On the Hausdorff dimension of geodesics that diverge on average
Abstract.
In this article we prove that the Hausdorff dimension of geodesic directions that are recurrent and diverge on average coincides with the entropy at infinity of the geodesic flow for any complete, pinched negatively curved Riemannian manifold. We derive an interesting consequence from this result, we prove that the entropy of a -finite, ergodic and conservative infinite invariant measure is bounded from above by the entropy at infinity of the geodesic flow.
2020 Mathematics Subject Classification:
37C45, 37D40, 37D35, 28A78, 28D20.1. Introduction
In 1984 Sullivan [S] established several groundbreaking results regarding the action of Kleinian groups on hyperbolic space. He proved that the Hausdorff dimension of the limit set of a geometrically finite group is equal to its critical exponent, and that the critical exponent of a convex cocompact group is equal to the topological entropy of the geodesic flow on the quotient manifold. These results highlight the close relationship among the Hausdorff dimension of limit sets, entropy theory and critical exponents. Since then, several generalizations of Sullivan’s results have been obtained. Bishop and Jones [BJ] proved that the Hausdorff dimension of the radial limit set of any non-elementary group coincides with its critical exponent. Shortly after, Paulin [Pau] generalized this result to Kleinian groups acting on Hadamard manifolds of any dimension. Otal and Peigné [OP] proved that the critical exponent of a non-elementary Kleinian group equals the topological entropy of the quotient manifold, where the compactness assumption is removed and the topological entropy is defined by means of the variational principle. In this article we study the relation between the Hausdorff dimension of a dynamically defined subset of the limit set, the diverging on average radial limit set, and the entropy at infinity of the geodesic flow. The entropy at infinity of a dynamical system measures the chaoticity of the system at the ends of phase space and plays a important role in the study of the ergodic theory and thermodynamic formalism of non-compact dynamical systems, for instance see [IRV, RV, ST, Ve, ITV, GST].
Consider a flow on a non-compact topological space. Given a point, we say that it is divergent if its trajectory eventually leaves any compact set, and it is said to be divergent on average if the time the trajectory spends in any compact set is asymptotically zero. The interest in the study of the set of divergent and divergent on average points, or directions in the case of the geodesic flow, is not new and has been a prominent subject of study in homogeneous dynamics and for the Teichmüller geodesic flow. For instance, Masur [Ma] proved that for a given quadratic differential on a genus surface with punctures, the Hausdorff dimension of the set of angles , for which diverges on average for the Teichmüller geodesic flow on the corresponding Teichmuller space is bounded from above by . This is also motivated by the fact that if the measured foliation associated to is minimal and not uniquely ergodic, then the quadratic differential is divergent [Ma, Theorem 1.1]. It was later proved by Apisa and Masur [AM] that the set of angles for which diverges on average is in fact equal to . In the context of homogeneous dynamics, Dani [D] studied the existence and classification of divergent orbits for diagonalizable flows, establishing a correspondence between divergent orbits and singular vectors and matrices. Notably, Cheung [Ch] calculated the exact Hausdorff dimension of divergent orbits for the action of diag() on . In the real rank one situation, Kadyrov and Pohl [KP] obtained both upper and lower bounds for the Hausdorff dimension of the set of points that diverge on average, see also [EKP]. In the higher rank case, for the action of on an upper bound was obtained by Kadyrov, Kleinbock, Lindenstrauss and Margulis [KKLM]. More recently, Das, Fishman, Simmons and Urbański [DFSU] proved that the upper bound obtained in [KKLM] is also a lower bound, and therefore establishing the equality. For countable Markov shifts an upper bound for the Hausdorff dimension of the set of recurrent and diverging on average points has been obtained in [ITV].
In a joint work with G. Iommi [IRV], the authors defined the entropy at infinity of a topological flow on a non-compact manifold as the supremum of limits of measure-theoretic entropies over sequences of measures that exhibit complete lose of mass. In [IRV] we proved that the entropy at infinity of the geodesic flow on extended Schottky manifolds coincides with the maximal critical exponent among parabolic subgroups of the fundamental group. This result was later generalized to any geometrically finite manifold in [RV]. For general pinched negatively curved manifolds the entropy at infinity has been studied from the topological and measure theoretic point of view in [ST, Ve, GST]. For symbolic dynamical system, related notions were already considered by Gurevich and Zargaryan [GZ], Ruette [Ru] and Buzzi [Bu].
In order to make precise statements of our results we start by introducing some notation. For further details we refer the reader to Section 2. Let be a complete, simply connected, pinched negatively curved manifold, and let be a discrete, torsion free, non-elementary group of isometries of . Consider the quotient manifold and let be the projection map between unit tangent spaces. The geodesic flow on is denoted by , where is the time -map. Let be the characteristic function of a subset .
Definition 1.1.
Let . We say that diverges on average if for any compact set , we have
We say that is divergent if for any compact set there exists such that , for any . A vector is recurrent if it is not divergent.
A vector in is said to be divergent on average, divergent or recurrent depending on whether its projection to under the map has the corresponding property. Given in or , we use to denote the set of recurrent vectors based at , and to denote the set of vectors that diverge on average based at . We define
Fix a reference point and denote by the canonical identification that sends to the geodesic ray that starts at with direction . Observe that , where is the radial limit set of . Define the diverging on average radial limit set of by
Since two asymptotic rays approach exponentially fast, the sets and are independent of the reference point . The entropy at infinity of the geodesic flow on is denoted by . The main result of this article is the following.
Theorem 1.2.
Let be a complete pinched negatively curved Riemannian manifold which is non-elementary. Then the Hausdorff dimension of the diverging on average radial limit set is equal to , that is,
where the Hausdorff dimension is computed using the Gromov-Bourdon visual metric on .
It worth noticing that if is hyperbolic, then the map is bi-Lipschitz considering endowed with the hyperbolic metric. In particular, Theorem 1.2 can be stated as
where the Hausdorff dimension is computed with respect to the hyperbolic metric and is any point.
In the case that is geometrically finite with parabolic, then it is known that , see [RV, ST]. In this case we obtain a formula for the Hausdorff dimension of the diverging on average radial limit set. In particular, if is non-compact, finite volume and hyperbolic, then , where . More generally, if is hyperbolic and geometrically finite with cusps, then the Hausdorff dimension is half the maximal rank of parabolic subgroups of .
Following Ledrappier [Led] we consider the entropy of infinite, -finite, ergodic and conservative measures. We denote by the entropy of the measure . As a consequence of Theorem 1.2 we obtain the following result.
Theorem 1.3.
Let be an infinite measure on which is -finite, invariant by the geodesic flow, ergodic and conservative. Then .
Let be the space of geodesic flow invariant -finite Borel measures on that are infinite, ergodic and conservative. Theorem 1.3 can be stated as
Establishing whether the aforementioned inequality equates to an equality is a interesting problem, this would indicate a variational principle for the entropy of infinite measures.
2. Preliminaries
Consider a complete, simply connected, pinched negatively curved Riemannian manifold of dimension . We assume that the sectional curvatures of the metric lie in the interval , for some . The group of isometries of is denoted by and consists of the set of diffeomorphisms that preserve the Riemannian metric. The Riemannian distance of will be denoted by .
2.1. Boundary at infinity
The boundary at infinity, or Gromov boundary of , which we denote by , is the set of equivalent classes of asymptotic geodesic rays on . We endow with the cone topology, which makes it homeomorphic to the unit sphere in . More concretely, consider the map that sends to the class of the geodesic ray based at with initial direction . The map defines a homeomorphism for every .
For and we will denote by
-
•
the geodesic segment starting at and ending at .
-
•
the geodesic ray starting at with end point at infinity .
-
•
the geodesic line having and as end points at infinity.
Given and consider the sets
Topologize the compactification such that sets of the form are a basis of neighborhoods of boundary points. The compactification is homeomorphic to the closed unit ball in .
In this article we are interested in calculating the Hausdorff dimension of subsets of with respect to certain metrics that we proceed to define.
Definition 2.1.
The Busemann cocycle is the map defined by
where is the parametrization of a geodesic ray ending at .
Definition 2.2.
Let . The Gromov product at between is defined as
The Gromov product at between is defined as
Definition 2.3.
Let . The Gromov-Bourdon visual distance on is defined as
Observe that , for any . Every Gromov-Bourdon visual distance induces the cone topology on . Moreover, they are mutually conformal and -invariant. More precisely,
(1) |
and
Given and , we define the shadow at infinity from of the ball as the set
Denote by
where and . For every we denote by to the end point of the geodesic ray based at passing through . The following lemma is due to Kaimanovich (see [Ka, Proposition 1.4]).
Lemma 2.4.
For any there exists such that for any , we have
where .
2.2. The geodesic flow
Let be a discrete, torsion free subgroup of isometries of and consider the quotient manifold . The quotient map is denoted by . Let be a reference point.
Elements of are classified into three types: hyperbolic, parabolic, or elliptic. We say that is non-elementary if it is neither generated by a hyperbolic element nor a parabolic subgroup. We say that is non-elementary if is non-elementary.
The geodesic flow on is denoted by , where is the time -map. A point is said to be non-wandering if for any neighborhood of and any , there exists such that . The set of all non-wandering points is called the non-wandering set of the geodesic flow. The non-wandering set is denoted by . This set is closed and invariant under the geodesic flow, meaning that if , then for all . Furthermore, the Poincaré recurrence theorem implies that if is a geodesic flow invariant probability measure, then . In particular, the support of any invariant probability measure is contained in .
We say is convex cocompact if is compact. We say that is geometrically finite if an -neighborhood of has finite Liouville volume (for a detailed discussion on the notion of geometrically finiteness, see [Bo]).
Let be the space of geodesic flow invariant probability measures on , and the measure-theoretic entropy of . The topological entropy of the geodesic flow is defined as
2.3. Limit sets
The limit set of , denoted by , is the set of accumulation points of a orbit on . More precisely,
The limit set is independent of the reference point . Observe that if there exists a sequence in and a point such that converges to .
A point is called a radial limit point if there exists an -neighborhood of the geodesic ray that contains infinitely many elements of for some . The radial limit set of , denoted by , is the collection of all radial limit points of . It is worth noting that if the projection to of the geodesic ray returns infinitely many times to the ball centered at of radius , for some . In other words, , where is the canonical identification that sends to the geodesic ray that starts at with direction , and is the set of recurrent directions based at .
In the introduction of this article, we introduced a subset of the radial limit set called the diverging on average radial limit set of . This is defined as
where is the set of recurrent and diverging on average vectors based at . Equivalently, if it is a radial limit point and
for every compact set , where is a vector whose associated geodesic ray converges to .
2.4. Critical exponents
The critical exponent of is defined by
This number does not depend on the reference point . Moreover, the critical exponent of is the exponent of convergence of the Poincaré series
Indeed, the Poincaré series converges for and diverges for .
The following two theorems emphasize the strong interconnection between the Hausdorff dimension of the radial limit set, the topological entropy of the geodesic flow, and the critical exponent of the group. Sullivan [S] initially proved the first result for geometrically finite groups. Later, Bishop and Jones [BJ] extended this result to arbitrary Kleinian groups acting on . Finally, Paulin [Pau] extended it further to include Hadamard manifolds of any dimension. The second result, originally established by Sullivan [S] for convex cocompact manifolds, was generalized to arbitrary manifolds by Otal and Peigné [OP].
Theorem 2.5.
Let be a non-elementary pinched negatively curved Riemannian manifold. Then
where the Hausdorff dimension is computed using a Gromov-Bourdon visual metric on .
Theorem 2.6.
Let be a non-elementary pinched negatively curved Riemannian manifold. Then .
In analogy to the critical exponent of we now define the critical exponent outside a compact set of , which quantifies the complexity of the ends of the manifold. Let be a compact, pathwise-connected set that is the closure of its interior and has a piecewise boundary. A nice preimage of is a compact set with a piecewise boundary such that and the restriction of to the interior of is injective. The following lemma was proved in [ST, Lemma 7.5].
Lemma 2.7.
Let be a compact pathwise connected set with piecewise boundary which is the closure of its interior. Then
-
(1)
A nice preimage of always exists,
-
(2)
If , then ,
-
(3)
The set is finite, and
-
(4)
If are compact sets as above, then they admit nice preimages .
Definition 2.8.
Let be a compact set and a nice preimage of . The fundamental group of out of is the set of elements for which there exists such that the geodesic segment touches only at and . In other words, such that
The definition of a fundamental group out of a compact set depends upon the choice of a nice preimage of . The following proposition ([ST, Proposition 7.9 (1) and (3)]) clarifies the dependence of this definition on the choice of . It is worth noting that may not form a group in general.
Proposition 2.9.
Let be a compact pathwise connected set with piecewise boundary.
-
(1)
If , then
-
(2)
If and are nice preimages of , then there exists a finite subset of , such that
Definition 2.10.
Let be a compact pathwise connected set with piecewise boundary. The critical exponent out of is the exponent of convergence of the Poincaré series .
It follows from Proposition 2.9 that is well-defined and independent of the nice preimage . The critical exponent coincides with the exponential growth of the orbits of , that is
If are compact sets with piecewise boundaries such that , then (see [ST, Proposition 7.9 (2)]).
Definition 2.11.
The critical exponent at infinity of is defined by
Observe that if is a sequence of compact pathwise connected sets with piecewise boundaries such that and , then
2.5. Hausdorff dimension
Let be a Borel subset of and . For every and , define
where the infimum is taken over all finite coverings of by balls of radius . Then define
Remarkably, defines a measure on , that is, the Hausdorff measure of dimension with respect to the metric . It is well known that there exists such that
The Hausdorff dimension of with respect to is defined to be equal to . Since and are comparable for every (see equation (1)), the Hausdorff dimension of subsets of is independent of . The Hausdorff dimension of with respect to any of the visual metrics is denoted by .
The following result is a useful tool that allows us to establish lower bounds for the Hausdorff dimension of a given set.
Lemma 2.12 (Frostman’s lemma).
Let be a Borel subset of . Assume that there exists and a positive measure on such that for every and , we have
Then .
We now recall the definition of the Hausdorff dimension of a measure.
Definition 2.13.
Let be a finite Borel measure on . The Hausdorff dimension of is defined by
We remark that the Hausdorff dimension of a measure is sometimes defined where the infimum runs over subsets with total mass; we are mostly interested in ergodic measures, in which case there is no difference. The following result can be found in [Led, Section 2].
Proposition 2.14.
Let be a finite Borel measure on . Then
where the essential infimum is considered with respect to .
The proposition mentioned above, along with Frostman’s Lemma, implies the following variational principle in relation to Hausdorff dimensions.
Theorem 2.15.
Let be a Borel set. Then
where the supremum runs over all finite measures on such that .
3. Piecewise geodesic curves
We say that is a piecewise geodesic curve, where , if it is a continuous map that parametrizes a collection of countably many geodesic segments. In this article we are mostly interested in the case where . In this case, there exists an increasing sequence of real numbers converging to , where and such that is a geodesic segment parametrized by arc-length for every . By definition, the distance between and is given by . The interior angle at is denoted by , where an interior angle is by definition in . Using this notation, we say that has lengths and angles .
Definition 3.1.
Let . A piecewise geodesic curve is called a -quasi-geodesic if for every , we have that
Definition 3.2.
Let and . A piecewise geodesic curve is called a -local-quasi-geodesic if for every such that then .
Let be fixed. In the proofs we will need to shadow a piecewise geodesic curve with angles at least and sufficiently large lenghts by a geodesic ray. To do so, we first need to establish that such a curve is a local quasi-geodesic. This is fairly standard but we provide details for completeness. Let be points such that the interior angle between and is at least . The manifold is a space, since the sectional curvature is bounded above by . Denote by the hyperbolic metric in . Let be points such that , , and such that the angle at of the geodesic triangle determined by is equal to . Then (see [BH, Proposition 1.7, Chapter 2]). Moreover, using the hyperbolic law of cosine and that , we get
Observe that the last term is at least , therefore
where the last inequality holds provided that , for large enough. Set . Therefore,
(2) |
Let us now consider a piecewise geodesic curve with lengths and angles , where and , for every . We claim that is a -local-quasi-geodesic. Note that if and
then either , or there exists such that In the first case the condition in Definition 3.2 is clearly satisfied; in the later case we can use inequality (2) to conclude the claim.
Following [CDP, Theorem 1.4, Chapter 3] a -local-quasi-geodesic is a -quasi-geodesic, where depends only on the constants . Here denotes a positive constant for which is a -hyperbolic space in the sense of Gromov. In [CDP, Theorem 3.1, Chapter 3] it is proved the stability of quasi-geodesics having infinite length. More precisely, they proved that a -quasi-geodesic is contained in a -neighbourhood of a geodesic line, where depends only on and . Combining these results and the discussion above we conclude the following result:
Theorem 3.3.
Given there exists such that any piecewise geodesic curve with lengths and angles , where and , for every , is contained in a -neighbourhood of a geodesic line.
The following proposition summarizes relevant properties of piecewise geodesic curves having interior angles bounded from below away from 0. This proposition will be useful at the end of Section 4.
Proposition 3.4.
Given there exist positive constants such that the following holds: if is a piecewise geodesic curve with and for every , then the interior angle at between the geodesic segments and is larger than , and
(3) |
for every .
Proof.
To begin, let us make some general observations. Suppose we have a geodesic triangle in with vertices , whose angle at vertex is greater than . Using the hyperbolic law of cosines and a comparison theorem, we can derive that there exists a constant such that
Furthermore, if (where is large enough), then the interior angle at vertex is at most . We will assume Next, consider a point such that the interior angle at vertex of is greater than and . It follows that the interior angle at of is at least and we have that
Note that the triangle satisfies similar conditions to , so we can iterate this procedure.
By induction, it follows that the interior angle between and is at most . Furthermore, we have
Since , we conclude that the interior angle at between the geodesic segments and is greater than .
∎
4. Proof of Theorem 1.2
In this section we will prove that the Hausdorff dimension of the diverging on average radial limit set is equal to We separate the proof into two parts: we first prove that is an upper bound for the Hausdorff dimension and then we prove it is a lower bound.
Fix a reference point . Let be an increasing sequence of compact pathwise connected sets with piecewise boundaries such that . Let be a nice pre-image of containing . We assume that , for every , and that . For , let be the unit vector based at pointing towards . Denote by the canonical projection. Observe that
4.1. The upper bound
Let and large enough such that , for every . Fix . To ease the notation we write , and set for the -neighbourhood of . Set and for , let be the arc-length parametrization of the geodesic segment . Define
and
Define as the set of diverging on average radial limit points such that returns infinitely many times to . For any set such that , where is the constant from Lemma 2.4 for . Set . It is straightforward to prove that the set of balls is a covering of by balls of diameter less than . Hence
We will prove that for small enough, we have
(4) |
which implies that as (in this case ) whenever . This implies that
for every . Finally, observe that , and therefore . Since we obtain the desired bound.
In order to establish (4), we will make use of an estimate derived from [GST, Section 5]. This estimate concerns the quantity of periodic orbits that spend a significant portion of their time outside of an -neighborhood of a specified compact set in . Their result in the context of counting closed geodesics, necessitating the consideration of multiplicities and accounting for the number of distinct lifts intersecting a nice preimage. Within their proof, they evaluate the cardinality of .
Proposition 4.1.
Let and . Then, for every and , there exists a positive number such that
(5) |
Moreover, for fixed, tends monotonically to 0 when tends to 0.
4.2. The bound from below
To establish the lower bound of the Hausdorff dimension, we will construct a subset and a positive measure on such that Frostman’s lemma applies with the appropriate exponent. Fix and set .
For every , let be the arc-length parametrization of the geodesic ray with . The closure of the set of points in that projects orthogonally into is denoted by . Set . In other words, the set is the convex-hull of on . The following corresponds to Lemma 2.5 in [Sc]. Let be the hyperbolicity constant asociated to .
Lemma 4.2.
For every , we have
We consider an increasing sequence of compact pathwise connected sets with piecewise boundaries such that , and is a nice preimage of containing . We further assume that , for every . Since and is compact, there exists such that for every and every we have
Since is non-elementary, there exists such that . Set and , where we choose large enough such that the closures of and are disjoints and are not in . There exists such that for every , the angle between the geodesic segments and at is bounded from below by . Let be the constant obtained from Theorem 3.3, and fix where is the constant from Lemma 2.4 associated to . Set and . The equation above and the triangle inequality imply that
(6) |
since contains .
Define . We claim that for every and we have
(7) |
Indeed, if this is not the case there would exist , and such that
for every Then
which contradicts (6). It follows from (7) that
(8) |
for all and .
The next result can be deduced from classical geometric arguments.
Lemma 4.3.
Given there exists such that for every , if verify , then and are disjoint.
Choose a sequence such that
(9) |
where is the diameter of . This can be done since can be chosen arbitrarily large once is given.
Since the set is finite, the ball of radius centered at contains at most points of . This implies that any set has a subset of size at least , where each pair of points has distance at least . In particular, by Lemma 4.3 and by equation (8), for each , and , there is and a finite subset of the set such that
-
(i)
for any , , the balls
are disjoint, and
-
(ii)
.
Now we describe the inductive construction of radial limit points defining geodesic rays escaping in average to infinity.
First step of induction. We consider a geodesic segment , where . Recall that , for some . There exists such that the geodesic segment forms an angle with the geodesic segment , for every . In particular, the piecewise geodesic path has interior angle at larger than . Note that , where , so the piecewise geodesic curve has the form
Second step of induction. We consider a piecewise geodesic curve of the form
(10) |
with interior angles larger than , where and , for every and . Note that is a piecewise geodesic curve ending at , so there exists such that and forms a angle larger than at , for every . Since , for some , a piecewise geodesic curve defined as in (10) has all interior angles larger than .
This inductive construction defines a tree in for which each vertex at the -step has at least children , for . By Theorem 3.3 each piecewise geodesic curves of is contained in some -neighbourhood of a geodesic ray. By (i) these geodesic rays are different. Let us denote by the set of extremities at infinity of them. We also denote the set of children of a vertex and the set of vertices at level . Note that .
Lemma 4.4.
The set is contained in .
Proof.
Let . By construction a -neighbourhood of the geodesic ray contains a unique piecewise geodesic curve in passing through infinitely many orbit elements , so by definition is a radial point.
To prove that escapes in average we will use Condition (9). Let be any compact set in containing a -neighbourhood of and denote by its diameter. Then for any large enough we have . We will assume without loss of generality that for any . Let be the arc-length parametrization of . For , the interval can be decomposed into the union
where is an increasing sequence of positive real numbers, and . Each time is defined to be the time for which is the closest point of from . Recall that , where , . By construction , in particular . With this information, we have
so we need to estimate now the number of times intersects . A priori, the number of times the geodesic segment intersects translations of might be large. We will prove this number to be bounded from above by a constant depending on and , not on . We will assume in addition that , with , which is the hardest situation since in the case the segment intersects translations of only at the beginning and at the end. Let and assume that a -neighbourhood of is contained in . Since the curvature of is negative, there exists such that for every with , then the geodesic segments and are -close for every and . We apply this to , and .
Suppose that intersects not only at the beginning and at the end. If , then a -neighbourhood of intersects . Since contains a -neighbourhood of , then also intersects . The later cannot happen since intersects a -translation of only at the beginning and at the end, so intersects only if . In particular, given , the number -translations of intersecting is bounded from above by a constant depending on , and . Hence
On the other hand, by construction, we have
so
Since as , Condition (9) implies that the limit above when is 0. Hence defines an escaping in average direction. ∎
For define and . Note that if are different, then Define , and observe that . By (i), we have
Moreover, if , then
(11) |
In order to prove (11) first note that is in the shadow at infinity from of the ball since both and are at distance at most of a geodesic curve with extremity at infinity in . In particular . By triangle inequality, if , then
Following the proof of Proposition 3.4, we know that , where is a constant depending only on (and the bounds of the sectional curvatures). Since we can choose and to be arbitrarily large, we will assume . Hence
Finally, we get , which ends the proof of equation (11).
We are now able to define a measure on by setting and
whenever .
It follows by induction and (ii) that , for every Let and the smallest natural number such that there is where intersects but it is not contained in . Let be the parent of . Since we have that
(12) |
Set and . Note that , and therefore . Note also that . By -hyperbolicity, we get
In conclusion, we obtain , and therefore
(13) |
We choose such that
where is chosen according to (3). This can be done as consequence of the following technical lemma.
Lemma 4.5.
Let be an increasing sequence of real numbers converging to . Then, there exists a sequence of positive integers such that for and , for , if
then
Proof.
Let assume for the moment that is any sequence of positive real numbers. Then, for , we have
So the numbers need to be chosen so that
∎
Remark 4.6.
The construction of and the numbers depend on the compact sets . Hence, according to the previous lemma, the suitable sequence can be obtained by making the sequence constant in some large enough intervals. In terms of the dynamics, we are constructing geodesic orbits escaping slowly in average. This does not change our key assumption (9).
We apply Lemma 4.5 to . We therefore get for large enough
Note that each is smaller than the length of a geodesic segment ending in , so
from (3). Using (12) and the fact that , this already implies that
By (13), we finally obtain
for some constant , which give us the lower bound through Frostman’s Lemma. Since and is arbitrary, we obtain that
5. Entropy of infinite measures
In this section we relate the diverging on average radial limit set and the entropy at infinity with the entropy of -finite, ergodic and conservative infinite measures.
A geodesic flow invariant Borel measure is said to be conservative if it is supported on the non-wandering set of the geodesic flow, and ergodic if every flow invariant set have either full or zero measure. We denote by the space of geodesic flow invariant -finite Borel measures on that are infinite, ergodic and conservative. Let be the set of diverging on average vectors in the non-wandering set.
Lemma 5.1.
If , then .
Proof.
It is a consequence of Hopf’s ergodic theorem that for every and -a.e. , we have that
In particular, this applies to , where is a compact set. We conclude that, on average, a generic point of spends zero proportion of time within any compact set. ∎
Following [Led] we now recall a notion of measure-theoretic entropy that allows us to include infinite measures. Recall that for , and , a -dynamical ball is the set
where denotes the Sasaki metric on .
Definition 5.2.
Let be a geodesic flow invariant -finite Borel measure on . The measure-theoretic entropy of is defined as
It worth noticing that Lipschitz equivalent metrics on lead to the same entropy value. Moreover, it is a consequence of Brin-Katok entropy formula that if is an ergodic probability measure, then coincides with the entropy of defined using partitions.
For a vector , we denote by the end point at infinity of the geodesic ray . Let be the map defined by . The proof of our next result follows closely the strategy considered by F. Ledrappier in [Led, Proposition 4.4].
Theorem 5.3.
If , then .
Proof.
Let be a point in the support of and smaller than the injectivity radius at . Let be a lift of to . The restriction of to can be lifted to a measure on . Set . Observe that since is conservative, the support of is a subset of . Moreover, it follows from Lemma 5.1 that the support of is a subset of . We claim that
(14) |
Inequality (14), together with Theorem 1.2 and Theorem 2.15, finishes the proof. Note that there exists such that for every , the image of under contains , where is the constant given by Lemma 2.4. Then,
and therefore (see Proposition 2.14). ∎
References
- [AM] P. Apisa and H. Masur, Divergent on average directions of Teichmüler geodesic flow, J. Eur. Math. Soc. (JEMS) 24 no. 3 (2022), pp 1007–1044.
- [BJ] C. Bishop and P. Jones, Hausdorff dimension and Kleinian groups, Acta. Math., 179 (1997), pp 1–39.
- [Bo] B. Bowditch, Geometrical finiteness with variable negative curvature, Duke Math. J., 77 (1995), pp 229–274.
- [BH] M. Bridson and A. Haefliger, Metric Spaces of Non-Positive Curvature, Springer (1999).
- [Bu] J. Buzzi, Puzzles of quasi-finite type, zeta functions and symbolic dynamics for multidimensional maps. Annales de l’Institut Fourier 60 (2010), pp 801–852.
- [Ch] Y. Cheung, Hausdorff dimension of the set of singular pairs, Ann. of Math. (2) 173 no. 1 (2011), pp 127–167.
- [CDP] M. Coornaert, T. Delzant and A. Papadopoulos, Géométrie et théorie des groupes, Lecture Notes in Mathematics, Springer-Verlag, 1441 (1990).
- [D] S. Dani, Divergent trajectories of flows on homogeneous spaces and Diophantine approximation. J. Reine Angew. Math. 359 (1985), pp 55–89.
- [DFSU] T. Das, L. Fishman, D. Simmons, M. Urbański, A variational principle in the parametric geometry of numbers, preprint.
- [EK] M. Einsiedler, S. Kadyrov, Entropy and escape of mass for . Israel J. Math. 190 (2012), pp 253–288.
- [EKP] M. Einsiedler, S. Kadyrov, A. Pohl, Escape of mass and entropy for diagonal flows in real rank one situations, Israel J. Math. 210 no. 1 (2015), pp 245–295.
- [GST] S. Gouezel, B. Schapira and S. Tapie, Pressure at infinity and strong positive recurrence in negative curvature, Preprint ArXiv:2007.08816.
- [GZ] B. Gurevich, A. Zargaryan, Conditions for the existence of a maximal measure for a countable symbolic Markov chain. Vestnik Moskov. Univ. Ser. I Mat. Mekh. 1988, no. 5, 14–18, 103; translation in Moscow Univ. Math. Bull. 43 no. 5 (1988), pp 18–23.
- [IRV] G. Iommi, F. Riquelme and A. Velozo Entropy in the cusp and phase transitions for geodesic flows, Israel J. Math., 225 (2) (2018), pp 609–659.
- [ITV] G. Iommi, M. Todd and A. Velozo Escape of entropy for countable Markov shifts, Adv. Math., 405 (2022), Paper No. 108507, 54pp.
- [KKLM] S. Kadyrov and D. Kleinbock and E. Lindenstrauss and G. A. Margulis, Singular systems of linear forms and non-escape of mass in the space of lattices, J. Anal. Math., 133 (2017), pp 253–277.
- [KP] S. Kadyrov, A. Pohl, Amount of failure of upper-semicontinuity of entropy in non-compact rank-one situations, and Hausdorff dimension. Ergodic Theory Dynam. Systems 37 no. 2 (2017), pp 539–563.
- [Ka] V. Kaimanovich, Invariant measures of the geodesic flow and measures at infinity on negatively curved manifolds, Ann. Inst. H. Poincaré Phys. Théor., 53 no 4 (1990), pp 361–393.
- [Led] F. Ledrappier, Entropie et principe variationnel pour le flot géodésique en courbure négative pincée, Monogr. Enseign. Math., 43 (2013), pp 117–144.
- [Ma] H. Masur, Hausdorff dimension of the set of nonergodic foliations of a quadratic differential, Duke Math. J., 66 no 3 (1992), pp 387–442.
- [OP] J. P. Otal and M. Peigné, Principe variationnel et groupes Kleiniens, Duke Math. J., 125 no 1 (2004), pp 15–44.
- [Pau] F. Paulin, On the critical exponent of a discrete group of hyperbolic isometries, Differential Geom. Appl., 7 no 3 (1997), pp 231–236.
- [PPS] F. Paulin, M. Pollicot, and B. Schapira, Equilibrium states in negative curvature, Astérisque, no 373 (2015).
- [RV] F. Riquelme and A. Velozo, Escape of mass and entropy for geodesic flows, Erg. Theo. Dyna. Syst 39 no 2 (2019), pp 446–473.
- [Ru] S. Ruette, On the Vere-Jones classification and existence of maximal measures for countable topological Markov chains. Pacific J. Math. 209 no. 2 (2003), pp 366–380.
- [Sc] B. Schapira, Lemme de l’ombre et non divergence des horosphères d’une variété géométriquement finie, Ann. Inst. Fourier (Grenoble) 54 no. 4 (2004), pp 939–987.
- [ST] B. Schapira and S. Tapie,Regularity of entropy, geodesic currents and entropy at infinity. Ann. Sci. Éc. Norm. Supér. (4) 54 no. 1 (2021), pp 1–68.
- [S] D. Sullivan, Entropy, Hausdorff measures old and new, and limit sets of geometrically finite Kleinian groups, Acta Math. 153 (1984), pp 259–277.
- [Ve] A. Velozo Thermodynamic formalism and the entropy at infinity of the geodesic flow, Preprint arXiv:1711.06796.