particular, harmonic measures on horofunction boundary which can arise as the hitting measure of a random walk on are among such examples.
Proof.
Let be the countable union of fixed points of all contracting elements . If is hyperbolic or CAT(0), we adjoin into the fixed points of all parabolic elements in , which are countably many by Lemma 6.4. Note that for has no atom. By Lemma 2.35, we will prove the –full set of points is contained in . By definition of an –conical point, there exists a sequence of –barriers on a geodesic ray , where is taken over the finite set . By definition, we have . Passing to a subsequence, we may assume that for all . Moreover, we can assume that are all distinct; otherwise lies in the limit set of the same for , so is fixed by a contracting element . This implies , contradicting the assumption.
Let be a compact confining subset for , so .
Choose so that . Now, let so that . Thus, . According to Lemma 6.1, we see that lies in the horoball .
We shall prove that is an infinite (discrete) subset, which thus converges to by Lemma 2.36. Arguing by contradiction, assume now that is a finite set. By taking a subsequence, the proper action of allows us to assume that for any .
Case 1. is finite. We may then assume that for all . We thus obtain . By Lemma 6.3, is a contracting element for any two . If is of infinite order, gives a contradiction, as any infinite order element in is a contracting element.
Case 2. may be an infinite compact set, but is assumed to be hyperbolic or geodesically complete CAT(0). Note that might be distinct in general. As is torsion-free, must be of infinite order. According to classification of isometries, is either hyperbolic or parabolic.
As , the convergence implies that and then fixes . So gives a contradiction. The proof is complete. ∎
In the following statement, we emphasize that is not necessarily a discrete subgroup. The proof relies on Lemma 5.8 applied with here, which does not use the proper action as well. Compared with Theorem 6.5, may contain torsion elements, but is assumed to be trivial.
Theorem 6.7.
Assume that is a subgroup confined by with a finite confining subset that intersects trivially . Then contains as a subset. Moreover, .
Proof.
By definition of an –conical point, there exists a sequence of –barriers on a geodesic ray . By definition, we have .
By Lemma 5.8, we can choose and so that labels an –admissible path. If for any , then by Lemma 2.12. Thus is an infinite subset. Setting
we argue exactly as in the proof of Theorem 6.5 and obtain that lies in the horoball (the main issue there was proving the infiniteness of ). Hence, converges to by Lemma 2.36, so is a big horospheric limit point. ∎
Proof.
Under the assumptions of Theorem 6.5, we proved that a –full subset of is contained in the horospheric limit set . Taking a countable intersection allows to assume that is –invariant. If is hyperbolic or CAT(0), it is well-known that the limit set is a –invariant minimal subset. Hence, the topological closure of recovers , thus verifying , so the proof is completed in this case.
Otherwise, is a finite set by assumption. As is assumed, the conclusion follows immediately from Lemma 5.10. ∎
To conclude this subsection, we record a stronger statement, provided that is a normal subgroup. The proof strategy is due to [27].
Theorem 6.9.
Suppose that is an infinite normal subgroup of . Then contains as a subset.
Proof.
Given , let be a geodesic ray starting at and ending at . Let us take a contracting element , as an infinite normal subgroup contains infinitely many ones. By definition of in (21), contains infinitely many –barriers for any . By normality of , we have forms a sequence of elements, which enters any given horoball based on . With Lemma 2.36, this implies that . ∎
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 is assumed to be of divergence type with contracting elements.
-
•
Let be a convergence boundary for and be a quasi-conformal, –equivariant density of dimension on .
-
•
Let be a discrete subgroup of a group , confined by , with a finite confining subset that intersects trivially .
Recall that is the –class stabilizer of in (Def. 5.5).
In this section, let be a set of three independent contracting elements, and the constants given by Lemma 5.7.
7.1. Shadow Principle for conformal density supported on boundary
Below, the constant is given in Definition 2.24, and is the convergence error of the horofunction in Assumption D. Note that for the horofunction boundary. By Corollary 6.8, we have .
Lemma 7.1 (Shadow Principle).
Let be a –dimensional –quasi-equivariant quasi-conformal density supported on . Then there exists such that
for any and .
Proof.
We start by proving the lower bound, which is the key part of the proof. Write explicitly and for . Let .
1. Lower Bound. Fix a Borel subset such that . For given , there exists a sequence of points such that . Since is a finite set, up to taking a subsequence of , there exist and given by Lemma 5.7 such that for any , is an –barrier for any geodesic . This implies , which tends to . In addition, we can choose according to definition of the confined subgroup so that . By definition of the shadow,
(29) |
Note that depends on , and consists of three elements.
Consequently, the set can be decomposed as a disjoint union of three sets : for each , there exist such that ; equivalently,
(30) |
Denote . As there exists by quasi-conformality (12) such that
(31) |
We apply the –quasi-equivariance (11) with to the right-hand side:
(32) |
Combining together (30), (31) and (32), we have
where the last line uses the –invariance of by Lemma 5.10.
Denote . We now conclude the proof for the lower bound via quasi-conformality (12):
2. Upper Bound. Fix . Given , there is a sequence of tending to such that for . Since Buseman cocyles extend continuously to the horofunction boundary, we obtain
The upper bound is given as follows:
The proof of lemma is complete. ∎
7.2. Shadow Principle for conformal density supported inside
Fix a basepoint and . Given , define supported on , we have
Lemma 7.2.
There are constants such that, given any and, we have
Proof.
Let and . According to Lemma 5.7, for given and , we pick such that for any , the word labels an –admissible path, where is an –barrier for . This implies , or equivalently . As can be arbitrarily chosen, the definition of confined subgroup allows us to choose for the element so that lies in and so lies in .
Set . If we have . We claim that this assignment
is uniformly finite to one. Indeed, assume for some . Applying multiplication by on the left for gives . As mentioned above, these two words label two –admissible paths with the same endpoints. Noting that , Lemma 2.12 implies . Consequently, at most elements have the same image under the map . Setting thus completes the proof. ∎
The following is the Shadow Principle we want for all .
Lemma 7.3.
Fix any . Let be a –dimensional –quasi-equivariant quasi-conformal density supported on . Then there exists such that
for any and .
Proof.
If we have , so . Thus, it remains to prove the lower bound. As the triangle inequality implies , it suffices to show for some .
By definition,
Assume that , and set for simplicity. Note that . We remark, however, that may not hold, as is only –conformal density but may not be in .
By Lemma 7.2, we have
Here, . Setting , this is equivalent to the following
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 has purely exponential growth.
We first give a uniform upper bound for all conjugates of a confined subgroup .
Lemma 8.1.
There exist constants so that the following holds
for any .
Proof.
By Lemma 5.9, there exists a finite set of contracting elements such that any contains a set, denoted by , of three pairwise independent contracting elements from . Applying Lemma 2.10 with , the conclusion follows by the same argument as in [71, Proposition 5.2], where does not depend on since there are only finitely many choices of independent of . We include the sketch below and refer to [71, Proposition 5.2] for full details.
Set . The idea is to prove the following for any
where do not depend on . This is proved by considering the map
where is chosen by Lemma 2.10. In particular, labels admissible path so where depends on . We proved there (also see Lemma 2.12) that fails to be injective, only if is greater than a constant depending only on . This constant determines the value of , so the proof of the above inequality is proved. ∎
Lemma 8.2.
There exist constants with the following property. For any and any , we have
Proof.
Let us consider the conjugate for given . Define a map as follows
where and are chosen for according to Lemma 5.8. Hence, so we obtain where .
We shall prove that is uniformly finite to one. That is to say, there exists a constant independent of and so that contains at most elements.
Indeed, if then . According to Lemma 5.8, the words and label respectively two –admissible paths with the same endpoints. Thus, any geodesic with the same endpoints –fellow travels them, so and have a distance at most to , where is given by Proposition 2.9. For , we have . Hence, the –closeness then implies that . Thus, if is larger than a constant given by Lemma 2.12, we would obtain a contradiction. The finite number is hence the desired upper bound on the preimage of , which is clearly uniform independent of .
To conclude the proof, since is assumed to have purely exponential growth, we have for some . Setting , we have . The lemma is proved. ∎
An immediate corollary is as follows. If 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 diverges at
In particular, .
Here is another corollary from the proof. For any , there exists such that . This defines a map as follows:
Lemma 8.4.
There exists a subset , on which the above map is injective, so that .
Proof.
The above defined map is exactly the one for in the proof of Lemma 8.2. As is uniformly finite to one and is of divergent type, we can then find with the desired divergence property. ∎
Lemma 8.5.
Suppose that . Then for any and , we have
where the implicit constant does not depend on .
Proof.
We can write for any and any ,
where the implicit constant depends on the width of the annulus, but is independent of .
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 of divergence type is confined by with a finite confining subset that intersects trivially with , then .
The proof presented here is similar to the proof of [72, Theorem 9.1(Case 1)], where is assumed to be normal in . We emphasize that is not necessarily to be contained in . We need some preparation before starting the proof.
Fix . Let be a PS measure on the horofunction boundary , which is an accumulation point of in (18) supported on for a sequence . This is a –dimensional –equivariant conformal density.
As is of divergence type, Lemma 2.37 implies that push forwarding for different 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
Recall that . From the definition of Poincaré series (15), we have
On account of Lemma 2.37, we fix in the following statement, a PS measure on the horofunction boundary . This is a limit point of in (18) supported on for a sequence .
Lemma 8.7.
If , then there exists a constant such that
(33) |
Proof.
For given , let us take the same sequence of (depending on ) so that and converge to an –conformal density and –conformal density respectively. The above relation then gives
(34) |
Push forward the limiting measures and to the quasi-conformal densities on the quotient , which we keep by the same notation. The constant in Definition 2.24 is universal for any with , as the difference of two horofunctions in the same locus of a Myrberg point is universal. Thus, Lemma 2.37 gives a constant independent of :
so the relation (34) implies .
Recall that and may be different limit points of , where is the PS measure fixed at the beginning of the proof. However, applying Lemma 2.37 again gives for any and thus for . Combinning with the above bound , we obtain the desired upper bound on in (33) depending only on . 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 , there exists such that for every , any is contained in at most shadows where .
We are ready to complete the proof of Proposition 8.6.
Proof of Proposition 8.6.
We assume by way of contradiction, so Lemma 8.7 could apply. With the Shadow Principal 7.1, we obtain
for any . The same argument as in [72, Prop. 6.6] proves . Indeed, by Lemma 8.8, that every point are contained in at most shadows for some uniform . Hence,
which by the Shadow Lemma 2.31 gives , and thus gives a contradiction. ∎
8.2. Completion of proof of Theorem 1.13
First of all, the non-strict inequality follows from Corollary 8.3. The difficulty is the strict part of this inequality.
9. Growth and co-growth Inequality
Again, is only assumed to be confined by , but not necessarily contained in . Denote the collection of right –cosets with representatives in . Note that and may be not necessarily contained entirely in . Consider the Hilbert space with over the coset space .
For each , define
Recall that is the convergence radius of the series
Thus, if then ; if then . For each , define
We now prove the following.
Lemma 9.1.
For any , we have .
Proof.
Consider the PS-measure for supported on :
We compute the norm
(35) | ||||
By Shadow Principle 7.3, we have
Given , any element in the following set
is contained at most members from the family of cones , where is a uniform number depending on .
It follows that
From (35), we thus obtain
The last term is the derivative of the Poincaré series , so is finite for . ∎
Let and . Recall that . Note that may be properly contained in the union of over . Summing up the elements of in the same and noting , we have
The Cauchy-Schwartz inequality gives the finiteness on the scalar product of and :
This implies that and thus . 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 acts properly on a proper geodesic space with a contracting element. Fix a basepoint . Without loss of generality, we may assume that the stabilizer of in is trivial: as we can always embed isometrically 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 , let be the set of points that are –closer to than any point in . Namely, if and only if . Equivalently, is a countable intersection of –half spaces defined as follows
This forms a locally finite family of closed sets (via the same proof of [59, Theorem 6.6.13]), so is a closed subset. It is obvious that for , and for . See Fig. 8 for an illustration of these notions.
We remark that the introduction of with negative is an essential novelty here, which is crucial in the Claim 2 in proof of Theorem 10.13.
Dirichlet domain
For , is the so-called Dirichlet domain centered at for the action . It is a fundamental domain in the following sense: and for any . Denote by the quotient map, whose quotient topology is induced by the metric . By [59, Theorem 6.6.7], is homeomorphic to , where the latter is equipped with quotient topology via the restriction . 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 in faces. We here adopt a more general version of bisectors. For , set , and then .
Examples 10.1.
Here are two examples to clarify the notion of bisectors.
-
(1)
In trees, it is readily checked that is not contained in for any . Indeed, the limit set of the former set properly contains that of the latter set in the end boundary.
-
(2)
In a simply connected negatively pinched Riemannian manifold, it can be shown that has finite Hausdorff distance (depending on ) to (using crucially the lower bound of curvature, i.e. the fatness of the comparison triangle). It follows that is quasiconvex and its limit set remains the same for any .
Let denote the closed –neighborhood of a subset in .
Lemma 10.2.
The following holds for any real number .
-
(1)
is star-shaped at : any geodesic for is contained in .
-
(2)
If , then .
Proof.
(1). For any , we wish to prove that any geodesic is contained in . Indeed, if is not in , there exists so that . This implies , a contradiction with .
(2). If , then for some , we have and . So , i.e.: . ∎
We now fix a convergence boundary for , denote by the set of accumulations points of in . By Definition 2.13(B), we have .
The main result of this subsection is the following.
Theorem 10.3.
For any large ,
-
(1)
-
(2)
.
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 be Gromov hyperbolic with Gromov boundary or CAT(0) with visual boundary . Then for any large .
-
(1)
-
(2)
.
The proof is achieved by a series of elementary lemmas.
Recall that a point is visual if a geodesic ray originating at terminates at , i.e. all the accumulation points of are contained in . We emphasize that may not be an accumulation point of , so it is necessary to make statements on –classes rather than boundary points. Any –conical point is visual by Lemma 2.35.
Remark 10.5.
The assumption in the next three results is used to guarantee the following. If , then any limiting geodesic ray of (a subsequence of) accumulates into . If is hyperbolic or CAT(0), this fact holds for any with being Gromov boundary or visual boundary. Therefore, could be replaced with 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 with the following property. Let be not a big horospheric limit point. Then there exist some and a geodesic which is contained in .
Proof.
According to definition, if , some horoball centered at contains only finitely many elements from , so is a finite value. Moreover, the horoball contains some , but for any . Let be the common contracting constant for elements in . Note that is a subset of by Lemma 2.35. For any sequence of , we obtain from (8) for :
That is, setting , for any , so we obtain by definition. See Fig. 9. Now, take a geodesic ray starting at and ending at by Lemma 2.35 (cf. Remark 10.5). If is taken on tending to , the star-sharpedness implies by Lemma 10.2. Thus . ∎
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 for some . Then contains a geodesic ray ending at starting at .
Proof.
By [72, Lemma 4.4], such a geodesic ray exists for any . We briefly recall the proof, and then explain can be taken into by using the star-shaped . First, a conical point by definition lies in infinitely many partial shadows for . By the proof of [72, Lemma 4.4], we can find a sequence of tending to , so that intersects for a common in a diameter comparable with (as ). By Ascoli-Arzela Lemma, the limiting geodesic ray of (a subsequence of) tends to by Definition 2.13(B).
Now take tending to . If , then converges locally uniformly to a geodesic ray by Lemma 10.2. In the general case, if is sufficiently large, the –contracting property of implies that intersects in a diameter comparable with . Hence, a subsequence of converges locally uniformly to a geodesic ray ending at . As we wanted, is contained in the star-shaped . ∎
We now prove the assertion (2) in Theorem 10.3
Lemma 10.8.
Let for some . Then .
Proof.
By Lemma 10.7, take a geodesic ray starting at and terminating at .
By way of contradiction, assume that is a small horospheric limit point. Hence, any horoball centered at contains a sequence of tending to . That is, as . Note that the Busemann cocycle associated to differs from up to a uniform error. This implies that as tends to . Consequently, the definition of shows that for any fixed . This contradicts the choice of in . ∎
10.2. Big and small horospheric limit set
In this and next subsections, assume that
-
•
The proper action is cocompact on a proper hyperbolic space , which is compactified with Gromov boundary endowed with maximal partition (i.e. each -class is singleton, so is omitted in what follows).
- •
Let a countable group (possibly not contained in ) act properly on . 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.
.
The key part is to prove that the set of Garnett points is –null. Recall that a point is called Garnett point if there exists a unique horoball (Definition 2.22) at so that
-
•
The interior of contains no point from : for any ;
-
•
any larger horoball contains infinitely many points of : for infinitely many and any ;
where . Denote by the set of Garnett points for .
We first recall a useful observation of Tukia [67, Appendix].
Lemma 10.10.
The set is –null.
Proof.
We include a short proof for completeness. Indeed, let be a non-Garnett point, so there exists a horoball whose boundary contains a non-empty finite set of elements in but whose interior contains no point in . Consequently, is a disjoint union of measurable subsets indexed over finite subsets in , which are invariant under , so we could choose a measurable fundamental domain for . By Lemma 2.48, is a subset of the conservative part , so . ∎
To complete the proof of Theorem 10.9, it suffices to prove that is a –null set.
Lemma 10.11.
.
Proof.
By way of contradiction, assume that .
We define a height function . Set , where is the unique horoball centered at whose interior does not contain any point of and any horoball centered at larger than that contains infinitely many points of . We verify that is a measurable function. Indeed, if , then ; if , then for any . This implies that the set with is a countable intersection of closed subsets, so is Borel measurable. By definition, is the limit supremum of such measurable sets. Moreover, the set of with is measurable.
As is a doubling measure, the Lebesgue density theorem holds: Let be the –full set of density points , so
-
(1)
is approximately continuous: implies ;
-
(2)
for any sequence of metric balls centered at with radius tending to 0, we have
As in [65], we are going to find a “forbidden” region , consisting of non-Garnett points, with a definite –measure. This contradicts the item (2). The co-compact action is crucial to apply Shadow Lemma 2.31: for any and any fixed large constant , we have
(36) |
For , there exists a unique horoball for some real number , so that for all , but for infinitely many distinct . Let be a projection point of onto , so by definition. As horoballs in hyperbolic spaces are quasi-convex, we have , and .
Denote . Choose with and for a large constant . Set and . See Fig. 10 for approximate tree configuration of these points.
As and , we see that . Shadow estimates (36) implies the existence of a positive constant depending only on :
We claim that for all , consists of non-Garnett points up to a -null set. If not, take a density point . Let be the unique horoball associated with , so follows as above. Let so that . Our goal is to prove must contain , contradicting the definition of a Garnett point.
Indeed, the approximate continuity implies for . As , we have . If , the thin-triangle inequality shows that and thus . Now, by the choice of and , we have , implying
Noting for some , we choose so that . Lemma 6.1 implies . Hence, we conclude that . Recall that , this contradicts the property that 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 on Teichmüller space for . The proof of Lemma 10.11 could be adapted to its subgroup , provided that (36) holds for all . (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 gives the same mass to Thurston boundary , and the volume growth in Teichmüller space does not depend on the basepoint. This provides positive evidence that (36) holds for all .
10.3. Characterizing maximal quotient growth
Let and denote respectively the –ball and –ball at and of radius . Note that : if and only if . We say that has slower growth than if given any maximal –separated subset (resp. ) of (resp. ) for some , we have
This definition is independent of the choice of and , 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 is a simply connected rank-1 manifold with a co-compact action, is comparable to the volume of . 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, is assumed to be geometric action on a hyperbolic space, and is the corresponding Hausdorff measure on the Gromov boundary .
Theorem 10.13.
Assume that acts properly on a proper hyperbolic space . Then the small/big horospheric limit set is –full, if and only if grows slower than . Moreover, is equivalent to the following
As an immediate corollary, we obtain a characterization of conservative actions.
Corollary 10.14.
Assume that is torsion-free. Then the action is conservative if and only if the action grows slower than .
Proof of Theorem 10.13.
Let be the projection. Fix a basepoint and consider the orbit . Then for some as the action is co-compact by assumption. The image is equipped with the metric for , and write
According to (3), we have
Recall . For each , choose so that . That is, is a closest point in to , so is an element in . Thus, every element in admits a (possibly non-unique) lift in the following intersection:
Thus, .
Claim.
If , then and
Proof of the Claim.
Define
Let be the limit supremum of . By construction, coincides with . Indeed, any point , there exists a geodesic ray in ending at . Then we can choose so that . This implies . The converse inclusion follows from the definition of . Hence, as .
As the family of shadows for has uniformly bounded multiplicity, we have by the Shadow Lemma
where the implicit constant depends on . Hence, the claim follows. ∎
Claim.
If , then .
Proof of the Claim.
Recall that is a fundamental domain for , so any point admits at least one translate into . If , we have and then lies on the bisector .
Consider the facet-like set
and the set of accumulation points of in . By Theorem 10.3, the intersection has no small horospheric limit point, so is –null by Theorem 10.9.
Note that for is contained in the interior of , where restricts as an injective map. See Fig. 8. By definition, is the union of with countably many facet-like sets , which are -null, so we obtain .
By injectivity of restricted as map, injects into . Consider similarly
Let be the limit of . As the above claim, the star-shapedness of shows that , and is covered with a uniform multiplicity by the family of for .
Consequently, for some uniform independent of , and
We thus obtain , and then the conclusion follows by . ∎
The proof of the theorem is completed by the above two claims. ∎
The first claim actually does not make use of the assumption that is cocompact and is doubling. These two assumptions are only required in Lemma 10.11. So we have.
Corollary 10.15.
Assume that is a proper action on a hyperbolic space and be the –dimensional –equivariant quasi-conformal density on . Let be a subgroup in . If , then .
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
-
•
admits a proper SCC action on with contracting elements.
-
•
Let be a convergence boundary for and be a quasi-conformal, –equivariant density of dimension on .
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 and form the contracting system , whose stabilizers are accordingly .
Projection complex
For a sufficiently large , the projection complex defined by Bestvina-Bromberg-Fujiwara [9] is a quasi-tree of infinite diameter ([9]):
-
•
The vertex set consists of all elements in .
-
•
Two vertices are connected by one edge if for any with , we have . Here the projection is understood between as axes in .
Remark 11.1.
Any two vertices are connected by a standard path consisting of consecutive vertices
where . See [9, Theorem 3.3]. This does not certifies the connectedness of , but also plays a crucial role in the whole theory. Among the useful facts, we mention
By the rotating family theory, Dahmani-Guirardel-Osin [24] proved that
Theorem 11.2.
There exists a (not possibly every) large so that the normal closure 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 at , so is fixed by . By definition, is generated by all conjugates of in . If is chosen so is normal in , then acts on with the stabilizers exactly from those conjugates of . In particular, the stabilizer of is now .
Construction of free base
We do it inductively and start with and . Let be the –neighborhood of and be the set of vertices up to –conjugacy in . Set and . Then is a fundamental set for the action , i.e. containing exactly one point from each orbit. Here denotes the vertex set of .
For , let and define to be the –neighborhood of . Let be the set of vertices up to –conjugacy in , so that each vertex in is adjacent to some in . This can be done, as . Then is generated by the stabilizers of vertices in .
Canoeing path
Recall that is a large constant chosen for . Any two vertices in can be connected by an –canoeing path for some (see [11, Def. 4.3 & Lemma 4.9]):
-
(1)
is a concatenation where is a geodesic or the union of two geodesics.
-
(2)
The common endpoint of and has –large angle: if are previous and next vertices, then , where are understood as axes in .
By [11, Prop. 4.4], –large angle points are contained in the standard path from to .
In [11], through the canoe path, it is inductively proven that is generated by with a free group generated by . Exhausting by , is obtained as the direct limit of , which is a free group generated by stabilizers of vertices in . Recall that and , and consider the splitting
where is a free group. We also need the associated Bass-Serre tree . More precisely, as the quotient graph is an interval, we partition with the stabilizers of vertices in being conjugate to and the stabilizers of vertices in to . Moreover,
-
•
the set of –vertices (resp. –vertices) in (resp. ) are labeled by left –cosets (resp. –cosets) in ;
-
•
the set of edges adjacent to (resp. ) are bijective to the elements in (resp. ).
11.2. Subgroups with nontrivial conservative part
We now state the main theorem in this section.
Theorem 11.3.
The subgroup has proper limit set . Moreover, If , then .
In particular, since 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 has non-trivial Hopf decomposition.
Remark 11.4.
In the construction, we have that is non-elementary acylindrically hyperbolic, in particular it is non-amenable. The amenability conjecture states that for a normal subgroup , if and only if the quotient group is amenable. If the amenability conjecture is verified for any SCC action, then the assumption that 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 .
The end boundary of the Bass-Serre tree associated to consists of all geodesic rays from a fixed point. As is locally infinite, it is non-compact. Note that for . In what follows, we understand as the quotient space by identifying each –class as one point. Recall that is the set of non-pinched points in (Definition 2.13(C)), and the points outside are pinched.
Proposition 11.5.
We have the following covering
(38) |
where a map 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 .
In a variety of examples, it is known that consists of non-pinched points, so the union in (38) becomes a disjoint union. We record this for further reference.
Corollary 11.6.
As shown in (39), the first points of any contracting elements other than must be contained in some –translate of .
In the next three steps, we shall describe in detail how a geodesic on the Bass-Serre tree produces a canoeing path in the projection complex , which is then lifted to an admissible path in the space . This is the crux of the proof of Proposition 11.5.
Let be a non-empty reduced word over the alphabet . We decompose into nonempty subwords separated by letters in :
(40) |
where and unless the last one . For each , we have a minimal so that represents an element in . Then the element is represented by . If is not in , then and .
Fix an oriented edge in with endpoint stabilizers and respectively. Denote the reversed edge. We consider the basepoint corresponding to the coset in , and the basepoint in stabilized by .
-
Step 1
The word gets unfolded as a geodesic in starting with the edge , at which rotates the edge at to . Rotating edges consecutively in this way, we write down
where the last edge terminates at the endpoint fixed by corresponding conjugate of . See Fig. 11.
-
Step 2
Recall and . As mentioned above, there exists an –canoeing path in from to . These and are adjacent at , around which rotates . By (37), for any , so is a –large angle point of the concatenation . We then obtain an –canoeing path from to as the concatenation of up to appropriate translation,
where the common endpoints of two consecutive paths are –large angle points fixed by the corresponding conjugates of . See Fig. 12.
Figure 12. Canoeing path in : join canoeing paths at large angle points . Those large points are contained in the standard path (possibly ) from to . -
Step 3
By [11, Prop. 4.4], those –large angle points lie on the standard path from to . We lift the standard path to get an –admissible path in as follows, with contracting sets given by the axis where :
Choose two points and in . Connect by concatenating geodesics in consecutive axis and then geodesics from to . See a schematic illustration (13) of a lifted path.
We first establish the desired map mentioned in Proposition 11.5.
Lemma 11.7.
There exists a map so that the image is contained in the conical limit set . Moreover, if for , then is outside the set of non-pinched points.
Proof.
By construction, , so any geodesic ray in from has the list of all vertices , which are alternating left and cosets. This determines a unique infinite word of form (40). By Step (3), the lifted –admissible path in intersects infinitely many translates of under , as contains infinitely many –cosets. By Definition 2.13(B), accumulates into a –class . If we choose a sequence , we see so lies in . Setting defines the desired map .
To justify the claim of injectivity, consider a distinct geodesic originating at with the set of vertices , so that . We shall show that is pinched. Let be the bi-infinite geodesic in the tree from to , i.e.: as a set, where is the vertex that departs from . As shown in the first paragraph, there are two sequences and tending to .
We read off from a bi-infinite alternating word in the form (40). As described in Step (3), we obtain from an –admissible path from to , where the contracting subsets correspond to the –vertices on . By Proposition 2.9, has at most –distance to the entry and exit points of the contracting subsets. See Fig. 3. Recall that is a bipartite graph with –vertices and –vertices. Look at the first –vertex on after , and its exit point of the contracting subset. By the above discussion, we have , that is is non-escaping. Hence, is pinched. ∎
We continue to examine the possible intersection in (38).
Lemma 11.8.
The following hold.
-
(1)
For any , the intersection consist of only pinched points.
-
(2)
intersects only in pinched points.
-
(3)
for any .
Proof.
(1). Without loss of generality, we may represent as a nonempty alternating word as in (40), ending with letters in . Let and tend to the same . We will show ia pinched: that is, intersects a compact set for all .
To this end, consider the nonempty alternating word representing . As and , the word has the form , where gives one letter in . We obtain from , as described in the Step (3) above, an –admissible path , which has the same endpoints as . Note that is the exit vertex of the admissible path . Up to translation, we obtain that has –distance to the fixed point by Proposition 2.9. This shows that lies outside of the non-pinched points .
(2). The same reasoning shows that intersects only in pinched points. Indeed, if lies in , then it is an accumulation point of , where lies in infinitely many –cosets (see the proof of Lemma 11.7). The same argument as above implies that is non-escaping.
(2). The same argument proves that must be pinched. However, since is assumed to be non-pinched points, we obtain . ∎
Proof of Proposition 11.5.
Note that is the union of all left –cosets and –cosets. Consider a limit point in the following complement
so that , for a sequence of belonging to infinitely many distinct –cosets and/or –cosets denoted by . By the above construction of Bass-Serre tree, corresponds to vertices on , so after passing to subsequence, we obtain a limiting geodesic ray in starting at : that is, the intersection of with becomes unbounded as . By the proof of Lemma 11.7, this implies is contained in .
The “moreover” statement follows from Lemma 11.8. The proof is now complete. ∎
Proof of Theorem 11.3.
First of all, as is an infinite normal subgroup, by Theorem 1.11.
By Proposition 11.5, is covered by a countable union with the remaining “exceptional” set :
As is contained in the conical points, so is shadowed by the elements in . By the assumption that , we have is –null by Lemma 2.34. This then proves .
This concludes the proof modulo of the following claim.
Claim.
A small/big horospheric limit points of is either in or a small/big horospheric limit point in a –translate of . That is,
Proof of the claim.
Indeed, let . Then there exists a sequence tending to in some (or any) horoball based at . As , lies in finitely many –cosets, and after passing to subsequence, we may assume for . By Lemma 11.8, for , so we conclude that and thus . ∎
As is –null, we thus obtain from Proposition 11.5 that . ∎
We conclude this section with a corollary which might be of independent interest. Define a topology on as follows (cf. [23]). A sequence of points if and only if the geodesic eventually passes no element in any given finite set of edges (which is required adjacent to if ). This makes homeomorphic to a Cantor space, in particular, it is compact. If is the Bass-Serre tree of , then is exactly the Bowditch boundary of a hyperbolic group relative to . Moreover, if both are one-ended, it is also the end boundary of .
Corollary 11.9.
In the setting of Corollary 11.6, there exists a continuous surjective –map from to , where can be identified with the Bowditch boundary of relative to factors and .
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 .
References
- [1] M. Abért, N. Bergeron, I. Biringer, T. Gelander, N. Nikolov, J. Raimbault, and I. Samet, On the growth of -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 -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, -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., Birkaser, 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 -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.