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

On the Hausdorff dimension of geodesics that diverge on average

Felipe Riquelme IMA, Pontificia Universidad Católica de Valparaíso, Blanco Viel 596, Valparaíso, Chile. felipe.riquelme@pucv.cl http://ima.ucv.cl/academicos/felipe-riquelme/  and  Anibal Velozo Facultad de Matemáticas, Pontificia Universidad Católica de Chile (PUC), Avenida Vicuña Mackenna 4860, Santiago, Chile apvelozo@mat.uc.cl
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 σ\sigma-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.
F.R. was supported by FONDECYT Iniciación 11190461 and Regular 1231257
A.V. was supported by FONDECYT Iniciación 11220409.

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 qq on a genus gg surface with nn punctures, the Hausdorff dimension of the set of angles θ\theta, for which eiθqe^{i\theta}q diverges on average for the Teichmüller geodesic flow on the corresponding Teichmuller space is bounded from above by 1/21/2. This is also motivated by the fact that if the measured foliation associated to eiθqe^{i\theta}q 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 eiθqe^{i\theta}q diverges on average is in fact equal to 1/21/2. 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(et,et,e2te^{t},e^{t},e^{-2t}) on SL(3,)/SL(3,)\text{SL}(3,{\mathbb{R}})/\text{SL}(3,{\mathbb{Z}}). 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 gt=diag(ent,,ent,emt,,emt)g_{t}=\text{diag}(e^{nt},\ldots,e^{nt},e^{-mt},\ldots,e^{-mt}) on SL(m+n,)/SL(m+n,)\text{SL}(m+n,{\mathbb{R}})/\text{SL}(m+n,{\mathbb{Z}}) 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 ~\widetilde{\mathcal{M}} be a complete, simply connected, pinched negatively curved manifold, and let Γ\Gamma be a discrete, torsion free, non-elementary group of isometries of ~\widetilde{\mathcal{M}}. Consider the quotient manifold =~/Γ\mathcal{M}=\widetilde{\mathcal{M}}/\Gamma and let p:T1~T1p:T^{1}\widetilde{\mathcal{M}}\to T^{1}\mathcal{M} be the projection map between unit tangent spaces. The geodesic flow on \mathcal{M} is denoted by (gt)t(g_{t})_{t\in{\mathbb{R}}}, where gt:T1T1g_{t}:T^{1}\mathcal{M}\to T^{1}\mathcal{M} is the time tt-map. Let χA\chi_{A} be the characteristic function of a subset AA.

Definition 1.1.

Let vT1v\in T^{1}\mathcal{M}. We say that vv diverges on average if for any compact set WT1W\subset T^{1}\mathcal{M}, we have

limT+1T0TχW(gtv)𝑑t=0.\lim_{T\to+\infty}\frac{1}{T}\int_{0}^{T}\chi_{W}(g_{t}v)dt=0.

We say that vv is divergent if for any compact set WT1W\subset T^{1}\mathcal{M} there exists T>0T>0 such that χW(gt(v))=0\chi_{W}(g_{t}(v))=0, for any tTt\geqslant T. A vector is recurrent if it is not divergent.

A vector in T1~T^{1}\widetilde{\mathcal{M}} is said to be divergent on average, divergent or recurrent depending on whether its projection to T1T^{1}\mathcal{M} under the map pp has the corresponding property. Given xx in \mathcal{M} or ~\widetilde{\mathcal{M}}, we use (x)\mathcal{R}(x) to denote the set of recurrent vectors based at xx, and 𝒟𝒜(x)\mathcal{DA}(x) to denote the set of vectors that diverge on average based at xx. We define

𝒟𝒜(x)=(x)𝒟𝒜(x).\mathcal{RDA}(x)=\mathcal{R}(x)\cap\mathcal{DA}(x).

Fix a reference point z~z\in\widetilde{\mathcal{M}} and denote by Θz:Tz1~~,\Theta_{z}:T^{1}_{z}\widetilde{\mathcal{M}}\to\partial_{\infty}\widetilde{\mathcal{M}}, the canonical identification that sends vTz1~v\in T^{1}_{z}\widetilde{\mathcal{M}} to the geodesic ray that starts at zz with direction vv. Observe that ΛΓrad=Θz((z))\Lambda_{\Gamma}^{rad}=\Theta_{z}(\mathcal{R}(z)), where ΛΓrad\Lambda_{\Gamma}^{rad} is the radial limit set of Γ\Gamma. Define the diverging on average radial limit set of Γ\Gamma by

ΛΓ,rad=Θz(𝒟𝒜(z)).\Lambda_{\Gamma}^{\infty,rad}=\Theta_{z}(\mathcal{RDA}(z)).

Since two asymptotic rays approach exponentially fast, the sets Θz((z))\Theta_{z}(\mathcal{R}(z)) and Θz(𝒟𝒜(z))\Theta_{z}(\mathcal{DA}(z)) are independent of the reference point zz. The entropy at infinity of the geodesic flow on \mathcal{M} is denoted by δΓ\delta_{\Gamma}^{\infty}. The main result of this article is the following.

Theorem 1.2.

Let \mathcal{M} 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 δΓ\delta^{\Gamma}_{\infty}, that is,

HD(ΛΓ,rad)=δΓ,\emph{HD}(\Lambda_{\Gamma}^{\infty,rad})=\delta_{\Gamma}^{\infty},

where the Hausdorff dimension is computed using the Gromov-Bourdon visual metric on ~\partial_{\infty}\widetilde{\mathcal{M}}.

It worth noticing that if =n/Γ\mathcal{M}={\mathbb{H}}^{n}/\Gamma is hyperbolic, then the map Θz\Theta_{z} is bi-Lipschitz considering Tz1nT^{1}_{z}{\mathbb{H}}^{n} endowed with the hyperbolic metric. In particular, Theorem 1.2 can be stated as

HD(𝒟𝒜(x))=δΓ,\textrm{HD}(\mathcal{RDA}(x))=\delta_{\Gamma}^{\infty},

where the Hausdorff dimension is computed with respect to the hyperbolic metric and xx\in\mathcal{M} is any point.

In the case that \mathcal{M} is geometrically finite with parabolic, then it is known that δΓ=max{δ𝒫:𝒫Γ parabolic}\delta^{\infty}_{\Gamma}=\max\{\delta_{\mathcal{P}}:{\mathcal{P}}\subset\Gamma\textrm{ parabolic}\}, 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 \mathcal{M} is non-compact, finite volume and hyperbolic, then HD(ΛΓ,rad)=n12\text{HD}(\Lambda_{\Gamma}^{\infty,rad})=\frac{n-1}{2}, where n=dimn=\dim\mathcal{M}. More generally, if \mathcal{M} is hyperbolic and geometrically finite with cusps, then the Hausdorff dimension is half the maximal rank of parabolic subgroups of Γ\Gamma.

Following Ledrappier [Led] we consider the entropy of infinite, σ\sigma-finite, ergodic and conservative measures. We denote by h(m)h(m) the entropy of the measure mm. As a consequence of Theorem 1.2 we obtain the following result.

Theorem 1.3.

Let mm be an infinite measure on T1T^{1}\mathcal{M} which is σ\sigma-finite, invariant by the geodesic flow, ergodic and conservative. Then h(m)δΓh(m)\leq\delta^{\Gamma}_{\infty}.

Let M(T1,g)M_{\infty}(T^{1}\mathcal{M},g) be the space of geodesic flow invariant σ\sigma-finite Borel measures on T1T^{1}\mathcal{M} that are infinite, ergodic and conservative. Theorem 1.3 can be stated as

supmM(T1,g)h(m)δΓ.\sup_{m\in M_{\infty}(T^{1}\mathcal{M},g)}h(m)\leqslant\delta_{\infty}^{\Gamma}.

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 ~\widetilde{\mathcal{M}} of dimension nn. We assume that the sectional curvatures of the metric lie in the interval [b2,1][-b^{2},-1], for some b1b\geqslant 1. The group of isometries of ~\widetilde{\mathcal{M}} is denoted by Iso(~)\text{Iso}(\widetilde{\mathcal{M}}) and consists of the set of diffeomorphisms that preserve the Riemannian metric. The Riemannian distance of ~\widetilde{\mathcal{M}} will be denoted by dd.

2.1. Boundary at infinity

The boundary at infinity, or Gromov boundary of ~\widetilde{\mathcal{M}}, which we denote by ~\partial_{\infty}\widetilde{\mathcal{M}}, is the set of equivalent classes of asymptotic geodesic rays on ~\widetilde{\mathcal{M}}. We endow ~\partial_{\infty}\widetilde{\mathcal{M}} with the cone topology, which makes it homeomorphic to the unit sphere in n{\mathbb{R}}^{n}. More concretely, consider the map Θz:Tz1~~\Theta_{z}:T^{1}_{z}\widetilde{\mathcal{M}}\to\partial_{\infty}\widetilde{\mathcal{M}} that sends vTz1~v\in T^{1}_{z}\widetilde{\mathcal{M}} to the class of the geodesic ray based at zz with initial direction vv. The map Θz\Theta_{z} defines a homeomorphism for every z~z\in\widetilde{\mathcal{M}}.

For x,x~x,x^{\prime}\in\widetilde{\mathcal{M}} and ξ,ξ~\xi,\xi^{\prime}\in\partial_{\infty}\widetilde{\mathcal{M}} we will denote by

  • [x,x][x,x^{\prime}] the geodesic segment starting at xx and ending at xx^{\prime}.

  • [x,ξ)[x,\xi) the geodesic ray starting at xx with end point at infinity ξ\xi.

  • (ξ,ξ)(\xi,\xi^{\prime}) the geodesic line having ξ\xi and ξ\xi^{\prime} as end points at infinity.

Given x,y~x,y\in\widetilde{\mathcal{M}} and r>0r>0 consider the sets

Ux,y,r={z~~:[x,z)B(y,r)}.U_{x,y,r}=\{z\in\widetilde{\mathcal{M}}\cup\partial_{\infty}\widetilde{\mathcal{M}}:[x,z)\cap B(y,r)\neq\emptyset\}.

Topologize the compactification ¯=~~\overline{\mathcal{M}}=\widetilde{\mathcal{M}}\cup\partial_{\infty}\widetilde{\mathcal{M}} such that sets of the form Ux,y,rU_{x,y,r} are a basis of neighborhoods of boundary points. The compactification ¯\overline{\mathcal{M}} is homeomorphic to the closed unit ball in n{\mathbb{R}}^{n}.

In this article we are interested in calculating the Hausdorff dimension of subsets of ~\partial_{\infty}\widetilde{\mathcal{M}} with respect to certain metrics that we proceed to define.

Definition 2.1.

The Busemann cocycle is the map β:~×~×~,\beta:\partial_{\infty}\widetilde{\mathcal{M}}\times\widetilde{\mathcal{M}}\times\widetilde{\mathcal{M}}\to{\mathbb{R}}, defined by

β:(ξ,x,y)βξ(x,y):=limt+d(x,ξt)d(y,ξt),\beta:(\xi,x,y)\mapsto\beta_{\xi}(x,y):=\lim_{t\to+\infty}d(x,\xi_{t})-d(y,\xi_{t}),

where tξtt\mapsto\xi_{t} is the parametrization of a geodesic ray ending at ξ\xi.

Definition 2.2.

Let x~x\in\widetilde{\mathcal{M}}. The Gromov product at xx between z,w~z,w\in\widetilde{\mathcal{M}} is defined as

z,wx=12[d(z,x)+d(x,w)d(z,w)]\langle z,w\rangle_{x}=\frac{1}{2}[d(z,x)+d(x,w)-d(z,w)]

The Gromov product at xx between ξ,η~\xi,\eta\in\partial_{\infty}\widetilde{\mathcal{M}} is defined as

ξ,ηx=limzξwηz,wx.\langle\xi,\eta\rangle_{x}=\lim_{\begin{subarray}{c}z\to\xi\\ w\to\eta\end{subarray}}\langle z,w\rangle_{x}.
Definition 2.3.

Let x~x\in\widetilde{\mathcal{M}}. The Gromov-Bourdon visual distance dxd_{x} on ~\partial_{\infty}\widetilde{\mathcal{M}} is defined as

dx(ξ,η)={eξ,ηx if ξη0 if ξ=η.d_{x}(\xi,\eta)=\begin{cases}e^{-\langle\xi,\eta\rangle_{x}}&\emph{ if }\xi\neq\eta\\ 0&\emph{ if }\xi=\eta.\end{cases}

Observe that ξ,ηx=12(βξ(x,y)+βη(x,y))\langle\xi,\eta\rangle_{x}=-\frac{1}{2}(\beta_{\xi}(x,y)+\beta_{\eta}(x,y)), for any y(ξ,η)y\in(\xi,\eta). Every Gromov-Bourdon visual distance induces the cone topology on ~\partial_{\infty}\widetilde{\mathcal{M}}. Moreover, they are mutually conformal and Iso(~)\text{Iso}(\widetilde{\mathcal{M}})-invariant. More precisely,

(1) dx(ξ,η)=exp(12[βξ(x,x)+βη(x,x)])dx(ξ,η)d_{x^{\prime}}(\xi,\eta)=\exp\left(\frac{1}{2}\left[\beta_{\xi}(x,x^{\prime})+\beta_{\eta}(x,x^{\prime})\right]\right)d_{x}(\xi,\eta)

and

dγx(γξ,γη)=dx(ξ,η).d_{\gamma x}(\gamma\xi,\gamma\eta)=d_{x}(\xi,\eta).

Given x,x~x,x^{\prime}\in\widetilde{\mathcal{M}} and r>0r>0, we define the shadow at infinity from xx of the ball B(x,r)B(x^{\prime},r) as the set

𝒪x(x,r)={η~:[x,η)B(x,r)}.\mathcal{O}_{x}(x^{\prime},r)=\{\eta\in\partial_{\infty}\widetilde{\mathcal{M}}:[x,\eta)\cap B(x^{\prime},r)\neq\emptyset\}.

Denote by

Bx(ξ,r)={η~:dx(η,ξ)<r},B_{x}(\xi,r)=\{\eta\in\partial_{\infty}\widetilde{\mathcal{M}}:d_{x}(\eta,\xi)<r\},

where ξ~\xi\in\partial_{\infty}\widetilde{\mathcal{M}} and r>0r>0. For every x,x~x,x^{\prime}\in\widetilde{\mathcal{M}} we denote by ξx,x\xi_{x,x^{\prime}} to the end point of the geodesic ray based at xx passing through xx^{\prime}. The following lemma is due to Kaimanovich (see [Ka, Proposition 1.4]).

Lemma 2.4.

For any r>0r>0 there exists c0(r)1c_{0}(r)\geq 1 such that for any x,x~x,x^{\prime}\in\widetilde{\mathcal{M}}, we have

Bx(ξx,x,c0(r)1et)𝒪x(x,r)Bx(ξx,x,c0(r)et),B_{x}(\xi_{x,x^{\prime}},c_{0}(r)^{-1}e^{-t})\subseteq\mathcal{O}_{x}(x^{\prime},r)\subseteq B_{x}(\xi_{x,x^{\prime}},c_{0}(r)e^{-t}),

where t=d(x,x)t=d(x,x^{\prime}).

2.2. The geodesic flow

Let Γ\Gamma be a discrete, torsion free subgroup of isometries of ~\widetilde{\mathcal{M}} and consider the quotient manifold =~/Γ\mathcal{M}=\widetilde{\mathcal{M}}/\Gamma. The quotient map is denoted by pΓ:~p_{\Gamma}:\widetilde{\mathcal{M}}\to\mathcal{M}. Let o~o\in\widetilde{\mathcal{M}} be a reference point.

Elements of Iso(~)\text{Iso}(\widetilde{\mathcal{M}}) are classified into three types: hyperbolic, parabolic, or elliptic. We say that Γ\Gamma is non-elementary if it is neither generated by a hyperbolic element nor a parabolic subgroup. We say that \mathcal{M} is non-elementary if Γ\Gamma is non-elementary.

The geodesic flow on \mathcal{M} is denoted by (gt)t(g_{t})_{t\in{\mathbb{R}}}, where gt:T1T1g_{t}:T^{1}\mathcal{M}\to T^{1}\mathcal{M} is the time tt-map. A point xT1x\in T^{1}\mathcal{M} is said to be non-wandering if for any neighborhood UU of xx and any N>0N>0, there exists t>Nt>N such that gt(U)Ug_{-t}(U)\cap U\neq\emptyset. The set of all non-wandering points is called the non-wandering set of the geodesic flow. The non-wandering set is denoted by Ω\Omega. This set is closed and invariant under the geodesic flow, meaning that if xΩx\in\Omega, then gt(x)Ωg_{t}(x)\in\Omega for all tt\in{\mathbb{R}}. Furthermore, the Poincaré recurrence theorem implies that if μ\mu is a geodesic flow invariant probability measure, then μ(Ω)=1\mu(\Omega)=1. In particular, the support of any invariant probability measure is contained in Ω\Omega.

We say \mathcal{M} is convex cocompact if Ω\Omega is compact. We say that \mathcal{M} is geometrically finite if an ϵ\epsilon-neighborhood of Ω\Omega has finite Liouville volume (for a detailed discussion on the notion of geometrically finiteness, see [Bo]).

Let M(T1,g)M(T^{1}\mathcal{M},g) be the space of geodesic flow invariant probability measures on T1T^{1}\mathcal{M}, and h(μ)h(\mu) the measure-theoretic entropy of μ(g)\mu\in\mathcal{M}(g). The topological entropy of the geodesic flow is defined as

htop(g)=supμM(T1,g)h(μ).h_{top}(g)=\sup_{\mu\in M(T^{1}\mathcal{M},g)}h(\mu).

2.3. Limit sets

The limit set of Γ\Gamma, denoted by ΛΓ\Lambda_{\Gamma}, is the set of accumulation points of a Γ\Gamma-orbit on ~\partial_{\infty}\widetilde{\mathcal{M}}. More precisely,

ΛΓ=Γo¯Γo.\Lambda_{\Gamma}=\overline{\Gamma\cdot o}\setminus\Gamma\cdot o.

The limit set ΛΓ\Lambda_{\Gamma} is independent of the reference point o~o\in\widetilde{\mathcal{M}}. Observe that ξΛΓ\xi\in\Lambda_{\Gamma} if there exists a sequence (γn)n(\gamma_{n})_{n} in Γ\Gamma and a point x~x\in\widetilde{\mathcal{M}} such that γnx\gamma_{n}x converges to ξ\xi.

A point ξΛΓ\xi\in\Lambda_{\Gamma} is called a radial limit point if there exists an RR-neighborhood of the geodesic ray [o,ξ)[o,\xi) that contains infinitely many elements of Γo\Gamma\cdot o for some R>0R>0. The radial limit set of Γ\Gamma, denoted by ΛΓrad\Lambda_{\Gamma}^{rad}, is the collection of all radial limit points of Γ\Gamma. It is worth noting that ξΛΓrad\xi\in\Lambda_{\Gamma}^{rad} if the projection to \mathcal{M} of the geodesic ray [o,ξ)[o,\xi) returns infinitely many times to the ball centered at pΓ(o)p_{\Gamma}(o) of radius RR, for some R>0R>0. In other words, ΛΓrad=Θz((z))\Lambda_{\Gamma}^{rad}=\Theta_{z}(\mathcal{R}(z)), where Θz:Tz1~~\Theta_{z}:T^{1}_{z}\widetilde{\mathcal{M}}\to\partial_{\infty}\widetilde{\mathcal{M}} is the canonical identification that sends vTz1~v\in T^{1}_{z}\widetilde{\mathcal{M}} to the geodesic ray that starts at zz with direction vv, and (z)Tz1~\mathcal{R}(z)\subset T^{1}_{z}\widetilde{\mathcal{M}} is the set of recurrent directions based at zz.

In the introduction of this article, we introduced a subset of the radial limit set called the diverging on average radial limit set of Γ\Gamma. This is defined as

ΛΓ,rad=Θz(𝒟𝒜(z)),\Lambda_{\Gamma}^{\infty,rad}=\Theta_{z}(\mathcal{RDA}(z)),

where 𝒟𝒜(z)\mathcal{RDA}(z) is the set of recurrent and diverging on average vectors based at z~z\in\widetilde{\mathcal{M}}. Equivalently, ξΛΓ,rad\xi\in\Lambda_{\Gamma}^{\infty,rad} if it is a radial limit point and

limT1T0TχpΓ1(K)(gtvξ)𝑑t=0,\lim_{T\to\infty}\frac{1}{T}\int_{0}^{T}\chi_{p_{\Gamma}^{-1}(K)}(g_{t}v_{\xi})dt=0,

for every compact set KK\subset\mathcal{M}, where vξT1~v_{\xi}\in T^{1}\widetilde{\mathcal{M}} is a vector whose associated geodesic ray converges to ξ\xi.

2.4. Critical exponents

The critical exponent of Γ\Gamma is defined by

δΓ=lim supR+log(#{γΓ:d(o,γo)R})R.\delta_{\Gamma}=\limsup_{R\to+\infty}\frac{\log(\#\{\gamma\in\Gamma:d(o,\gamma\cdot o)\leq R\})}{R}.

This number does not depend on the reference point o~o\in\widetilde{\mathcal{M}}. Moreover, the critical exponent of Γ\Gamma is the exponent of convergence of the Poincaré series

PΓ(s)=γΓesd(o,γo).P_{\Gamma}(s)=\sum_{\gamma\in\Gamma}e^{-sd(o,\gamma\cdot o)}.

Indeed, the Poincaré series converges for s>δΓs>\delta_{\Gamma} and diverges for s<δΓs<\delta_{\Gamma}.

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 3{\mathbb{H}}^{3}. 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 =~/Γ\mathcal{M}=\widetilde{\mathcal{M}}/\Gamma be a non-elementary pinched negatively curved Riemannian manifold. Then

HD(ΛΓrad)=δΓ,\emph{HD}(\Lambda^{rad}_{\Gamma})=\delta_{\Gamma},

where the Hausdorff dimension is computed using a Gromov-Bourdon visual metric on ~\partial_{\infty}\widetilde{\mathcal{M}}.

Theorem 2.6.

Let =~/Γ\mathcal{M}=\widetilde{\mathcal{M}}/\Gamma be a non-elementary pinched negatively curved Riemannian manifold. Then htop(g)=δΓh_{top}(g)=\delta_{\Gamma}.

In analogy to the critical exponent of Γ\Gamma we now define the critical exponent outside a compact set of \mathcal{M}, which quantifies the complexity of the ends of the manifold. Let KK\subseteq\mathcal{M} be a compact, pathwise-connected set that is the closure of its interior and has a piecewise C1C^{1} boundary. A nice preimage of KK is a compact set K~~\widetilde{K}\subseteq\widetilde{\mathcal{M}} with a piecewise C1C^{1} boundary such that pΓ(K~)=Kp_{\Gamma}(\widetilde{K})=K and the restriction of pΓp_{\Gamma} to the interior of K~\widetilde{K} is injective. The following lemma was proved in [ST, Lemma 7.5].

Lemma 2.7.

Let KK\subseteq\mathcal{M} be a compact pathwise connected set with piecewise C1C^{1} boundary which is the closure of its interior. Then

  1. (1)

    A nice preimage K~\widetilde{K} of KK always exists,

  2. (2)

    If γId\gamma\neq\emph{Id}, then γint(K~)int(K~)=\gamma\cdot\emph{int}(\widetilde{K})\cap\emph{int}(\widetilde{K})=\emptyset,

  3. (3)

    The set {γΓ:γK~K~}\{\gamma\in\Gamma:\gamma\cdot\widetilde{K}\cap\widetilde{K}\neq\emptyset\} is finite, and

  4. (4)

    If K1K2K_{1}\subseteq K_{2} are compact sets as above, then they admit nice preimages K~1K~2\widetilde{K}_{1}\subseteq\widetilde{K}_{2}.

Definition 2.8.

Let KK\subseteq\mathcal{M} be a compact set and K~\widetilde{K} a nice preimage of KK. The fundamental group of \mathcal{M} out of KK is the set ΓK~c\Gamma_{\widetilde{K}^{c}} of elements γΓ\gamma\in\Gamma for which there exists x,yK~x,y\in\widetilde{K} such that the geodesic segment [x,γy][x,\gamma\cdot y] touches pΓ1(K)p_{\Gamma}^{-1}(K) only at K~\widetilde{K} and γK~\gamma\cdot\widetilde{K}. In other words, such that

[x,γy]pΓ1(K)K~γK~.[x,\gamma\cdot y]\cap p_{\Gamma}^{-1}(K)\subseteq\widetilde{K}\cup\gamma\cdot\widetilde{K}.

The definition of a fundamental group out of a compact set KK depends upon the choice of a nice preimage K~\widetilde{K} of KK. The following proposition ([ST, Proposition 7.9 (1) and (3)]) clarifies the dependence of this definition on the choice of K~\widetilde{K}. It is worth noting that ΓKc\Gamma_{K^{c}} may not form a group in general.

Proposition 2.9.

Let KK\subseteq\mathcal{M} be a compact pathwise connected set with piecewise C1C^{1} boundary.

  1. (1)

    If γΓ\gamma\in\Gamma, then Γ(γK~)c=γΓK~cγ1\Gamma_{(\gamma\cdot\widetilde{K})^{c}}=\gamma\Gamma_{\widetilde{K}^{c}}\gamma^{-1}

  2. (2)

    If K~1\widetilde{K}_{1} and K~2\widetilde{K}_{2} are nice preimages of KK, then there exists a finite subset {γ1,,γk}\{\gamma_{1},\ldots,\gamma_{k}\} of Γ\Gamma, such that

    ΓK~2ci,j=1kγiΓK~1cγj1.\Gamma_{\widetilde{K}^{c}_{2}}\subset\bigcup_{i,j=1}^{k}\gamma_{i}\Gamma_{\widetilde{K}^{c}_{1}}\gamma_{j}^{-1}.
Definition 2.10.

Let KK\subseteq\mathcal{M} be a compact pathwise connected set with piecewise C1C^{1} boundary. The critical exponent out of KK is the exponent of convergence δKc\delta_{K^{c}} of the Poincaré series γΓK~cesd(o,γo)\sum_{\gamma\in\Gamma_{\widetilde{K}^{c}}}e^{-sd(o,\gamma\cdot o)}.

It follows from Proposition 2.9 that δKc\delta_{K^{c}} is well-defined and independent of the nice preimage K~\widetilde{K}. The critical exponent δKc\delta_{K^{c}} coincides with the exponential growth of the orbits of ΓK~c\Gamma_{\widetilde{K}^{c}}, that is

δKc=lim supRlog(#{γΓK~c:d(o,γo)R})R.\delta_{K^{c}}=\limsup_{R\to\infty}\frac{\log(\#\{\gamma\in\Gamma_{\widetilde{K}^{c}}:d(o,\gamma\cdot o)\leq R\})}{R}.

If K1,K2K_{1},K_{2} are compact sets with piecewise C1C^{1} boundaries such that K1int(K2)K_{1}\subseteq\text{int}(K_{2}), then δK2cδK1c\delta_{K_{2}^{c}}\leq\delta_{K_{1}^{c}} (see [ST, Proposition 7.9 (2)]).

Definition 2.11.

The critical exponent at infinity of \mathcal{M} is defined by

δΓ=inf{δKc:K is a compact set}.\delta_{\Gamma}^{\infty}=\inf\{\delta_{K^{c}}:K\subset\mathcal{M}\emph{ is a compact set}\}.

Observe that if (Kn)n(K_{n})_{n\in{\mathbb{N}}} is a sequence of compact pathwise connected sets with piecewise C1C^{1} boundaries such that Knint(Kn+1)K_{n}\subset\text{int}(K_{n+1}) and =nKn\mathcal{M}=\bigcup_{n\in{\mathbb{N}}}K_{n}, then

δΓ=limnδKnc.\delta_{\Gamma}^{\infty}=\lim_{n\to\infty}\delta_{K_{n}^{c}}.

2.5. Hausdorff dimension

Let EE be a Borel subset of ~\partial_{\infty}\widetilde{\mathcal{M}} and x~x\in\widetilde{\mathcal{M}}. For every s>0s>0 and δ>0\delta>0, define

δ,xs(E)=infiris,\mathcal{H}^{s}_{\delta,x}(E)=\inf\sum_{i}r_{i}^{s},

where the infimum is taken over all finite coverings {Bx(ξi,ri)}\{B_{x}(\xi_{i},r_{i})\} of EE by balls of radius riδr_{i}\leq\delta. Then define

xs(E)=limδ0δ,xs(E).\mathcal{H}^{s}_{x}(E)=\lim_{\delta\to 0}\mathcal{H}^{s}_{\delta,x}(E).

Remarkably, xs()\mathcal{H}_{x}^{s}(\cdot) defines a measure on ~\partial_{\infty}\widetilde{\mathcal{M}}, that is, the Hausdorff measure of dimension ss with respect to the metric dxd_{x}. It is well known that there exists s0s^{\star}\geqslant 0 such that

xs(E)={+ if 0s<s0 if s<s+.\mathcal{H}^{s}_{x}(E)=\begin{cases}+\infty&\text{ if }0\leq s<s^{\star}\\ 0&\text{ if }s^{\star}<s\leq+\infty.\end{cases}

The Hausdorff dimension of EE with respect to dxd_{x} is defined to be equal to ss^{\star}. Since dxd_{x} and dxd_{x^{\prime}} are comparable for every x,x~x,x^{\prime}\in\widetilde{\mathcal{M}} (see equation (1)), the Hausdorff dimension of subsets of ~\partial_{\infty}\widetilde{\mathcal{M}} is independent of x~x\in\widetilde{\mathcal{M}}. The Hausdorff dimension of E~E\subset\partial_{\infty}\widetilde{\mathcal{M}} with respect to any of the visual metrics is denoted by HD(E)\text{HD}(E).

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 EE be a Borel subset of ~\partial_{\infty}\widetilde{\mathcal{M}}. Assume that there exists C>0C>0 and a positive measure μ\mu on EE such that for every ξ~\xi\in\partial_{\infty}\widetilde{\mathcal{M}} and r>0r>0, we have

μ(Bx(ξ,r))Crs.\mu(B_{x}(\xi,r))\leq Cr^{s}.

Then HD(E)s\emph{HD}(E)\geq s.

We now recall the definition of the Hausdorff dimension of a measure.

Definition 2.13.

Let ν\nu be a finite Borel measure on ~\partial_{\infty}\widetilde{\mathcal{M}}. The Hausdorff dimension of ν\nu is defined by

HD(ν)=inf{HD(A):A~,ν(A)>0}.\emph{HD}(\nu)=\inf\{\emph{HD}(A):A\subseteq\partial_{\infty}\widetilde{\mathcal{M}},\ \nu(A)>0\}.

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 ν\nu be a finite Borel measure on ~\partial_{\infty}\widetilde{\mathcal{M}}. Then

HD(ν)=essinflim infε0logν(Bo(ξ,ε))logε,\emph{HD}(\nu)={\mathrm{ess}\inf}\liminf_{\varepsilon\to 0}\frac{\log\nu(B_{o}(\xi,\varepsilon))}{\log\varepsilon},

where the essential infimum is considered with respect to ν\nu.

The proposition mentioned above, along with Frostman’s Lemma, implies the following variational principle in relation to Hausdorff dimensions.

Theorem 2.15.

Let A~A\subseteq\partial_{\infty}\widetilde{\mathcal{M}} be a Borel set. Then

HD(A)=supνHD(ν),\emph{HD}(A)=\sup_{\nu}\emph{HD}(\nu),

where the supremum runs over all finite measures ν\nu on ~\partial_{\infty}\widetilde{\mathcal{M}} such that ν(~A)=0\nu(\partial_{\infty}\widetilde{\mathcal{M}}\setminus A)=0.

3. Piecewise geodesic curves

We say that ϕ:[0,T]~\phi:[0,T]\to\widetilde{\mathcal{M}} is a piecewise geodesic curve, where T+{+}T\in{\mathbb{R}}^{+}\cup\{+\infty\}, 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 T=+T=+\infty. In this case, there exists an increasing sequence of real numbers (tn)n(t_{n})_{n} converging to ++\infty, where t0=0t_{0}=0 and such that ϕ|[tn1,tn]\phi|_{[t_{n-1},t_{n}]} is a geodesic segment parametrized by arc-length for every nn\in{\mathbb{N}}. By definition, the distance between ϕ(tn1)\phi(t_{n-1}) and ϕ(tn)\phi(t_{n}) is given by n=tntn1\ell_{n}=t_{n}-t_{n-1}. The interior angle at ϕ(tn)\phi(t_{n}) is denoted by θn\theta_{n}, where an interior angle is by definition in [0,π][0,\pi]. Using this notation, we say that ϕ\phi has lengths (n)n(\ell_{n})_{n} and angles (θn)n(\theta_{n})_{n}.

Definition 3.1.

Let λ1\lambda\geq 1. A piecewise geodesic curve ϕ:[0,T]~\phi:[0,T]\to\widetilde{\mathcal{M}} is called a λ\lambda-quasi-geodesic if for every [c,d][0,T][c,d]\subset[0,T], we have that

length(ϕ([c,d]))λd(ϕ(c),ϕ(d)).\emph{length}(\phi([c,d]))\leq\lambda d(\phi(c),\phi(d)).
Definition 3.2.

Let λ1\lambda\geq 1 and L>0L>0. A piecewise geodesic curve ϕ:[0,T]~\phi:[0,T]\to\widetilde{\mathcal{M}} is called a (λ,L)(\lambda,L)-local-quasi-geodesic if for every [c,d][0,T][c,d]\subset[0,T] such that length(ϕ([c,d]))L\emph{length}(\phi([c,d]))\leqslant L then length(ϕ([c,d]))λd(ϕ(c),ϕ(d))\emph{length}(\phi([c,d]))\leqslant\lambda d(\phi(c),\phi(d)).

Let α(0,π)\alpha\in(0,\pi) be fixed. In the proofs we will need to shadow a piecewise geodesic curve with angles at least α\alpha 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 x,y,z~x,y,z\in\widetilde{\mathcal{M}} be points such that the interior angle θ\theta between [x,y][x,y] and [y,z][y,z] is at least α\alpha. The manifold ~\widetilde{\mathcal{M}} is a CAT(1)\text{CAT}(-1) space, since the sectional curvature is bounded above by 1-1. Denote by d2d_{\mathbb{H}^{2}} the hyperbolic metric in 2\mathbb{H}^{2}. Let x¯,y¯,z¯2\bar{x},\bar{y},\bar{z}\in\mathbb{H}^{2} be points such that d(x,y)=d2(x¯,y¯)=:bd(x,y)=d_{\mathbb{H}^{2}}(\bar{x},\bar{y})=:b, d(y,z)=d2(y¯,z¯)=:cd(y,z)=d_{\mathbb{H}^{2}}(\bar{y},\bar{z})=:c, and such that the angle at y¯\bar{y} of the geodesic triangle determined by x¯,y¯,z¯\bar{x},\bar{y},\bar{z} is equal to θ\theta. Then d(x,z)d2(x¯,z¯)=:ad(x,z)\geq d_{\mathbb{H}^{2}}(\bar{x},\bar{z})=:a (see [BH, Proposition 1.7, Chapter 2]). Moreover, using the hyperbolic law of cosine and that θα\theta\geq\alpha, we get

cosh(a)\displaystyle\cosh(a) =\displaystyle= cosh(b)cosh(c)sinh(b)sinh(c)cos(θ)\displaystyle\cosh(b)\cosh(c)-\sinh(b)\sinh(c)\cos(\theta)
\displaystyle\geq cosh(b)cosh(c)sinh(b)sinh(c)cos(α).\displaystyle\cosh(b)\cosh(c)-\sinh(b)\sinh(c)\cos(\alpha).

Observe that the last term is at least sin2(α/2)cosh(b+c)\sin^{2}(\alpha/2)\cosh(b+c), therefore

d(x,z)\displaystyle d(x,z) \displaystyle\geq a\displaystyle a
\displaystyle\geq Arccosh(sin2(α/2)cosh(b+c))\displaystyle\text{Arccosh}\left(\sin^{2}(\alpha/2)\cosh(b+c)\right)
\displaystyle\geq sin2(α/2)(b+c)\displaystyle\sin^{2}(\alpha/2)(b+c)
=\displaystyle= sin2(α/2)(d(x,y)+d(y,z)),\displaystyle\sin^{2}(\alpha/2)(d(x,y)+d(y,z)),

where the last inequality holds provided that b+cL0b+c\geq L_{0}, for L0=L0(α)L_{0}=L_{0}(\alpha) large enough. Set λ0(α)=1/sin2(α/2)\lambda_{0}(\alpha)=1/\sin^{2}(\alpha/2). Therefore,

(2) d(x,y)+d(y,z)λ0(α)d(x,z).\displaystyle d(x,y)+d(y,z)\leqslant\lambda_{0}(\alpha)d(x,z).

Let us now consider a piecewise geodesic curve ϕ:[0,]~\phi:[0,\infty]\to\widetilde{\mathcal{M}} with lengths (n)n(\ell_{n})_{n} and angles (θn)n(\theta_{n})_{n}, where θnα\theta_{n}\geqslant\alpha and nL0(α)\ell_{n}\geqslant L_{0}(\alpha), for every nn\in{\mathbb{N}}. We claim that ϕ\phi is a (λ0(α),L0(α))(\lambda_{0}(\alpha),L_{0}(\alpha))-local-quasi-geodesic. Note that if [c,d][0,][c,d]\subseteq[0,\infty] and

length(ϕ([c,d]))L0(α),\text{length}(\phi([c,d]))\leq L_{0}(\alpha),

then either ϕ([c,d])=[ϕ(c),ϕ(d)]\phi([c,d])=[\phi(c),\phi(d)], or there exists e(c,d)e\in(c,d) such that ϕ([c,d])=[ϕ(c),ϕ(e)][ϕ(e),ϕ(d)].\phi([c,d])=[\phi(c),\phi(e)]\cup[\phi(e),\phi(d)]. 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 (λ,L)(\lambda,L)-local-quasi-geodesic is a λ\lambda^{\prime}-quasi-geodesic, where λ1\lambda^{\prime}\geq 1 depends only on the constants λ,L,ϱ\lambda,L,\varrho. Here ϱ>0\varrho>0 denotes a positive constant for which ~\widetilde{\mathcal{M}} is a ϱ\varrho-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 λ\lambda-quasi-geodesic is contained in a DD-neighbourhood of a geodesic line, where DD depends only on λ\lambda and ϱ\varrho. Combining these results and the discussion above we conclude the following result:

Theorem 3.3.

Given α>0,\alpha>0, there exists D(α)>0D(\alpha)>0 such that any piecewise geodesic curve ϕ:[0,]~\phi:[0,\infty]\to\widetilde{\mathcal{M}} with lengths (n)n(\ell_{n})_{n} and angles (θn)n(\theta_{n})_{n}, where θnα\theta_{n}\geqslant\alpha and nL0(α)\ell_{n}\geqslant L_{0}(\alpha), for every nn\in{\mathbb{N}}, is contained in a D(α)D(\alpha)-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 α>0,\alpha>0, there exist positive constants L=L(α),C=C(α)L=L(\alpha),C=C(\alpha) such that the following holds: if ϕ:[0,+)~\phi:[0,+\infty)\to\widetilde{\mathcal{M}} is a piecewise geodesic curve with θnα\theta_{n}\geq\alpha and nL\ell_{n}\geq L for every nn\in{\mathbb{N}}, then the interior angle at ϕ(tn1)\phi(t_{n-1}) between the geodesic segments [ϕ(t0),ϕ(tn1)][\phi(t_{0}),\phi(t_{n-1})] and [ϕ(tn1),ϕ(tn)][\phi(t_{n-1}),\phi(t_{n})] is larger than 2α3\frac{2\alpha}{3}, and

(3) d(ϕ(t0),ϕ(tn))k=1nk(n1)C,d(\phi(t_{0}),\phi(t_{n}))\geq\sum_{k=1}^{n}\ell_{k}-(n-1)C,

for every nn\in{\mathbb{N}}.

Proof.

To begin, let us make some general observations. Suppose we have a geodesic triangle in ~\widetilde{\mathcal{M}} with vertices x,y,z{x,y,z}, whose angle at vertex yy is greater than 2α3\frac{2\alpha}{3}. Using the hyperbolic law of cosines and a comparison theorem, we can derive that there exists a constant C=C(α)>0C=C(\alpha)>0 such that

d(x,z)d(x,y)+d(y,z)C.d(x,z)\geqslant d(x,y)+d(y,z)-C.

Furthermore, if d(y,z),d(x,y)Ld(y,z),d(x,y)\geq L (where L=L(α)L=L(\alpha) is large enough), then the interior angle at vertex zz is at most α3\frac{\alpha}{3}. We will assume LC.L\geqslant C. Next, consider a point w~w\in\widetilde{\mathcal{M}} such that the interior angle at vertex zz of [y,z][z,w][y,z]\cup[z,w] is greater than α\alpha and d(z,w)Ld(z,w)\geqslant L. It follows that the interior angle at zz of [x,z][z,w][x,z]\cup[z,w] is at least 2α3\frac{2\alpha}{3} and we have that

d(x,w)d(x,z)+d(z,w)Cd(x,y)+d(y,z)+d(z,w)2C.d(x,w)\geqslant d(x,z)+d(z,w)-C\geqslant d(x,y)+d(y,z)+d(z,w)-2C.

Note that the triangle {x,z,w}\{x,z,w\} satisfies similar conditions to {x,y,z}\{x,y,z\}, so we can iterate this procedure.

By induction, it follows that the interior angle between [ϕ(t0),ϕ(tn)][\phi(t_{0}),\phi(t_{n})] and [ϕ(tn),ϕ(tn1)][\phi(t_{n}),\phi(t_{n-1})] is at most α3\frac{\alpha}{3}. Furthermore, we have

d(ϕ(t0),ϕ(tn))k=1n1k(n2)C.d(\phi(t_{0}),\phi(t_{n}))\geqslant\sum_{k=1}^{n-1}\ell_{k}-(n-2)C.

Since θnα\theta_{n}\geqslant\alpha, we conclude that the interior angle at ϕ(tn)\phi(t_{n}) between the geodesic segments [ϕ(t0),ϕ(tn)][\phi(t_{0}),\phi(t_{n})] and [ϕ(tn),ϕ(tn+1)][\phi(t_{n}),\phi(t_{n+1})] is greater than 2α3\frac{2\alpha}{3}.

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 δΓ.\delta_{\Gamma}^{\infty}. We separate the proof into two parts: we first prove that δΓ\delta_{\Gamma}^{\infty} is an upper bound for the Hausdorff dimension and then we prove it is a lower bound.

Fix a reference point o~o\in\widetilde{\mathcal{M}}. Let (Kn)n(K_{n})_{n\in{\mathbb{N}}} be an increasing sequence of compact pathwise connected sets with piecewise C1C^{1} boundaries such that =nKn\mathcal{M}=\bigcup_{n}K_{n}. Let K~n\widetilde{K}_{n} be a nice pre-image of KnK_{n} containing oo. We assume that K~nK~n+1\widetilde{K}_{n}\subset\widetilde{K}_{n+1}, for every nn\in{\mathbb{N}}, and that δΓ=limnδKnc\delta_{\Gamma}^{\infty}=\lim_{n\to\infty}\delta_{K_{n}^{c}}. For ξ~\xi\in\partial_{\infty}\widetilde{\mathcal{M}}, let vξT1~v_{\xi}\in T^{1}\widetilde{\mathcal{M}} be the unit vector based at oo pointing towards ξ\xi. Denote by π:T1~~\pi:T^{1}\widetilde{\mathcal{M}}\to\widetilde{\mathcal{M}} the canonical projection. Observe that

ΛΓ,radn1{ξ~:limT1T0TχΓK~n(π(gtvξ))𝑑t=0}.\Lambda_{\Gamma}^{\infty,rad}\subset\bigcap_{n\geq 1}\left\{\xi\in\partial_{\infty}\widetilde{\mathcal{M}}:\lim_{T\to\infty}\frac{1}{T}\int_{0}^{T}\chi_{\Gamma\cdot\widetilde{K}_{n}}(\pi(g_{t}v_{\xi}))dt=0\right\}.

4.1. The upper bound

Let ε>0\varepsilon>0 and nn\in{\mathbb{N}} large enough such that δKpc<δΓ+ε\delta_{K_{p}^{c}}<\delta_{\Gamma}^{\infty}+\varepsilon, for every pnp\geqslant n. Fix mnm\geqslant n. To ease the notation we write K=KmK=K_{m}, K~=K~m\widetilde{K}=\widetilde{K}_{m} and set K~R\widetilde{K}_{R} for the RR-neighbourhood of K~\widetilde{K}. Set α(0,1)\alpha\in(0,1) and for γΓ\gamma\in\Gamma, let γ^:[0,d(o,γo)]~\hat{\gamma}:[0,d(o,\gamma\cdot o)]\to\widetilde{\mathcal{M}} be the arc-length parametrization of the geodesic segment [o,γo][o,\gamma\cdot o]. Define

Sα={γΓ:1d(o,γo)0d(o,γo)χΓK~R(γ^(t))𝑑tα}S_{\alpha}=\Big{\{}\gamma\in\Gamma:\frac{1}{d(o,\gamma\cdot o)}\int_{0}^{d(o,\gamma\cdot o)}\chi_{\Gamma\cdot\widetilde{K}_{R}}(\hat{\gamma}(t))dt\leq\alpha\Big{\}}

and

Aα={γSα:d(o,γo)<+1}.A^{\alpha}_{\ell}=\{\gamma\in S_{\alpha}:\ell\leq d(o,\gamma\cdot o)<\ell+1\}.

Define 𝒟p\mathcal{D}_{p} as the set of diverging on average radial limit points ξΛΓ,rad\xi\in\Lambda_{\Gamma}^{\infty,rad} such that [o,ξ)[o,\xi) returns infinitely many times to ΓK~p\Gamma\cdot\widetilde{K}_{p}. For any δ>0\delta>0 set d1d\geq 1 such that c0edδ/2c_{0}e^{-d}\leq\delta/2, where c01c_{0}\geq 1 is the constant from Lemma 2.4 for r=diam(K)r={\rm diam}(K). Set Sα,d={γSα:d(o,γo)d}S_{\alpha,d}=\{\gamma\in S_{\alpha}:d(o,\gamma\cdot o)\geq d\}. It is straightforward to prove that the set of balls {Bo(ξo,γo,c0ed(o,γo))}γSα,d\{B_{o}(\xi_{o,\gamma o},c_{0}e^{-d(o,\gamma\cdot o)})\}_{\gamma\in S_{\alpha,d}} is a covering of 𝒟m\mathcal{D}_{m} by balls of diameter less than δ\delta. Hence

δs(𝒟m)\displaystyle\mathcal{H}^{s}_{\delta}(\mathcal{D}_{m}) \displaystyle\leq γSα,d(Ced(o,γo))s\displaystyle\sum_{\gamma\in S_{\alpha,d}}(Ce^{-d(o,\gamma\cdot o)})^{s}
\displaystyle\leq Csdesl#Aα.\displaystyle C^{s}\sum_{\ell\geq d}e^{-sl}\#A^{\alpha}_{\ell}.

We will prove that for α>0\alpha>0 small enough, we have

(4) lim sup1log(#Aα)δΓ+2ε,\limsup_{\ell\to\infty}\frac{1}{\ell}\log(\#A^{\alpha}_{\ell})\leq\delta^{\infty}_{\Gamma}+2\varepsilon,

which implies that δs(𝒟m)0\mathcal{H}^{s}_{\delta}(\mathcal{D}_{m})\to 0 as δ0\delta\to 0 (in this case dd\to\infty) whenever s>δΓ+2εs>\delta^{\infty}_{\Gamma}+2\varepsilon. This implies that

HD(𝒟m)δΓ+2ε,{\rm HD}(\mathcal{D}_{m})\leq\delta^{\infty}_{\Gamma}+2\varepsilon,

for every mnm\geqslant n. Finally, observe that ΛΓ,rad=mn𝒟m\Lambda_{\Gamma}^{\infty,rad}=\bigcup_{m\geqslant n}\mathcal{D}_{m}, and therefore HD(ΛΓ,rad)δΓ+2ε{\rm HD}(\Lambda_{\Gamma}^{\infty,rad})\leq\delta^{\infty}_{\Gamma}+2\varepsilon. Since ε>0\varepsilon>0 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 RR-neighborhood of a specified compact set in T1~T^{1}\widetilde{\mathcal{M}}. 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 AαA^{\alpha}_{\ell}.

Proposition 4.1.

Let T0>0T_{0}>0 and η>0\eta>0. Then, for every α(0,1]\alpha\in(0,1] and R2R\geq 2, there exists a positive number ψ=ψ(K~,η,α/R)\psi=\psi(\widetilde{K},\eta,\alpha/R) such that

(5) lim sup1log(#Aα)(1α)δKc+αδΓ+η+ψ(K~,η,α/R).\limsup_{\ell\to\infty}\frac{1}{\ell}\log(\#A^{\alpha}_{\ell})\leq(1-\alpha)\delta_{K^{c}}+\alpha\delta_{\Gamma}+\eta+\psi(\widetilde{K},\eta,\alpha/R).

Moreover, for η>0\eta>0 fixed, ψ(K~,η,α/R)\psi(\widetilde{K},\eta,\alpha/R) tends monotonically to 0 when α/R\alpha/R tends to 0.

Fix R=2R=2, η=ε/2\eta=\varepsilon/2 and let α\alpha be small enough so that

(1α)δKc+αδΓ+ψ(K~,η,α/R)<δKc+ε.(1-\alpha)\delta_{K^{c}}+\alpha\delta_{\Gamma}+\psi(\widetilde{K},\eta,\alpha/R)<\delta_{K^{c}}+\varepsilon.

By (5) and the choice of variables we obtain inequality (4).

4.2. The bound from below

To establish the lower bound of the Hausdorff dimension, we will construct a subset EΛΓ,radE\subset\Lambda_{\Gamma}^{\infty,rad} and a positive measure μ\mu on EE such that Frostman’s lemma applies with the appropriate exponent. Fix ε>0\varepsilon>0 and set s=δΓ2εs=\delta_{\Gamma}^{\infty}-2\varepsilon.

For every ξ~\xi\in\partial_{\infty}\widetilde{\mathcal{M}}, let ξt\xi_{t} be the arc-length parametrization of the geodesic ray [o,ξ)[o,\xi) with ξ0=o\xi_{0}=o. The closure of the set of points in ~\widetilde{\mathcal{M}} that projects orthogonally into [ξt,ξ)[\xi_{t},\xi) is denoted by S(ξ,t)S(\xi,t). Set D(ξ,t)=S(ξ,t)~D(\xi,t)=S(\xi,t)\cap\partial_{\infty}\widetilde{\mathcal{M}}. In other words, the set S(ξ,t)S(\xi,t) is the convex-hull of D(ξ,t)D(\xi,t) on ¯\overline{\mathcal{M}}. The following corresponds to Lemma 2.5 in [Sc]. Let ϱ>0\varrho>0 be the hyperbolicity constant asociated to ~\widetilde{\mathcal{M}}.

Lemma 4.2.

For every t2ϱt\geq 2\varrho, we have

Bo(ξ,e(t+ϱ))D(ξ,t)Bo(ξ,e(t2ϱ)).B_{o}\left(\xi,e^{-(t+\varrho)}\right)\subset D(\xi,t)\subset B_{o}\left(\xi,e^{-(t-2\varrho)}\right).

We consider an increasing sequence (Kn)n(K_{n})_{n} of compact pathwise connected sets with piecewise C1C^{1} boundaries such that =nKn\mathcal{M}=\bigcup_{n}K_{n}, and K~n\widetilde{K}_{n} is a nice preimage of KnK_{n} containing oo. We further assume that K~nK~n+1\widetilde{K}_{n}\subset\widetilde{K}_{n+1}, for every nn\in{\mathbb{N}}. Since s+ε<δΓs+\varepsilon<\delta_{\Gamma}^{\infty} and ~\partial_{\infty}\widetilde{\mathcal{M}} is compact, there exists ξ1~\xi_{1}\in\partial_{\infty}\widetilde{\mathcal{M}} such that for every t>2δt>2\delta and every n1n\geq 1 we have

γΓK~nc|γoS(ξ1,t)e(s+ε)d(o,γo)=.\sum_{\gamma\in\Gamma_{\widetilde{K}_{n}^{c}}|\gamma\cdot o\in S(\xi_{1},t)}e^{-(s+\varepsilon)d(o,\gamma\cdot o)}=\infty.

Since Γ\Gamma is non-elementary, there exists gΓg\in\Gamma such that ξ2:=gξ1ξ1\xi_{2}:=g\cdot\xi_{1}\neq\xi_{1}. Set B1:=S(ξ1,t)B_{1}:=S(\xi_{1},t) and B2:=S(ξ2,t)B_{2}:=S(\xi_{2},t), where we choose t2ϱt\geq 2\varrho large enough such that the closures of B1B_{1} and B2B_{2} are disjoints and o,goo,g\cdot o are not in B1B2B_{1}\cup B_{2}. There exists α(0,π)\alpha\in(0,\pi) such that for every xB1x\in B_{1}, yB2y\in B_{2} the angle between the geodesic segments [o,x][o,x] and [o,y][o,y] at oo is bounded from below by 2α2\alpha. Let D=D(α)>0D=D(\alpha)>0 be the constant obtained from Theorem 3.3, and fix c=max{c0,e2δ}c=\max\{c_{0},e^{2\delta}\} where c01c_{0}\geq 1 is the constant from Lemma 2.4 associated to r=2Dr=2D. Set ΓK~nc1=ΓK~nc\Gamma^{1}_{\widetilde{K}_{n}^{c}}=\Gamma_{\widetilde{K}_{n}^{c}} and ΓK~nc2=gΓK~nc\Gamma^{2}_{\widetilde{K}_{n}^{c}}=g\Gamma_{\widetilde{K}_{n}^{c}}. The equation above and the triangle inequality imply that

(6) γΓK~nc1|γoB1e(s+ε)d(o,γo)=,andγΓK~nc2|γoB2e(s+ε)d(o,γo)=,\sum_{\gamma\in\Gamma^{1}_{\widetilde{K}_{n}^{c}}|\gamma\cdot o\in B_{1}}e^{-(s+\varepsilon)d(o,\gamma\cdot o)}=\infty,\quad\mbox{and}\quad\sum_{\gamma\in\Gamma^{2}_{\widetilde{K}_{n}^{c}}|\gamma\cdot o\in B_{2}}e^{-(s+\varepsilon)d(o,\gamma\cdot o)}=\infty,

since B2B_{2} contains gS(ξ1,t+d(o,go))g\cdot S(\xi_{1},t+d(o,g\cdot o)).

Define A={x~:d(o,x)<+1}A_{\ell}=\{x\in\widetilde{\mathcal{M}}:\ell\leq d(o,x)<\ell+1\}. We claim that for every m{1,2}m\in\{1,2\} and nn\in{\mathbb{N}} we have

(7) lim supγΓK~ncm|γoBmAesd(o,γo)=.\displaystyle\limsup_{\ell\to\infty}\sum_{\gamma\in\Gamma^{m}_{\widetilde{K}_{n}^{c}}|\gamma\cdot o\in B_{m}\cap A_{\ell}}e^{-sd(o,\gamma\cdot o)}=\infty.

Indeed, if this is not the case there would exist nn\in{\mathbb{N}}, m{0,1}m\in\{0,1\} and C>0C>0 such that

γΓK~ncm|γoBmAesd(o,γo)C,\sum_{\gamma\in\Gamma^{m}_{\widetilde{K}_{n}^{c}}|\gamma\cdot o\in B_{m}\cap A_{\ell}}e^{-sd(o,\gamma\cdot o)}\leqslant C,

for every .\ell\in{\mathbb{N}}. Then

γΓK~ncm|γoBme(s+ε)d(o,γo)\displaystyle\sum_{\gamma\in\Gamma^{m}_{\widetilde{K}_{n}^{c}}|\gamma\cdot o\in B_{m}}e^{-(s+\varepsilon)d(o,\gamma o)} =\displaystyle= 1γΓK~ncm|γoBmAe(s+ε)d(o,γo)\displaystyle\sum_{\ell\geq 1}\sum_{\gamma\in\Gamma^{m}_{\widetilde{K}_{n}^{c}}|\gamma\cdot o\in B_{m}\cap A_{\ell}}e^{-(s+\varepsilon)d(o,\gamma o)}
\displaystyle\leq C1eε<,\displaystyle C\sum_{\ell\geq 1}e^{-\varepsilon\ell}<\infty,

which contradicts (6). It follows from (7) that

(8) lim supes#{γΓK~ncm|γoBmA}=\limsup_{\ell\to\infty}e^{-\ell s}\#\{\gamma\in\Gamma^{m}_{\widetilde{K}_{n}^{c}}|\gamma\cdot o\in B_{m}\cap A_{\ell}\}=\infty

for all m{1,2}m\in\{1,2\} and nn\in{\mathbb{N}}.

The next result can be deduced from classical geometric arguments.

Lemma 4.3.

Given q>0q>0 there exists L1L\geq 1 such that for every L\ell\geq L, if x,yAx,y\in A_{\ell} verify d(x,y)>qd(x,y)>q, then Bo(ξo,x,3ced(o,x))B_{o}(\xi_{o,x},3ce^{-d(o,x)}) and Bo(ξo,y,3ced(o,y))B_{o}(\xi_{o,y},3ce^{-d(o,y)}) are disjoint.

Choose a sequence ((n))n(\ell(n))_{n} such that

(9) limnnΔnk=1n(k)2D=0,\lim_{n\to\infty}\frac{n\Delta_{n}}{\sum_{k=1}^{n}\ell(k)-2D}=0,

where Δn\Delta_{n} is the diameter of K~n\widetilde{K}_{n}. This can be done since (n)\ell(n) can be chosen arbitrarily large once nn is given.

Since the set Bq={γΓ:d(γo,o)<q}B_{q}=\{\gamma\in\Gamma:d(\gamma\cdot o,o)<q\} is finite, the ball of radius qq centered at xΓox\in\Gamma\cdot o contains at most #Bq\#B_{q} points of Γo\Gamma\cdot o. This implies that any set AΓoA\subseteq\Gamma\cdot o has a subset of size at least #A/#Bq\#A/\#B_{q}, where each pair of points has distance at least qq. In particular, by Lemma 4.3 and by equation (8), for each m{1,2}m\in\{1,2\}, nn\in{\mathbb{N}} and =(n)L\ell=\ell(n)\geq L, there is c1c\geq 1 and a finite subset Vm,nBmA(n)V_{m,n}\subset B_{m}\cap A_{\ell(n)} of the set ΓKncmo\Gamma^{m}_{K_{n}^{c}}\cdot o such that

  1. (i)

    for any x,xVm,nx,x^{\prime}\in V_{m,n}, xxx\neq x^{\prime}, the balls

    Bo(ξo,x,3cesd(o,x))andBo(ξo,x,3cesd(o,x))B_{o}(\xi_{o,x},3ce^{-sd(o,x)})\quad\mbox{and}\quad B_{o}(\xi_{o,x^{\prime}},3ce^{-sd(o,x^{\prime})})

    are disjoint, and

  2. (ii)

    γ|γoVm,nesd(o,γo)es(n)#{γ|γoVm,n}1\sum_{\gamma|\gamma\cdot o\in V_{m,n}}e^{-sd(o,\gamma\cdot o)}\geq e^{-s\ell(n)}\#\{\gamma|\gamma\cdot o\in V_{m,n}\}\geq 1.

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 [o,x1][o,x_{1}], where x1V1,1x_{1}\in V_{1,1}. Recall that x1=γx1ox_{1}=\gamma_{x_{1}}\cdot o, for some γx1ΓK~1c1\gamma_{x_{1}}\in\Gamma^{1}_{\widetilde{K}_{1}^{c}}. There exists m1{1,2}m_{1}\in\{1,2\} such that the geodesic segment γx11[o,x1]=[γx11o,o]\gamma_{x_{1}}^{-1}[o,x_{1}]=[\gamma_{x_{1}}^{-1}\cdot o,o] forms an angle α\geq\alpha with the geodesic segment [o,x2][o,x_{2}], for every x2Vm1,2x_{2}\in V_{m_{1},2}. In particular, the piecewise geodesic path [o,x1]γx1[o,x2][o,x_{1}]\cup\gamma_{x_{1}}[o,x_{2}] has interior angle at x1x_{1} larger than α\alpha. Note that x2=γx2ox_{2}=\gamma_{x_{2}}\cdot o, where γx2ΓK~2cm1\gamma_{x_{2}}\in\Gamma^{m_{1}}_{\widetilde{K}_{2}^{c}}, so the piecewise geodesic curve has the form

[o,γx1o][γx1o,γx1γx2o].[o,\gamma_{x_{1}}\cdot o]\cup[\gamma_{x_{1}}\cdot o,\gamma_{x_{1}}\gamma_{x_{2}}\cdot o].

Second step of induction. We consider a piecewise geodesic curve of the form

(10) ϕn:=k=1n[gn1o,gno],\phi_{n}:=\bigcup_{k=1}^{n}[g_{n-1}\cdot o,g_{n}\cdot o],

with interior angles larger than α\alpha, where gn=γx1γxng_{n}=\gamma_{x_{1}}\cdot\ldots\cdot\gamma_{x_{n}} and γxkoVmk1,k\gamma_{x_{k}}\cdot o\in V_{m_{k-1},k}, mk1{1,2}m_{k-1}\in\{{1,2}\} for every k{1,,n}k\in\{1,\ldots,n\} and m0=1m_{0}=1. Note that gn1ϕng_{n}^{-1}\phi_{n} is a piecewise geodesic curve ending at oo, so there exists mn{1,2}m_{n}\in\{1,2\} such that gn1ϕng_{n}^{-1}\phi_{n} and [o,xn+1][o,x_{n+1}] forms a angle larger than α\alpha at oo, for every xn+1Vmn,n+1x_{n+1}\in V_{m_{n},n+1}. Since xn+1=γxn+1ox_{n+1}=\gamma_{x_{n+1}}\cdot o, for some γxn+1ΓK~n+1cmn\gamma_{x_{n+1}}\in\Gamma^{m_{n}}_{\widetilde{K}_{n+1}^{c}}, a piecewise geodesic curve ϕn+1\phi_{n+1} defined as in (10) has all interior angles larger than α\alpha.

This inductive construction defines a tree 𝒯\mathcal{T} in ~\widetilde{\mathcal{M}} for which each vertex xx at the (n1)(n-1)-step has at least #Vm,n\#V_{m,n} children yy, for m{1,2}m\in\{1,2\}. By Theorem 3.3 each piecewise geodesic curves of 𝒯\mathcal{T} is contained in some D(α)D(\alpha)-neighbourhood of a geodesic ray. By (i) these geodesic rays are different. Let us denote by EE the set of extremities at infinity of them. We also denote 𝒯(x)\mathcal{T}(x) the set of children of a vertex xx and 𝒯n\mathcal{T}_{n} the set of vertices at level nn. Note that 𝒯0={o}\mathcal{T}_{0}=\{o\}.

Lemma 4.4.

The set EE is contained in ΛΓ,rad\Lambda_{\Gamma}^{\infty,rad}.

Proof.

Let ξE\xi\in E. By construction a DD-neighbourhood of the geodesic ray [o,ξ)[o,\xi) contains a unique piecewise geodesic curve in 𝒯\mathcal{T} passing through infinitely many orbit elements γo\gamma\cdot o, so by definition ξ\xi is a radial point.

To prove that [o,ξ)[o,\xi) escapes in average we will use Condition (9). Let K~\widetilde{K} be any compact set in ~\widetilde{\mathcal{M}} containing a DD-neighbourhood of oo and denote by Δ\Delta its diameter. Then for any large enough n1n\geq 1 we have K~K~n\widetilde{K}\subset\widetilde{K}_{n}. We will assume without loss of generality that K~K~n\widetilde{K}\subset\widetilde{K}_{n} for any n1n\geq 1. Let (ξt)(\xi_{t}) be the arc-length parametrization of [o,ξ)[o,\xi). For T>0T>0, the interval [0,T][0,T] can be decomposed into the union

[0,T]=[0,t1][t1,t2][tk1,tk][tk,r][0,T]=[0,t_{1}]\cup[t_{1},t_{2}]\cup\ldots\cup[t_{k-1},t_{k}]\cup[t_{k},r]

where (ti)(t_{i}) is an increasing sequence of positive real numbers, k=k(T)k=k(T) and r>0r>0. Each time tit_{i} is defined to be the time for which ξt\xi_{t} is the closest point of [o,ξ)[o,\xi) from giog_{i}\cdot o. Recall that gi=γx1γxig_{i}=\gamma_{x_{1}}\cdot\ldots\cdot\gamma_{x_{i}}, where γxjΓKjcmj\gamma_{x_{j}}\in\Gamma^{m_{j}}_{K_{j}^{c}}, j{1,,i}j\in\{1,\ldots,i\}. By construction d(gio,ξti)Dd(g_{i}\cdot o,\xi_{t_{i}})\leq D, in particular |ti(mi1)|2D|t_{i}-\ell(m_{i-1})|\leq 2D. With this information, we have

1T0TχΓK~(ξt)𝑑t\displaystyle\frac{1}{T}\int_{0}^{T}\chi_{\Gamma\cdot\widetilde{K}}(\xi_{t})dt =\displaystyle= 1T0tjχΓK~(ξt)+1TtkrχΓK~(ξt)\displaystyle\frac{1}{T}\int_{0}^{t_{j}}\chi_{\Gamma\cdot\widetilde{K}}(\xi_{t})+\frac{1}{T}\int_{t_{k}}^{r}\chi_{\Gamma\cdot\widetilde{K}}(\xi_{t})
\displaystyle\leq 1T0tkχΓK~(ξt)+2TΔ,\displaystyle\frac{1}{T}\int_{0}^{t_{k}}\chi_{\Gamma\cdot\widetilde{K}}(\xi_{t})+\frac{2}{T}\Delta,

so we need to estimate now the number of times ξt\xi_{t} intersects ΓK~\Gamma\cdot\widetilde{K}. A priori, the number of times the geodesic segment [ξti,ξti+1][\xi_{t_{i}},\xi_{t_{i+1}}] intersects translations of K~\widetilde{K} might be large. We will prove this number to be bounded from above by a constant depending on gg and KK, not on ii. We will assume in addition that γxi=gγxi\gamma_{x_{i}}=g\cdot\gamma^{\prime}_{x_{i}}, with γxiΓKic\gamma^{\prime}_{x_{i}}\in\Gamma_{K^{c}_{i}}, which is the hardest situation since in the case γxiΓKic\gamma_{x_{i}}\in\Gamma_{K^{c}_{i}} the segment [ξti,ξti+1][\xi_{t_{i}},\xi_{t_{i+1}}] intersects translations of K~\widetilde{K} only at the beginning and at the end. Let h>0h>0 and assume that a hh-neighbourhood of K~\widetilde{K} is contained in K~1\widetilde{K}_{1}. Since the curvature of ~\widetilde{\mathcal{M}} is negative, there exists R>0R>0 such that for every z~z\in\widetilde{\mathcal{M}} with d(z,o)Rd(z,o)\geq R, then the geodesic segments [x,z][x,z] and [y,z][y,z] are hh-close for every xK~x\in\widetilde{K} and ygK~y\in g\widetilde{K}. We apply this to x=gi1ξtix=g_{i}^{-1}\xi_{t_{i}}, y=goy=g\cdot o and z=gi1ξti+1z=g_{i}^{-1}\xi_{t_{i+1}}.

\begin{overpic}[scale={.3}]{esc.png} \put(10.0,38.0){$\xi_{t_{i}}$} \put(9.0,30.0){$g_{i}\cdot o$} \put(3.0,19.0){$g_{i}\cdot\widetilde{K}$} \put(25.0,8.0){$g_{i}g\cdot o$} \put(11.0,2.0){$g_{i}g\cdot\widetilde{K}$} \put(85.0,38.0){$\xi_{t_{i+1}}$} \put(84.0,30.0){$g_{i+1}\cdot o$} \put(84.0,19.0){$g_{i+1}\cdot\widetilde{K}$} \put(66.0,5.0){$\partial B(g_{i}\cdot o,R)$} \put(30.0,43.0){$g_{i}[x,z]$} \put(35.0,28.0){$g_{i}[y,z]$} \end{overpic}

Suppose that γK~\gamma\cdot\widetilde{K} intersects [x,z][x,z] not only at the beginning and at the end. If d(o,γo)R+Δd(o,\gamma\cdot o)\geq R+\Delta, then a hh-neighbourhood of γK~\gamma\cdot\widetilde{K} intersects [y,z][y,z]. Since K~i\widetilde{K}_{i} contains a hh-neighbourhood of K~\widetilde{K}, then γK~i\gamma\cdot\widetilde{K}_{i} also intersects [y,z][y,z]. The later cannot happen since [g1y,g1z][g^{-1}y,g^{-1}z] intersects a Γ\Gamma-translation of K~i\widetilde{K}_{i} only at the beginning and at the end, so γK~i\gamma\cdot\widetilde{K}_{i} intersects [y,z][y,z] only if d(o,γo)R+Δd(o,\gamma\cdot o)\leq R+\Delta. In particular, given h>0h>0, the number Γ\Gamma-translations of K~\widetilde{K} intersecting [ξti,ξti+1][\xi_{t_{i}},\xi_{t_{i+1}}] is bounded from above by a constant σ\sigma depending on hh, K~\widetilde{K} and gg. Hence

1T0TχΓK~(ξt)𝑑t1Ti=1kti1tiχΓK~i(ξt)1TσkΔk.\frac{1}{T}\int_{0}^{T}\chi_{\Gamma\cdot\widetilde{K}}(\xi_{t})dt\leq\frac{1}{T}\sum_{i=1}^{k}\int_{t_{i-1}}^{t_{i}}\chi_{\Gamma\cdot\widetilde{K}_{i}}(\xi_{t})\leq\frac{1}{T}\sigma k\Delta_{k}.

On the other hand, by construction, we have

Ti=1ktii=1k(i)2DT\geq\sum_{i=1}^{k}t_{i}\geq\sum_{i=1}^{k}\ell(i)-2D

so

1T0TχΓK~(ξt)𝑑tσkΔki=1k(i)2D\frac{1}{T}\int_{0}^{T}\chi_{\Gamma\cdot\widetilde{K}}(\xi_{t})dt\leq\frac{\sigma k\Delta_{k}}{\sum_{i=1}^{k}\ell(i)-2D}

Since k+k\to+\infty as T+T\to+\infty, Condition (9) implies that the limit above when T+T\to+\infty is 0. Hence ξ\xi defines an escaping in average direction. ∎

For x~x\in\widetilde{\mathcal{M}} define rx:=ced(o,x)r_{x}:=ce^{-d(o,x)} and β(x):=Bo(ξo,x,rx)\beta(x):=B_{o}(\xi_{o,x},r_{x}). Note that if x,y𝒯nx,y\in\mathcal{T}_{n} are different, then β(x)β(y)=.\beta(x)\cap\beta(y)=\emptyset. Define En:=x𝒯nβ(x)E_{n}:=\bigcup_{x\in\mathcal{T}_{n}}\beta(x), and observe that E=nEnE=\bigcap_{n\in{\mathbb{N}}}E_{n}. By (i), we have

Bo(ξo,x1,2rx1)Bo(ξo,x2,2rx2)=,for x1x2 in 𝒯n.B_{o}(\xi_{o,x_{1}},2r_{x_{1}})\cap B_{o}(\xi_{o,x_{2}},2r_{x_{2}})=\emptyset,\quad\mbox{for }x_{1}\neq x_{2}\text{ in }\mathcal{T}_{n}.

Moreover, if y𝒯(x)y\in\mathcal{T}(x), then

(11) Bo(ξo,y,2ry)Bo(ξo,x,2rx).B_{o}(\xi_{o,y},2r_{y})\subseteq B_{o}(\xi_{o,x},2r_{x}).

In order to prove (11) first note that ξo,y\xi_{o,y} is in the shadow at infinity from oo of the ball B(x,2D)B(x,2D) since both xx and yy are at distance at most DD of a geodesic curve with extremity at infinity in EE. In particular ξo,yβ(x)\xi_{o,y}\in\beta(x). By triangle inequality, if ηBo(ξo,y,2ry)\eta\in B_{o}(\xi_{o,y},2r_{y}), then

do(η,ξo,x)\displaystyle d_{o}(\eta,\xi_{o,x}) \displaystyle\leq do(η,ξo,y)+do(ξo,y,ξo,x)\displaystyle d_{o}(\eta,\xi_{o,y})+d_{o}(\xi_{o,y},\xi_{o,x})
\displaystyle\leq 2ry+rx\displaystyle 2r_{y}+r_{x}
=\displaystyle= 2ced(o,y)+ced(o,x).\displaystyle 2ce^{-d(o,y)}+ce^{-d(o,x)}.

Following the proof of Proposition 3.4, we know that d(o,y)d(o,x)+d(x,y)Cd(o,y)\geq d(o,x)+d(x,y)-C, where C>0C>0 is a constant depending only on α\alpha (and the bounds of the sectional curvatures). Since we can choose xx and yy to be arbitrarily large, we will assume d(x,y)Clog(2)d(x,y)-C\geq\log(2). Hence

ed(o,y)12ed(o,x).e^{-d(o,y)}\leq\frac{1}{2}e^{-d(o,x)}.

Finally, we get do(η,ξo,x)2ced(o,x)=2rxd_{o}(\eta,\xi_{o,x})\leq 2ce^{-d(o,x)}=2r_{x}, which ends the proof of equation (11).

We are now able to define a measure μ\mu on EE by setting μ(E0)=1\mu(E_{0})=1 and

μ(β(y))=esd(o,y)z𝒯(x)esd(o,z)μ(β(x)),\mu(\beta(y))=\frac{e^{-sd(o,y)}}{\sum_{z\in\mathcal{T}(x)}e^{-sd(o,z)}}\mu(\beta(x)),

whenever y𝒯(x)y\in\mathcal{T}(x).

It follows by induction and (ii) that μ(β(x))esd(o,x)\mu(\beta(x))\leqslant e^{-sd(o,x)}, for every x𝒯n.x\in\mathcal{T}_{n}. Let β=Bo(ξo,z,t)\beta=B_{o}(\xi_{o,z},t) and nn the smallest natural number such that there is y𝒯ny\in\mathcal{T}_{n} where β\beta intersects β(y)\beta(y) but it is not contained in Bo(ξo,y,2ry)B_{o}(\xi_{o,y},2r_{y}). Let xx be the parent of yy. Since βBo(ξo,x,2rx)\beta\subseteq B_{o}(\xi_{o,x},2r_{x}) we have that

(12) μ(β)μ(Bo(ξo,x,2rx))=μ(β(x))esd(o,x)esd(o,y)esd(x,y).\mu(\beta)\leqslant\mu(B_{o}(\xi_{o,x},2r_{x}))=\mu(\beta(x))\leqslant e^{-sd(o,x)}\leqslant e^{-sd(o,y)}e^{sd(x,y)}.

Set ηββ(y)\eta\in\beta\cap\beta(y) and ηβBo(ξo,y,2ry)c\eta^{\prime}\in\beta\cap B_{o}(\xi_{o,y},2r_{y})^{c}. Note that do(η,η)ryd_{o}(\eta,\eta^{\prime})\geqslant r_{y}, and therefore η,ηologry=d(o,y)logc\langle\eta,\eta^{\prime}\rangle_{o}\leqslant-\log r_{y}=d(o,y)-\log c. Note also that d(ξo,z,η),d(ξo,z,η)td(\xi_{o,z},\eta),d(\xi_{o,z},\eta^{\prime})\leqslant t. By ϱ\varrho-hyperbolicity, we get

η,ηoinf{η,ξo,zo,ξo,z,ηo}ϱlogtϱ.\langle\eta,\eta^{\prime}\rangle_{o}\geqslant\inf\{\langle\eta,\xi_{o,z}\rangle_{o},\langle\xi_{o,z},\eta^{\prime}\rangle_{o}\}-\varrho\geqslant-\log t-\varrho.

In conclusion, we obtain logtϱd(o,y)logc-\log t-\varrho\leqslant d(o,y)-\log c, and therefore

(13) c1eϱted(o,y).c^{-1}e^{\varrho}t\geqslant e^{-d(o,y)}.

We choose ((n))n(\ell(n))_{n} such that

limn(n)k=1n(k)nC=0,\lim_{n\to\infty}\frac{\ell(n)}{\sum^{n}_{k=1}\ell(k)-nC}=0,

where C>0C>0 is chosen according to (3). This can be done as consequence of the following technical lemma.

Lemma 4.5.

Let (ln)n(l_{n})_{n} be an increasing sequence of real numbers converging to ++\infty. Then, there exists a sequence (hn)n(h_{n})_{n} of positive integers such that for r1=0r_{1}=0 and rn=k=1n1hkr_{n}=\sum_{k=1}^{n-1}h_{k}, for n>1n>1, if

Lrn+k:=ln,for all 1khn,L_{r_{n}+k}:=l_{n},\ \text{for all}\ 1\leq k\leq h_{n},

then

limnLnk=1nLknC=0.\lim_{n\to\infty}\frac{L_{n}}{\sum^{n}_{k=1}L_{k}-nC}=0.
Proof.

Let assume for the moment that (hn)n(h_{n})_{n} is any sequence of positive real numbers. Then, for 1mhn+11\leq m\leq h_{n+1}, we have

Lrn+mk=1rn+mLk(rn+m)C\displaystyle\frac{L_{r_{n}+m}}{\sum_{k=1}^{r_{n}+m}L_{k}-(r_{n}+m)C} \displaystyle\leq ln+1i=1nhi(liC)\displaystyle\frac{l_{n+1}}{\sum_{i=1}^{n}h_{i}(l_{i}-C)}
\displaystyle\leq ln+1hn(lnC).\displaystyle\frac{l_{n+1}}{h_{n}(l_{n}-C)}.

So the numbers (hn)n(h_{n})_{n} need to be chosen so that

limn+ln+1hn(lnC)=0.\lim_{n\to+\infty}\frac{l_{n+1}}{h_{n}(l_{n}-C)}=0.

Remark 4.6.

The construction of EE and the numbers (n)\ell(n) depend on the compact sets (Kn)n(K_{n})_{n}. Hence, according to the previous lemma, the suitable sequence (n)\ell(n) can be obtained by making the sequence (Kn)n(K_{n})_{n} 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 ln=(n)l_{n}=\ell(n). We therefore get for large enough n1n\geq 1

(n+1)ε(k=1n+1(k)(n+1)C).\ell(n+1)\leq\varepsilon\left(\sum_{k=1}^{n+1}\ell(k)-(n+1)C\right).

Note that each (k)\ell(k) is smaller than the length of a geodesic segment ending in 𝒯k\mathcal{T}_{k}, so

k=0n+1(k)(n+1)Cd(o,y)\sum_{k=0}^{n+1}\ell(k)-(n+1)C\leq d(o,y)

from (3). Using (12) and the fact that d(x,y)(n+1)+1d(x,y)\leqslant\ell(n+1)+1, this already implies that

μ(β)ese(sε)d(o,y).\mu(\beta)\leq e^{s}e^{-(s-\varepsilon)d(o,y)}.

By (13), we finally obtain

μ(β)Kts(1ε)\mu(\beta)\leq Kt^{s(1-\varepsilon)}

for some constant K>0K>0, which give us the lower bound HD(E)(δΓ2ε)(1ε)\text{HD}(E)\geq(\delta_{\Gamma}^{\infty}-2\varepsilon)(1-\varepsilon) through Frostman’s Lemma. Since EΛΓ,radE\subset\Lambda_{\Gamma}^{\infty,rad} and ε>0\varepsilon>0 is arbitrary, we obtain that HD(ΛΓ,rad)δΓ.\text{HD}(\Lambda_{\Gamma}^{\infty,rad})\geq\delta_{\Gamma}^{\infty}.

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 σ\sigma-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 Ω\Omega of the geodesic flow, and ergodic if every flow invariant set have either full or zero measure. We denote by M(T1,g)M_{\infty}(T^{1}\mathcal{M},g) the space of geodesic flow invariant σ\sigma-finite Borel measures on T1T^{1}\mathcal{M} that are infinite, ergodic and conservative. Let ΩDA\Omega_{DA} be the set of diverging on average vectors in the non-wandering set.

Lemma 5.1.

If mM(T1,g)m\in M_{\infty}(T^{1}\mathcal{M},g), then m(ΩΩDA)=0m(\Omega\setminus\Omega_{DA})=0.

Proof.

It is a consequence of Hopf’s ergodic theorem that for every fL1(m)f\in L^{1}(m) and mm-a.e. vT1v\in T^{1}\mathcal{M}, we have that

limT1T0Tf(gtv)𝑑t=0.\lim_{T\to\infty}\frac{1}{T}\int_{0}^{T}f(g_{t}v)dt=0.

In particular, this applies to f=χWf=\chi_{W}, where WT1W\subset T^{1}\mathcal{M} is a compact set. We conclude that, on average, a generic point of mm 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 vT1v\in T^{1}\mathcal{M}, r>0r>0 and T>0T>0, a (T,r)(T,r)-dynamical ball is the set

BT(v,r)={wT1:dS(gtv,gtw)r for every 0tT},B_{T}(v,r)=\{w\in T^{1}\mathcal{M}:d_{S}(g_{t}v,g_{t}w)\leq r\text{ for every }0\leq t\leq T\},

where dSd_{S} denotes the Sasaki metric on T1T^{1}\mathcal{M}.

Definition 5.2.

Let mm be a geodesic flow invariant σ\sigma-finite Borel measure on T1T^{1}\mathcal{M}. The measure-theoretic entropy h(m)h(m) of mm is defined as

h(m)=essinfmlimr0lim infT1Tlogm(BT(v,r)).h(m)={\mathrm{ess}\inf}_{m}\lim_{r\to 0}\liminf_{T\to\infty}-\frac{1}{T}\log m(B_{T}(v,r)).

It worth noticing that Lipschitz equivalent metrics on T1T^{1}\mathcal{M} lead to the same entropy value. Moreover, it is a consequence of Brin-Katok entropy formula that if mm is an ergodic probability measure, then h(m)h(m) coincides with the entropy of mm defined using partitions.

For a vector vT1v\in T^{1}\mathcal{M}, we denote by v(+)v(+\infty) the end point at infinity of the geodesic ray {π(gt(v))}t0\{\pi(g_{t}(v))\}_{t\geqslant 0}. Let Φ:T1~~\Phi:T^{1}\widetilde{\mathcal{M}}\to\partial_{\infty}\widetilde{\mathcal{M}} be the map defined by Φ(v)=v(+)\Phi(v)=v(+\infty). The proof of our next result follows closely the strategy considered by F. Ledrappier in [Led, Proposition 4.4].

Theorem 5.3.

If mM(T1,g)m\in M_{\infty}(T^{1}\mathcal{M},g), then h(m)δΓh(m)\leq\delta_{\Gamma}^{\infty}.

Proof.

Let vT1~v\in T^{1}\widetilde{\mathcal{M}} be a point in the support of mm and r>0r>0 smaller than the injectivity radius at vv. Let v~\tilde{v} be a lift of vv to T1~T^{1}\widetilde{\mathcal{M}}. The restriction of mm to B(v,r)B(v,r) can be lifted to a measure m~\tilde{m} on B(v~,r)B(\tilde{v},r). Set ν:=Φ(m~)\nu:=\Phi_{*}(\tilde{m}). Observe that since mm is conservative, the support of ν\nu is a subset of ΛΓrad\Lambda_{\Gamma}^{rad}. Moreover, it follows from Lemma 5.1 that the support of ν\nu is a subset of ΛΓ,rad\Lambda_{\Gamma}^{\infty,rad}. We claim that

(14) h(m)HD(ν).\displaystyle h(m)\leqslant\text{HD}(\nu).

Inequality (14), together with Theorem 1.2 and Theorem 2.15, finishes the proof. Note that there exists C=C(v,o)>0C=C(v,o)>0 such that for every w~B(v~,r)\tilde{w}\in B(\tilde{v},r), the image of BT(w~,r)B_{T}(\tilde{w},r) under Φ\Phi contains Bo(w~(+),c0(r)1eT/C)B_{o}(\tilde{w}(+\infty),c_{0}(r)^{-1}e^{-T}/C), where c0(r)c_{0}(r) is the constant given by Lemma 2.4. Then,

lim infT1Tlogm(BT(w~,r))lim infT1Tlogν(Bo(w~(+),c0(r)1eT/C)),\liminf_{T\to\infty}-\frac{1}{T}\log m(B_{T}(\tilde{w},r))\leqslant\liminf_{T\to\infty}-\frac{1}{T}\log\nu(B_{o}(\tilde{w}(+\infty),c_{0}(r)^{-1}e^{-T}/C)),

and therefore h(m)HD(ν)h(m)\leqslant\text{HD}(\nu) (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 SL3()\SL3()SL3({\mathbb{Z}})\backslash SL3({\mathbb{R}}). 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.