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

Ruelle’s inequality and Pesin’s formula for Anosov geodesic flows in non-compact manifolds

Alexander Cantoral Instituto de Matemática, Universidade Federal do Rio de Janeiro, CEP 21941-909, Rio de Janeiro, Brazil [email protected]  and  Sergio Romaña Instituto de Matemática, Universidade Federal do Rio de Janeiro, CEP 21941-909, Rio de Janeiro, Brazil [email protected]
Abstract.

In this paper we prove Ruelle’s inequality for the geodesic flow in non-compact manifolds with Anosov geodesic flow and some assumptions on the curvature. In the same way, we obtain the Pesin’s formula for Anosov geodesic flow in non-compact manifolds with finite volume.

Keywords: Anosov geodesic flow, Jacobi field, Lyapunov exponents, Ruelle’s inequality, Pesin’s formula.
Mathematics Subject Classification (2010): 37D40, 53C20.

1. Introduction

Ruelle in [19] proved an important result in ergodic theory relating entropy and Lyapunov exponents. More precisely, if f:MMf:M\rightarrow M is a C1C^{1}-diffeomorphism on a compact manifold and μ\mu is an ff-invariant probability measure on MM, then

(1) hμ(f)𝒳i(x)>0𝒳i(x)dim(Hi(x))dμ(x),\displaystyle h_{\mu}(f)\leq\int\sum_{\mathcal{X}_{i}(x)>0}\mathcal{X}_{i}(x)\cdot\dim(H_{i}(x))d\mu(x),

where hμ(f)h_{\mu}(f) is the entropy, {𝒳i(x)}\left\{\mathcal{X}_{i}(x)\right\} is the set of Lyapunov exponents at xMx\in M and dim(Hi(x))\dim(H_{i}(x)) is the multiplicity of 𝒳i(x)\mathcal{X}_{i}(x). In situations involving non-compact manifolds, Ruelle’s inequality may be compromised. For example, Riquelme in [17] constructed diffeomorphisms defined on non-compact manifolds with an invariant measure with positive entropy and the sum of the positive Lyapunov exponents was equal to zero. However, in recent years, certain findings have been achieved that, in particular situations, offer the possibility of verifying Ruelle’s inequality in non-compact contexts. Liao and Qiu in [9] showed Ruelle’s inequality for general Riemannian manifolds under an integrable condition. Riquelme in [18] showed Ruelle’s inequality for the geodesic flow in manifolds with pinched negative sectional curvature with some condition about the derivatives of the sectional curvature.
The main goal of this work is to prove Ruelle’s inequality for the geodesic flow on the unit tangent bundle of a non-compact manifold with Anosov geodesic flow and some assumptions on the curvature. More precisely,

Theorem 1.1.

Let MM be a complete Riemannian manifold with Anosov geodesic flow. Assume that the curvature tensor and the derivative of the curvature tensor are both uniformly bounded. Then, for every ϕt\phi^{t}-invariant probability measure μ\mu on SMSM, we have

hμ(ϕ)SM𝒳i(θ)>0𝒳i(θ)dim(Hi(θ))dμ(θ).\displaystyle h_{\mu}(\phi)\leq\int\limits_{SM}\sum_{\mathcal{X}_{i}(\theta)>0}\mathcal{X}_{i}(\theta)\cdot\dim(H_{i}(\theta))d\mu(\theta).

We can see that this result generalizes what Riquelme demonstrated in [18] since the Anosov geodesic flows encompass the manifolds with pinched negative curvature.
The question arises as to under what conditions equality can be achieved in (1). For example, when the manifold is compact, the diffeomorphism is C1+αC^{1+\alpha} and the measure is absolutely continuous with respect to the Lebesgue measure, Pesin showed in [16] that (1) is actually an equality, called Pesin’s formula. Our second result deals with the equality case of Theorem 1.1. In this case, we suppose that the manifold has finite volume.

Theorem 1.2.

Let MM be a complete Riemannian manifold with finite volume and Anosov geodesic flow, where the flow is C1C^{1}-Hölder. Assume that the curvature tensor and the derivative of the curvature tensor are both uniformly bounded. Then, for every ϕt\phi^{t}-invariant probability measure μ\mu on SMSM which is absolutely continuous relative to the Lebesgue measure, we have

hμ(ϕ)=SM𝒳i(θ)>0𝒳i(θ)dim(Hi(θ))dμ(θ).\displaystyle h_{\mu}(\phi)=\int\limits_{SM}\sum_{\mathcal{X}_{i}(\theta)>0}\mathcal{X}_{i}(\theta)\cdot\dim(H_{i}(\theta))d\mu(\theta).

Structure of the Paper:

In section 2, we introduce the notations and geometric tools that we use in the paper. In section 3, we prove the existence of Oseledec’s decomposition for the flow at time t=1t=1. In section 4, we explore certain results that will allow us to deal with the challenge of non-compactness of the manifold. Using the strategies exhibited in [1] to prove the Ruelle’s inequality for diffeomorphisms in the compact case, we prove Theorem 1.1 in section 5. Finally, in section 6 we prove Theorem 1.2 using techniques applied by Mañe in [11].

2. Preliminaries and notation

Throughout this paper, M=(M,g)M=(M,g) will denote a complete Riemannian manifold without boundary of dimension n2n\geq 2, TMTM is the tangent bundle, SMSM its unit tangent bundle and π:TMM\pi:TM\rightarrow M will denote the canonical projection, that is, π(x,v)=x\pi(x,v)=x for (x,v)TM(x,v)\in TM.

2.1. Geodesic flow

Given θ=(x,v)TM\theta=(x,v)\in TM, we denote by γθ\gamma_{\theta} the unique geodesic with initial conditions γθ(0)=x\gamma_{\theta}(0)=x and γθ(0)=v\gamma^{\prime}_{\theta}(0)=v. The geodesic flow is a family of CC^{\infty}-diffeomorphisms ϕt:TMTM\phi^{t}:TM\rightarrow TM, where tt\in\mathbb{R}, given by

ϕt(θ)=(γθ(t),γθ(t)).\displaystyle\phi^{t}(\theta)=(\gamma_{\theta}(t),\gamma^{\prime}_{\theta}(t)).

Since geodesics travel with constant speed, we have that ϕt\phi^{t} leaves SMSM invariant. The geodesic flow generates a vector field GG on TMTM given by

G(θ)=ddt|t=0ϕt(θ)=ddt|t=0(γθ(t),γθ(t)).\displaystyle G(\theta)=\left.\dfrac{d}{dt}\right|_{t=0}\phi^{t}(\theta)=\left.\dfrac{d}{dt}\right|_{t=0}\left(\gamma_{\theta}(t),\gamma^{\prime}_{\theta}(t)\right).

For each θ=(x,v)TM\theta=(x,v)\in TM, let VV be the vertical subbundle of TMTM whose fiber at θ\theta is given by Vθ=kerdπθV_{\theta}=\ker d\pi_{\theta}. Let K:TTMTMK:TTM\rightarrow TM be the connection map induced by the Riemannian metric (see [15]) and denotes by HH the horizontal subbundle of TMTM whose fiber at θ\theta is given by Hθ=kerKθH_{\theta}=\ker K_{\theta}. The maps dπθ|Hθ:HθTxM\left.d\pi_{\theta}\right|_{H_{\theta}}:H_{\theta}\rightarrow T_{x}M and Kθ|Vθ:VθTxM\left.K_{\theta}\right|_{V_{\theta}}:V_{\theta}\rightarrow T_{x}M are linear isomorphisms. This implies that TθTM=HθVθT_{\theta}TM=H_{\theta}\oplus V_{\theta} and the map jθ:TθTMTxM×TxMj_{\theta}:T_{\theta}TM\rightarrow T_{x}M\times T_{x}M given by

(2) jθ(ξ)=(dπθ(ξ),Kθ(ξ))\displaystyle j_{\theta}(\xi)=(d\pi_{\theta}(\xi),K_{\theta}(\xi))

is a linear isomorphism. Furthermore, we can identify every element ξTθTM\xi\in T_{\theta}TM with the pair jθ(ξ)j_{\theta}(\xi). Using the decomposition TθTM=HθVθT_{\theta}TM=H_{\theta}\oplus V_{\theta}, we endow the tangent bundle TMTM with a special Riemannian metric that makes HθH_{\theta} and VθV_{\theta} orthogonal. This metric is called the Sasaki metric and it’s given by

ξ,ηθ=dπθ(ξ),dπθ(η)x+Kθ(ξ),Kθ(η)x.\displaystyle\left\langle\xi,\eta\right\rangle_{\theta}=\left\langle d\pi_{\theta}(\xi),d\pi_{\theta}(\eta)\right\rangle_{x}+\left\langle K_{\theta}(\xi),K_{\theta}(\eta)\right\rangle_{x}.

From now on, we work with the Sasaki metric restricted to the unit tangent bundle SMSM. To begin with, it is valid to ask if SMSM is a complete Riemannian manifold with this metric.

Lemma 2.1.

Let MM be a complete Riemannian manifold. Then SMSM is a complete metric space with the Sasaki metric.

Proof.

Let θ,ωSM\theta,\omega\in SM and γ:[0,1]SM\gamma:[0,1]\rightarrow SM be a curve joining θ\theta and ω\omega. By the identification (2) we can write

l(γ)\displaystyle l(\gamma) =01γ(t)𝑑t\displaystyle=\int_{0}^{1}\left\|\gamma^{\prime}(t)\right\|dt
=01(dπγ(t)(γ(t))2+Kγ(t)(γ(t))2)1/2𝑑t\displaystyle=\int_{0}^{1}\left(\left\|d\pi_{\gamma(t)}(\gamma^{\prime}(t))\right\|^{2}+\left\|K_{\gamma(t)}(\gamma^{\prime}(t))\right\|^{2}\right)^{1/2}dt
01dπγ(t)(γ(t))𝑑t\displaystyle\geq\int_{0}^{1}\left\|d\pi_{\gamma(t)}(\gamma^{\prime}(t))\right\|dt
=01(πγ)(t)𝑑t\displaystyle=\int_{0}^{1}\left\|\left(\pi\circ\gamma\right)^{\prime}(t)\right\|dt
=l(πγ).\displaystyle=l(\pi\circ\gamma).

This implies that

(3) d(θ,ω)d(π(θ),π(ω))\displaystyle d(\theta,\omega)\geq d(\pi(\theta),\pi(\omega))

for any two points θ,ωSM\theta,\omega\in SM. Let {(pn,vn)}n\left\{(p_{n},v_{n})\right\}_{n\in\mathbb{N}} be a Cauchy sequence in SMSM. By (3) we have that {pn}n\left\{p_{n}\right\}_{n\in\mathbb{N}} is a Cauchy sequence in MM. Since MM is complete, there is pMp\in M such that limn+pn=p\displaystyle\lim_{n\rightarrow+\infty}p_{n}=p. If we consider the compact set X={(q,v)SM:d(q,p)1}X=\left\{(q,v)\in SM:d(q,p)\leq 1\right\}, for nn0n\geq n_{0} we have that (pn,vn)X(p_{n},v_{n})\in X and therefore the Cauchy sequence converges in SMSM. ∎

The sectional curvature of SMSM with the Sasaki metric can be calculated from the curvature tensor and the derivative of the curvature tensor of MM as explained in [7]: Let Π\Pi be a plane in T(x,v)SMT_{(x,v)}SM and choose an orthonormal basis {(v1,w1),(v2,w2)}\left\{(v_{1},w_{1}),(v_{2},w_{2})\right\} for Π\Pi satisfying vi2+wi2=1\left\|v_{i}\right\|^{2}+\left\|w_{i}\right\|^{2}=1, for i=1,2i=1,2, and v1,v2=w1,w2=0\left\langle v_{1},v_{2}\right\rangle=\left\langle w_{1},w_{2}\right\rangle=0. Then the Sasaki sectional curvature of Π\Pi is given by

KSas(Π)=\displaystyle K_{Sas}(\Pi)= Rx(v1,v2)v1,v2+3Rx(v1,v2)w1,w2+w12w22\displaystyle\left\langle R_{x}(v_{1},v_{2})v_{1},v_{2}\right\rangle+3\left\langle R_{x}(v_{1},v_{2})w_{1},w_{2}\right\rangle+\left\|w_{1}\right\|^{2}\left\|w_{2}\right\|^{2}
34Rx(v1,v2)v2+14Rx(v,w2)v12+14Rx(v,w1)v22\displaystyle-\dfrac{3}{4}\left\|R_{x}(v_{1},v_{2})v\right\|^{2}+\dfrac{1}{4}\left\|R_{x}(v,w_{2})v_{1}\right\|^{2}+\dfrac{1}{4}\left\|R_{x}(v,w_{1})v_{2}\right\|^{2}
+12Rx(v,w1)w2,Rx(v,w2)v1Rx(v,w1)v1,Rx(v,w2)v2\displaystyle+\dfrac{1}{2}\left\langle R_{x}(v,w_{1})w_{2},R_{x}(v,w_{2})v_{1}\right\rangle-\left\langle R_{x}(v,w_{1})v_{1},R_{x}(v,w_{2})v_{2}\right\rangle
(4) +(v1R)x(v,w2)v2,v1+(v2R)x(v,w1)v1,v2.\displaystyle+\left\langle(\nabla_{v_{1}}R)_{x}(v,w_{2})v_{2},v_{1}\right\rangle+\left\langle(\nabla_{v_{2}}R)_{x}(v,w_{1})v_{1},v_{2}\right\rangle.

This equality shows that if the curvature tensor of MM and its derivatives are bounded, then the sectional curvature of SMSM with the Sasaki metric is also bounded. This property is crucial as it allows us to compare volumes between subsets of TSMTSM and subsets of SMSM using the exponential map of SMSM (see Lemma 5.3).

The types of geodesic flows discussed in this paper are the Anosov geodesic flows, whose definition follows below.
We say that the geodesic flow ϕt:SMSM\phi^{t}:SM\rightarrow SM is of Anosov type if T(SM)T(SM) has a continuous splitting T(SM)=EsGEuT(SM)=E^{s}\oplus\left\langle G\right\rangle\oplus E^{u} such that

dϕθt(Es(u)(θ))\displaystyle d\phi^{t}_{\theta}(E^{s(u)}(\theta)) =\displaystyle= Es(u)(ϕt(θ)),\displaystyle E^{s(u)}(\phi^{t}(\theta)),
dϕθt|Es\displaystyle\left\|d\phi^{t}_{\theta}\big{|}_{E^{s}}\right\| \displaystyle\leq Cλt,\displaystyle C\lambda^{t},
dϕθt|Eu\displaystyle\left\|d\phi^{-t}_{\theta}\big{|}_{E^{u}}\right\| \displaystyle\leq Cλt,\displaystyle C\lambda^{t},

for all t0t\geq 0 with C>0C>0 and λ(0,1)\lambda\in(0,1), where GG is the geodesic vector field. It’s known that if the geodesic flow is Anosov, then the subspaces Es(θ)E^{s}(\theta) and Eu(θ)E^{u}(\theta) are Lagrangian for every θSM\theta\in SM (see [15] for more details).

2.2. Jacobi fields

To study the differential of the geodesic flow with geometric arguments, let us recall the definition of a Jacobi field. A vector field JJ along a geodesic γ\gamma of MM is a Jacobi field if it satisfies the Jacobi equation

(5) J′′(t)+R(γ(t),J(t))γ(t)=0,\displaystyle J^{\prime\prime}(t)+R(\gamma^{\prime}(t),J(t))\gamma^{\prime}(t)=0,

where RR denotes the curvature tensor of MM and """^{\prime}" denotes the covariant derivative along γ\gamma. A Jacobi field is determined by the initial values J(t0)J(t_{0}) and J(t0)J^{\prime}(t_{0}), for any given t0t_{0}\in\mathbb{R}. If we denote by SS the orthogonal complement of the subspace spanned by GG, for every θSM\theta\in SM, the map ξJξ\xi\rightarrow J_{\xi} defines an isomorphism between S(θ)S(\theta) and the space of perpendicular Jacobi fields along γθ\gamma_{\theta}, where Jξ(0)=dπθ(ξ)J_{\xi}(0)=d\pi_{\theta}(\xi) and Jξ(0)=Kθ(ξ)J^{\prime}_{\xi}(0)=K_{\theta}(\xi). The differential of the geodesic flow is determined by the behavior of the Jacobi fields and, therefore, by the curvature. More precisely, for θSM\theta\in SM and ξTθSM\xi\in T_{\theta}SM we have (in the horizontal and vertical coordinates)

dϕθt(ξ)=(Jξ(t),Jξ(t)),t.\displaystyle d\phi^{t}_{\theta}(\xi)=(J_{\xi}(t),J^{\prime}_{\xi}(t)),\hskip 14.22636ptt\in\mathbb{R}.

In the context of an Anosov geodesic flow, if ξEs(θ)\xi\in E^{s}(\theta) (respectively, ξEu(θ)\xi\in E^{u}(\theta)), the Jacobi field associated Jξ(t)J_{\xi}(t) is called a stable (respectively, unstable) Jacobi field along γθ(t)\gamma_{\theta}(t).
The following proposition allows us to uniformly limit the derivative of the exponential map from certain conditions on the curvature of the manifold.

Proposition 2.2.

Let NN be a complete Riemannian manifold and suppose that the curvature tensor is uniformly bounded. Then there exists t0>0t_{0}>0 such that for all xNx\in N and for all v,wTxNv,w\in T_{x}N with v=w=1\left\|v\right\|=\left\|w\right\|=1 we have

d(expx)tvw52,|t|t0.\displaystyle\left\|d(exp_{x})_{tv}w\right\|\leq\dfrac{5}{2},\hskip 14.22636pt\forall\left|t\right|\leq t_{0}.
Proof.

If wvw\in\left\langle v\right\rangle, then w=vw=v or w=vw=-v. In both cases, by Gauss Lemma (see [8]) we have that

d(expx)tvw2\displaystyle\left\|d(exp_{x})_{tv}w\right\|^{2} =d(expx)tvw,d(expx)tvw\displaystyle=\left\langle d(exp_{x})_{tv}w,d(exp_{x})_{tv}w\right\rangle
=d(expx)tvv,d(expx)tvv\displaystyle=\left\langle d(exp_{x})_{tv}v,d(exp_{x})_{tv}v\right\rangle
=1t2d(expx)tvtv,d(expx)tvtv\displaystyle=\dfrac{1}{t^{2}}\left\langle d(exp_{x})_{tv}tv,d(exp_{x})_{tv}tv\right\rangle
=1t2tv,tv\displaystyle=\dfrac{1}{t^{2}}\left\langle tv,tv\right\rangle
=1.\displaystyle=1.

Now assume that wvw\in\left\langle v\right\rangle^{\perp}. Consider the Jacobi field

J(t)=d(expx)tvtw,t[1,1]\displaystyle J(t)=d(exp_{x})_{tv}tw,\hskip 8.5359ptt\in[-1,1]

with initial conditions J(0)=0J(0)=0 and J(0)=wJ^{\prime}(0)=w. By Lemma 8.3 of [3] there exists t0>0t_{0}>0, independent of the point xx, such that

d(expx)tvw=J(t)|t|32,t(t0,t0){0}.\displaystyle\left\|d(exp_{x})_{tv}w\right\|=\dfrac{\left\|J(t)\right\|}{\left|t\right|}\leq\dfrac{3}{2},\hskip 14.22636pt\forall t\in(-t_{0},t_{0})\setminus\left\{0\right\}.

As TxN=v+vT_{x}N=\left\langle v\right\rangle+\left\langle v\right\rangle^{\perp}, the last inequality completes the proof. ∎

2.3. No conjugate points

Let γ\gamma be a geodesic joining p,qMp,q\in M, pqp\neq q. We say that p,qp,q are conjugate along γ\gamma if there exists a non-zero Jacobi field along γ\gamma vanishing at pp and qq. A manifold MM has no conjugate points if any pair of points are not conjugate. This is equivalent to the fact that the exponential map is non-singular at every point of MM. There are examples of manifolds without conjugate points obtained from the hyperbolic behavior of the geodesic flow. In [6], Klingenberg proved that a compact Riemannian manifold with Anosov geodesic flow has no conjugate points. Years later, Mañé (see [10]) generalized this result to the case of manifolds of finite volume. In the case of infinite volume, Melo and Romaña in [12] extended the result of Mañé over the assumption of sectional curvature bounded below and above. These results show the relationship that exists between the geometry and dynamic of an Anosov geodesic flow.

Let MM be a complete Riemannian manifold without conjugate points and sectional curvature bounded below by c2-c^{2}, for some c>0c>0. When the geodesic flow ϕt:SMSM\phi^{t}:SM\rightarrow SM is of Anosov type, Bolton in [2] showed that there is a positive constant δ\delta such that, for every θSM\theta\in SM, the angle between Es(θ)E^{s}(\theta) and Eu(θ)E^{u}(\theta) is greater than δ\delta. Moreover, Eberlein in [4] showed that

  • 1.

    Kθ(ξ)cdπθ(ξ)\left\|K_{\theta}(\xi)\right\|\leq c\left\|d\pi_{\theta}(\xi)\right\| for every ξEs(θ)\xi\in E^{s}(\theta) or Eu(θ)E^{u}(\theta), where K:TTMTMK:TTM\rightarrow TM is the connection map.

  • 2.

    If ξEs(θ)\xi\in E^{s}(\theta) or Eu(θ)E^{u}(\theta), then Jξ(t)0J_{\xi}(t)\neq 0 for every tt\in\mathbb{R}.

2.4. Lyapunov exponents

Let (M.g)(M.g) be a Riemannian manifold and f:MMf:M\rightarrow M a C1C^{1}-diffeomorphism. The point xx is said to be (Lyapunov-Perron) regular if there exist numbers {𝒳i(x)}i=1l(x)\left\{\mathcal{X}_{i}(x)\right\}_{i=1}^{l(x)}, called Lyapunov exponents, and a decomposition of the tangent space at xx into TxM=i=1l(x)HiT_{x}M=\bigoplus_{i=1}^{l(x)}H_{i} such that for every vector vHi(x){0}v\in H_{i}(x)\setminus\left\{0\right\}, we have

limn±1nlogdfxnv=𝒳i(x)\displaystyle\lim_{n\rightarrow\pm\infty}\dfrac{1}{n}\log\left\|df^{n}_{x}v\right\|=\mathcal{X}_{i}(x)

and

limn±1nlog|det(dfxn)|=i=1l(x)𝒳i(x)dim(Hi(x)).\displaystyle\lim_{n\rightarrow\pm\infty}\dfrac{1}{n}\log\left|\det\left(df^{n}_{x}\right)\right|=\sum_{i=1}^{l(x)}\mathcal{X}_{i}(x)\cdot\dim(H_{i}(x)).

Let Λ\Lambda be the set of regular points. By Oseledec’s Theorem (see [14]), if μ\mu is an ff-invariant probability measure on MM such that log+df±1\log^{+}\left\|df^{\pm 1}\right\| is μ\mu-integrable, then the set Λ\Lambda has full μ\mu-measure. Moreover, the functions x𝒳i(x)x\rightarrow\mathcal{X}_{i}(x) and xdim(Hi(x))x\rightarrow\dim(H_{i}(x)) are μ\mu-measurable and ff-invariant. In particular, if μ\mu is ergodic, they are μ\mu-almost everywhere constant.

3. Existence of Lyapunov exponents

In this section, we will prove that when the geodesic flow is Anosov and the sectional curvature is bounded below, the norm dϕθ±1\left\|d\phi^{\pm 1}_{\theta}\right\| is bounded by a positive constant independent of θ\theta. This boundedness is crucial as it ensures, for a given probability measure, the existence of Lyapunov exponents by Oseledec’s Theorem. More precisely,

Theorem 3.1.

Let MM be a complete Riemannian manifold without conjugate points, sectional curvature bounded below by c2-c^{2}, for some c>0c>0, and μ\mu an ϕt\phi^{t}-invariant probability measure in SMSM. If the geodesic flow is of Anosov type, then logdϕ±1L1(μ)\log\left\|d\phi^{\pm 1}\right\|\in L^{1}(\mu).

Before giving a proof of Theorem 3.1, it is essential to establish the following two lemmas.

Lemma 3.2.

Let MM be a complete Riemannian manifold without conjugate points, sectional curvature bounded below by c2-c^{2}, for some c>0c>0, and geodesic flow of Anosov type. For every θSM\theta\in SM, there exists a constant P>0P>0 such that for every ξEs(θ)\xi\in E^{s}(\theta), ηEu(θ)\eta\in E^{u}(\theta) with ξ=η=1\left\|\xi\right\|=\left\|\eta\right\|=1, we have

Jη(1)P and Jξ(1)P.\displaystyle\left\|J_{\eta}(1)\right\|\leq P\hskip 5.69046pt\text{ and }\hskip 5.69046pt\left\|J_{\xi}(-1)\right\|\leq P.
Proof.

Fix θSM\theta\in SM and let ηEu(θ)\eta\in E^{u}(\theta) with η=1\left\|\eta\right\|=1. Consider a stable Jacobi field JsJ_{s} along γθ\gamma_{\theta} such that Jη(0)=Js(0)J_{\eta}(0)=J_{s}(0) and put ω=(Js(0),Js(0))\omega=(J_{s}(0),J^{\prime}_{s}(0)). By item 1 of section 2.3 we have

Js(0)cJs(0)=cJη(0)c\displaystyle\left\|J^{\prime}_{s}(0)\right\|\leq c\left\|J_{s}(0)\right\|=c\left\|J_{\eta}(0)\right\|\leq c

and

ω2=Js(0)2+Js(0)21+c2.\displaystyle\left\|\omega\right\|^{2}=\left\|J_{s}(0)\right\|^{2}+\left\|J_{s}^{\prime}(0)\right\|^{2}\leq 1+c^{2}.

Define the Jacobi field J(t)=Jη(t)Js(t)J(t)=J_{\eta}(t)-J_{s}(t). We can see that JJ is a perpendicular Jacobi field along γθ\gamma_{\theta} satisfying J(0)=0J(0)=0. By Rauch’s comparison Theorem (see [8]) we have that

(6) J(1)sinhccJ(0).\displaystyle\left\|J(1)\right\|\leq\dfrac{\sinh c}{c}\left\|J^{\prime}(0)\right\|.

Since the geodesic flow is Anosov and ωEs(θ)\omega\in E^{s}(\theta),

(7) Js(1)dϕθ1(ω)CλωCλ1+c2.\displaystyle\left\|J_{s}(1)\right\|\leq\left\|d\phi^{1}_{\theta}(\omega)\right\|\leq C\lambda\left\|\omega\right\|\leq C\lambda\sqrt{1+c^{2}}.

From (6) and (7) we have that

Jη(1)\displaystyle\left\|J_{\eta}(1)\right\| J(1)+Js(1)\displaystyle\leq\left\|J(1)\right\|+\left\|J_{s}(1)\right\|
sinhccJ(0)+Cλ1+c2\displaystyle\leq\dfrac{\sinh c}{c}\left\|J^{\prime}(0)\right\|+C\lambda\sqrt{1+c^{2}}
sinhcc(Jη(0)+Js(0))+Cλ1+c2\displaystyle\leq\dfrac{\sinh c}{c}\left(\left\|J^{\prime}_{\eta}(0)\right\|+\left\|J^{\prime}_{s}(0)\right\|\right)+C\lambda\sqrt{1+c^{2}}
(1+cc)sinhc+Cλ1+c2:=P1.\displaystyle\leq\left(\dfrac{1+c}{c}\right)\sinh c+C\lambda\sqrt{1+c^{2}}:=P_{1}.

Using the same technique for the stable case, there exists P2>0P_{2}>0 such that

Jξ(1)P2\displaystyle\left\|J_{\xi}(-1)\right\|\leq P_{2}

for every ξEs(θ)\xi\in E^{s}(\theta) with ξ=1\left\|\xi\right\|=1. Considering P=max{P1,P2}P=\max\left\{P_{1},P_{2}\right\}, the conclusion of the lemma follows. ∎

We know that, with the hypothesis of Theorem 3.1, there exists a constant δ>0\delta>0 such that the angle between the stable and unstable subspaces is uniformly bounded below by δ\delta. As a direct consequence of this result, we have the following lemma.

Lemma 3.3.

Let MM be a complete Riemannian manifold without conjugate points, sectional curvature bounded below by c2-c^{2}, for some c>0c>0, and geodesic flow of Anosov type. Define the function f:SMf:SM\rightarrow\mathbb{R} as

f(θ)=sup{|ξ,η|:ξEs(θ),ηEu(θ),ξ=η=1}.\displaystyle f(\theta)=\sup\left\{\left|\left\langle\xi,\eta\right\rangle\right|:\xi\in E^{s}(\theta),\eta\in E^{u}(\theta),\left\|\xi\right\|=\left\|\eta\right\|=1\right\}.

Then there exists Q>0Q>0 such that

supθSMf(θ)Q<1.\displaystyle\sup_{\theta\in SM}f(\theta)\leq Q<1.

Proof of Theorem 3.1. Fix θSM\theta\in SM and consider ξTθSM\xi\in T_{\theta}SM with ξ=1\left\|\xi\right\|=1. Since the geodesic flow is of Anosov type, we can write

ξ=sξ1+rξ2+ξ3,\displaystyle\xi=s\xi_{1}+r\xi_{2}+\xi_{3},

where ξ1Es(θ)\xi_{1}\in E^{s}(\theta), ξ2Eu(θ)\xi_{2}\in E^{u}(\theta) and ξ3G(θ)\xi_{3}\in\left\langle G(\theta)\right\rangle with ξ1=ξ2=1\left\|\xi_{1}\right\|=\left\|\xi_{2}\right\|=1. Then

1=sξ1+rξ22+ξ32.\displaystyle 1=\left\|s\xi_{1}+r\xi_{2}\right\|^{2}+\left\|\xi_{3}\right\|^{2}.

This implies that ξ31\left\|\xi_{3}\right\|\leq 1 and sξ1+rξ21\left\|s\xi_{1}+r\xi_{2}\right\|\leq 1. We have

sξ1+rξ22=s2+r2+2srξ1,ξ21.\displaystyle\left\|s\xi_{1}+r\xi_{2}\right\|^{2}=s^{2}+r^{2}+2sr\left\langle\xi_{1},\xi_{2}\right\rangle\leq 1.

It follows from Lemma 3.3 that the regions

Eβ={(s,r):s2+r2+2srβ1}\displaystyle E_{\beta}=\left\{(s,r):s^{2}+r^{2}+2sr\beta\leq 1\right\}

with QβQ-Q\leq\beta\leq Q are bounded ellipses. If we consider L=diam(EQ)2+1>0L=\frac{\text{diam}(E_{Q})}{2}+1>0, the ball BB centered in 0 and radius LL contains these ellipses (see Figure 1). In particular, the parameters s,rs,r are bounded, that is |s|,|r|L\left|s\right|,\left|r\right|\leq L. By Lemma 3.2 we have that

dϕθ1(ξ2)\displaystyle\left\|d\phi^{1}_{\theta}(\xi_{2})\right\| =Jξ2(1)2+Jξ2(1)2\displaystyle=\sqrt{\left\|J_{\xi_{2}}(1)\right\|^{2}+\left\|J^{\prime}_{\xi_{2}}(1)\right\|^{2}}
1+c2Jξ2(1)\displaystyle\leq\sqrt{1+c^{2}}\left\|J_{\xi_{2}}(1)\right\|
1+c2P.\displaystyle\leq\sqrt{1+c^{2}}P.
Refer to caption
Figure 1. Bounded ellipses for Q<1Q<1.

Then

dϕθ1(ξ)\displaystyle\left\|d\phi^{1}_{\theta}(\xi)\right\| |s|dϕθ1(ξ1)+|r|dϕθ1(ξ2)+dϕθ1(ξ3)\displaystyle\leq\left|s\right|\left\|d\phi^{1}_{\theta}(\xi_{1})\right\|+\left|r\right|\left\|d\phi^{1}_{\theta}(\xi_{2})\right\|+\left\|d\phi^{1}_{\theta}(\xi_{3})\right\|
|s|Cλ+|r|1+c2P+1\displaystyle\leq\left|s\right|C\lambda+\left|r\right|\sqrt{1+c^{2}}P+1
(8) LCλ+L1+c2P+1\displaystyle\leq LC\lambda+L\sqrt{1+c^{2}}P+1

for every ξTθSM\xi\in T_{\theta}SM with ξ=1\left\|\xi\right\|=1. This implies that dϕθ1\left\|d\phi^{1}_{\theta}\right\| is bounded and therefore the function logdϕ1\log\left\|d\phi^{1}\right\| is μ\mu-integrable, since the constants LL and PP are independent of the point θ\theta. Using the second inequality of Lemma 3.2 we obtain that logdϕ1\log\left\|d\phi^{-1}\right\| is μ\mu-integrable.\hfill\square

4. Consequences of a geodesic flow being of Anosov type

In this section, we explore some results, based on the hyperbolicity of a geodesic flow, that will allow us to address the challenge of the non-compactness of the manifold in the proof of Ruelle’s inequality.

From now on, let us assume that MM is a complete Riemannian manifold without conjugate points, sectional curvature bounded below by c2-c^{2}, for some c>0c>0, and the geodesic flow ϕt:SMSM\phi^{t}:SM\rightarrow SM is of Anosov type. For every ωTθSM\omega\in T_{\theta}SM we can write

ω=ωs+ωu+ωc,\displaystyle\omega=\omega^{s}+\omega^{u}+\omega^{c},

where ωsEs(θ)\omega^{s}\in E^{s}(\theta), ωuEu(θ)\omega^{u}\in E^{u}(\theta) and ωcG(θ)\omega^{c}\in\left\langle G(\theta)\right\rangle.

Lemma 4.1.

For mm\in\mathbb{N} large enough, there is τ1>1\tau_{1}>1 such that for every θSM\theta\in SM

dϕθmτ1dϕθm(η)\displaystyle\left\|d\phi^{m}_{\theta}\right\|\leq\tau_{1}\left\|d\phi^{m}_{\theta}(\eta)\right\|

for some ηEu(θ)\eta\in E^{u}(\theta) with η=1\left\|\eta\right\|=1.

Proof.

Fix θSM\theta\in SM and let ω=ωs+ωu+ωcTθSM\omega=\omega^{s}+\omega^{u}+\omega^{c}\in T_{\theta}SM with ω=1\left\|\omega\right\|=1. This implies that ωs+ωu1\left\|\omega^{s}+\omega^{u}\right\|\leq 1 and ωc1\left\|\omega^{c}\right\|\leq 1. Moreover, we know that ωsL\left\|\omega^{s}\right\|\leq L and ωuL\left\|\omega^{u}\right\|\leq L (see Section 3). Consider mm\in\mathbb{N} large enough such that Cλm<1/2C\lambda^{m}<1/2.
Case 1: ωu=0\omega^{u}=0.
Since the geodesic flow is Anosov we have that

dϕθm(ω)\displaystyle\left\|d\phi^{m}_{\theta}(\omega)\right\| dϕθm(ωs)+dϕθm(ωc)\displaystyle\leq\left\|d\phi^{m}_{\theta}(\omega^{s})\right\|+\left\|d\phi^{m}_{\theta}(\omega^{c})\right\|
Cλm+1\displaystyle\leq C\lambda^{m}+1
<C1λm\displaystyle<C^{-1}\lambda^{-m}
dϕθm(η)\displaystyle\leq\left\|d\phi^{m}_{\theta}(\eta)\right\|

for every ηEu(θ)\eta\in E^{u}(\theta) with η=1\left\|\eta\right\|=1.
Case 2: ωu0\omega^{u}\neq 0.
Since the geodesic flow is Anosov we have that

dϕθm(ωs)CλmL<C1λmLLdϕθm(ωu)ωu.\displaystyle\left\|d\phi^{m}_{\theta}(\omega^{s})\right\|\leq C\lambda^{m}L<C^{-1}\lambda^{-m}L\leq L\dfrac{\left\|d\phi^{m}_{\theta}(\omega^{u})\right\|}{\left\|\omega^{u}\right\|}.

Then

dϕθm(ω)\displaystyle\left\|d\phi^{m}_{\theta}(\omega)\right\| dϕθm(ωs)+dϕθm(ωu)+dϕθm(ωc)\displaystyle\leq\left\|d\phi^{m}_{\theta}(\omega^{s})\right\|+\left\|d\phi^{m}_{\theta}(\omega^{u})\right\|+\left\|d\phi^{m}_{\theta}(\omega^{c})\right\|
Ldϕθm(ωu)ωu+Ldϕθm(ωu)ωu+1\displaystyle\leq L\dfrac{\left\|d\phi^{m}_{\theta}(\omega^{u})\right\|}{\left\|\omega^{u}\right\|}+L\dfrac{\left\|d\phi^{m}_{\theta}(\omega^{u})\right\|}{\left\|\omega^{u}\right\|}+1
<(2L+1)dϕθm(ωu)ωu.\displaystyle<(2L+1)\dfrac{\left\|d\phi^{m}_{\theta}(\omega^{u})\right\|}{\left\|\omega^{u}\right\|}.

If we consider τ1=2L+1\tau_{1}=2L+1, in both cases we have that

dϕθm(ω)τ1dϕθm|Eu(θ)\displaystyle\left\|d\phi^{m}_{\theta}(\omega)\right\|\leq\tau_{1}\left\|\left.d\phi^{m}_{\theta}\right|_{E^{u}(\theta)}\right\|

for every ωTθSM\omega\in T_{\theta}SM with ω=1\left\|\omega\right\|=1. Since the norm is always attained in a finite-dimensional space, we conclude the proof of the lemma. ∎

Lemma 4.2.

For mm\in\mathbb{N} large enough, there is τ2(0,1)\tau_{2}\in(0,1) such that for every θSM\theta\in SM

dϕθmτ2dϕθm(ξ)\displaystyle\left\|d\phi^{m}_{\theta}\right\|^{*}\geq\tau_{2}\left\|d\phi^{m}_{\theta}(\xi)\right\|

for some ξEs(θ)\xi\in E^{s}(\theta) with ξ=1\left\|\xi\right\|=1, where dϕθm=infv=1dϕθm(v)\displaystyle\left\|d\phi^{m}_{\theta}\right\|^{*}=\inf_{\left\|v\right\|=1}\left\|d\phi^{m}_{\theta}(v)\right\|.

Proof.

Let ε>0\varepsilon>0 and consider mm\in\mathbb{N} large enough such that ε(L+1)Cλm\varepsilon\geq(L+1)C\lambda^{m} and 1ε2>εCλm\sqrt{1-\varepsilon^{2}}>\varepsilon C\lambda^{m}, where LL comes from Section 3. Fix θSM\theta\in SM and define the following set

Γθ,ε,m:={ωTθSM:ω=1,ω=ωs+ωu+ωcanddϕθm(ωu+ωc)ε}.\displaystyle\Gamma_{\theta,\varepsilon,m}:=\left\{\omega\in T_{\theta}SM:\left\|\omega\right\|=1,\omega=\omega^{s}+\omega^{u}+\omega^{c}\hskip 5.69046pt\text{and}\hskip 5.69046pt\left\|d\phi^{m}_{\theta}(\omega^{u}+\omega^{c})\right\|\geq\varepsilon\right\}.

Case 1: ωΓθ,ε,m\omega\in\Gamma_{\theta,\varepsilon,m} with ωs0\omega^{s}\neq 0.
Since the geodesic flow is Anosov,

dϕθm(ω)\displaystyle\left\|d\phi^{m}_{\theta}(\omega)\right\| dϕθm(ωu+ωc)dϕθm(ωs)\displaystyle\geq\left\|d\phi^{m}_{\theta}(\omega^{u}+\omega^{c})\right\|-\left\|d\phi^{m}_{\theta}(\omega^{s})\right\|
(9) dϕθm(ωu+ωc)Cλmωs.\displaystyle\geq\left\|d\phi^{m}_{\theta}(\omega^{u}+\omega^{c})\right\|-C\lambda^{m}\left\|\omega^{s}\right\|.

As ωΓθ,ε,m\omega\in\Gamma_{\theta,\varepsilon,m} we have that

(10) dϕθm(ωu+ωc)ε(L+1)Cλm(ωs+1)Cλm.\displaystyle\left\|d\phi^{m}_{\theta}(\omega^{u}+\omega^{c})\right\|\geq\varepsilon\geq(L+1)C\lambda^{m}\geq(\left\|\omega^{s}\right\|+1)C\lambda^{m}.

Then from (4) and (10)

dϕθm(ω)dϕθm(ωu+ωc)CλmωsCλmdϕθm(ωs)ωs.\displaystyle\left\|d\phi^{m}_{\theta}(\omega)\right\|\geq\left\|d\phi^{m}_{\theta}(\omega^{u}+\omega^{c})\right\|-C\lambda^{m}\left\|\omega^{s}\right\|\geq C\lambda^{m}\geq\dfrac{\left\|d\phi^{m}_{\theta}(\omega^{s})\right\|}{\left\|\omega^{s}\right\|}.

Case 2: ωΓθ,ε,m\omega\in\Gamma_{\theta,\varepsilon,m} with ωs=0\omega^{s}=0.
Since the geodesic flow is Anosov,

dϕθm(ω)=dϕθm(ωu+ωc)ε>Cλmdϕθm(ξ)\displaystyle\left\|d\phi^{m}_{\theta}(\omega)\right\|=\left\|d\phi^{m}_{\theta}(\omega^{u}+\omega^{c})\right\|\geq\varepsilon>C\lambda^{m}\geq\left\|d\phi^{m}_{\theta}(\xi)\right\|

for every ξEs(θ)\xi\in E^{s}(\theta) with ξ=1\left\|\xi\right\|=1.
Case 3: ωΓθ,ε,m\omega\notin\Gamma_{\theta,\varepsilon,m} and w=1\left\|w\right\|=1.
We have that

ε2\displaystyle\varepsilon^{2} >dϕθm(ωu+ωc)2=dϕθm(ωu)2+dϕθm(ωc)2=dϕθm(ωu)2+ωc2.\displaystyle>\left\|d\phi^{m}_{\theta}(\omega^{u}+\omega^{c})\right\|^{2}=\left\|d\phi^{m}_{\theta}(\omega^{u})\right\|^{2}+\left\|d\phi^{m}_{\theta}(\omega^{c})\right\|^{2}=\left\|d\phi^{m}_{\theta}(\omega^{u})\right\|^{2}+\left\|\omega^{c}\right\|^{2}.

Then ωc<ε\left\|\omega^{c}\right\|<\varepsilon and dϕθm(ωu)<ε\left\|d\phi^{m}_{\theta}(\omega^{u})\right\|<\varepsilon. Since the geodesic flow is Anosov,

C1λmωudϕθm(ωu)<ε.\displaystyle C^{-1}\lambda^{-m}\left\|\omega^{u}\right\|\leq\left\|d\phi^{m}_{\theta}(\omega^{u})\right\|<\varepsilon.

This implies that ωu<εCλm\left\|\omega^{u}\right\|<\varepsilon C\lambda^{m}. On the other hand, as w=1\left\|w\right\|=1, then

ωu+ωsωu+ωs=1ωc2>1ε2.\displaystyle\left\|\omega^{u}\right\|+\left\|\omega^{s}\right\|\geq\left\|\omega^{u}+\omega^{s}\right\|=\sqrt{1-\left\|\omega^{c}\right\|^{2}}>\sqrt{1-\varepsilon^{2}}.

Furthermore

(11) Lωs>1ε2εCλm>0.\displaystyle L\geq\left\|\omega^{s}\right\|>\sqrt{1-\varepsilon^{2}}-\varepsilon C\lambda^{m}>0.

In particular, ωs0\omega^{s}\neq 0. Denote by

Ecu(θ):=Eu(θ)G(θ)\displaystyle E^{cu}(\theta):=E^{u}(\theta)\oplus\left\langle G(\theta)\right\rangle

and define the following linear map

Pθ:TθSMEs(θ)\displaystyle P_{\theta}:T_{\theta}SM\rightarrow E^{s}(\theta)

as the parallel projection onto Es(θ)E^{s}(\theta) along Ecu(θ)E^{cu}(\theta). Since the angle between the stable and unstable subspaces is uniformly away from 0 for every θSM\theta\in SM, then there is δ1\delta\geq 1 such that

Pθ(ω)δω\displaystyle\left\|P_{\theta}(\omega)\right\|\leq\delta\left\|\omega\right\|

for every θSM\theta\in SM and ωTθSM\omega\in T_{\theta}SM (see Theorem 3.1 in [5]). Then

dϕθm(ωs)=Pϕm(θ)(dϕθm(ω))δdϕθm(ω).\displaystyle\left\|d\phi^{m}_{\theta}(\omega^{s})\right\|=\left\|P_{\phi^{m}(\theta)}(d\phi^{m}_{\theta}(\omega))\right\|\leq\delta\left\|d\phi^{m}_{\theta}(\omega)\right\|.

By (11), if we choose ε>0\varepsilon>0 such that ωs1/2\left\|\omega^{s}\right\|\geq 1/2, we have that

dϕθm(ω)12δdϕθm(ωs)ωs.\displaystyle\left\|d\phi^{m}_{\theta}(\omega)\right\|\geq\dfrac{1}{2\delta}\dfrac{\left\|d\phi^{m}_{\theta}(\omega^{s})\right\|}{\left\|\omega^{s}\right\|}.

If we consider τ2=1/2δ\tau_{2}=1/2\delta, in all cases we have that for all for every ωTθSM\omega\in T_{\theta}SM with ω=1\left\|\omega\right\|=1 there is ξEs(θ)\xi\in E^{s}(\theta), with ξ=1\left\|\xi\right\|=1 such that

dθϕm(ω)τ2dθϕm(ξ).\left\|d_{\theta}\phi^{m}(\omega)\right\|\geq\tau_{2}\left\|d_{\theta}\phi^{m}(\xi)\right\|.

Since the infimum is always attained in a finite-dimensional space, the last inequality concludes the proof of the lemma. ∎

Clearly dϕθmdϕθm\left\|d\phi^{m}_{\theta}\right\|^{*}\leq\left\|d\phi^{m}_{\theta}\right\| for every θSM\theta\in SM. From Lemmas 4.1 and 4.2, we can obtain a positive constant, independent of θ\theta, such that the direction of the inequality changes.

Proposition 4.3.

For mm\in\mathbb{N} large enough, there is κ>1\kappa>1, depending on mm, such that

dϕθmκdϕθm\displaystyle\left\|d\phi^{m}_{\theta}\right\|\leq\kappa\left\|d\phi^{m}_{\theta}\right\|^{*}

for every θSM\theta\in SM.

Proof.

From Lemma 4.1 we have that

dϕθmτ1dϕθm(η)\displaystyle\left\|d\phi^{m}_{\theta}\right\|\leq\tau_{1}\left\|d\phi^{m}_{\theta}(\eta)\right\|

for some ηEu(θ)\eta\in E^{u}(\theta) with η=1\left\|\eta\right\|=1. Denote by JηJ_{\eta} the Jacobi field associated to η\eta. Since the geodesic flow is of Anosov type and dϕθt(η)=(Jη(t),Jη(t))d\phi^{t}_{\theta}(\eta)=(J_{\eta}(t),J^{\prime}_{\eta}(t)), we have from item 1 of Section 2.3 that

(12) dϕθmτ1dϕθm(η)=τ1Jη(m)2+Jη(m)2τ11+c2Jη(m).\displaystyle\left\|d\phi^{m}_{\theta}\right\|\leq\tau_{1}\left\|d\phi^{m}_{\theta}(\eta)\right\|=\tau_{1}\sqrt{\left\|J_{\eta}(m)\right\|^{2}+\left\|J^{\prime}_{\eta}(m)\right\|^{2}}\leq\tau_{1}\sqrt{1+c^{2}}\left\|J_{\eta}(m)\right\|.

In the same way, by Lemma 4.2 we have that

(13) dϕθmτ2dϕθm(ξ)=τ21+Jξ(m)2Jξ(m)2Jξ(m),\displaystyle\left\|d\phi^{m}_{\theta}\right\|^{*}\geq\tau_{2}\left\|d\phi^{m}_{\theta}(\xi)\right\|=\tau_{2}\sqrt{1+\dfrac{\left\|J^{\prime}_{\xi}(m)\right\|^{2}}{\left\|J_{\xi}(m)\right\|^{2}}}\left\|J_{\xi}(m)\right\|,

for some ξEs(θ)\xi\in E^{s}(\theta) with ξ=1\left\|\xi\right\|=1, where JξJ_{\xi} is the Jacobi field associated to ξ\xi. Moreover,

(14) 1+c21+c21+Jξ(m)2Jξ(m)2.\displaystyle\sqrt{1+c^{2}}\leq\sqrt{1+c^{2}}\sqrt{1+\dfrac{\left\|J^{\prime}_{\xi}(m)\right\|^{2}}{\left\|J_{\xi}(m)\right\|^{2}}}.

Define the function

r:\displaystyle r: [0,+)\displaystyle[0,+\infty)\rightarrow\hskip 11.38092pt\mathbb{R}
tλtJξ(t)λtJη(t).\displaystyle\hskip 17.07182ptt\hskip 18.49411pt\rightarrow\dfrac{\lambda^{-t}\left\|J_{\xi}(t)\right\|}{\lambda^{t}\left\|J_{\eta}(t)\right\|}.

This function is well-defined because the stable and unstable Jacobi fields are never zero since the manifold has no conjugate points (see Section 2). We have that

r(t)=r(t)(2logλ+Jξ(t),Jξ(t)Jξ(t),Jξ(t)Jη(t),Jη(t)Jη(t),Jη(t)).\displaystyle r^{\prime}(t)=r(t)\left(-2\log\lambda+\dfrac{\left\langle J^{\prime}_{\xi}(t),J_{\xi}(t)\right\rangle}{\left\langle J_{\xi}(t),J_{\xi}(t)\right\rangle}-\dfrac{\left\langle J^{\prime}_{\eta}(t),J_{\eta}(t)\right\rangle}{\left\langle J_{\eta}(t),J_{\eta}(t)\right\rangle}\right).

Also

A(t)=Jξ(t),Jξ(t)Jξ(t),Jξ(t)[c,c]andB(t)=Jη(t),Jη(t)Jη(t),Jη(t)[c,c].\displaystyle A(t)=\dfrac{\left\langle J^{\prime}_{\xi}(t),J_{\xi}(t)\right\rangle}{\left\langle J_{\xi}(t),J_{\xi}(t)\right\rangle}\in[-c,c]\hskip 11.38092pt\text{and}\hskip 11.38092ptB(t)=\dfrac{\left\langle J^{\prime}_{\eta}(t),J_{\eta}(t)\right\rangle}{\left\langle J_{\eta}(t),J_{\eta}(t)\right\rangle}\in[-c,c].

Since the curvature is bounded below by c2-c^{2}, then λec\lambda\geq e^{-c} (see [13]). Therefore

2logλ2c2logλ+A(t)B(t)2logλ+2c.\displaystyle-2\log\lambda-2c\leq-2\log\lambda+A(t)-B(t)\leq-2\log\lambda+2c.

This implies that

2logλ2cr(t)r(t)2logλ+2c\displaystyle-2\log\lambda-2c\leq\dfrac{r^{\prime}(t)}{r(t)}\leq-2\log\lambda+2c

and

r(0)e(2logλ2c)tr(t)r(0)e(2logλ+2c)t.\displaystyle r(0)\cdot e^{(-2\log\lambda-2c)t}\leq r(t)\leq r(0)\cdot e^{(-2\log\lambda+2c)t}.

Therefore

r(0)1e(2logλ2c)t1r(t)r(0)1e(2logλ+2c)t.\displaystyle r(0)^{-1}\cdot e^{(2\log\lambda-2c)t}\leq\dfrac{1}{r(t)}\leq r(0)^{-1}\cdot e^{(2\log\lambda+2c)t}.

For t=mt=m we have that

(15) Jη(m)r(0)1e(2logλ+2c)mλ2mJξ(m)=r(0)1e2cmJξ(m).\displaystyle\left\|J_{\eta}(m)\right\|\leq r(0)^{-1}\cdot e^{(2\log\lambda+2c)m}\cdot\lambda^{-2m}\left\|J_{\xi}(m)\right\|=r(0)^{-1}\cdot e^{2cm}\left\|J_{\xi}(m)\right\|.

From (12), (13), (14) and (15)

dϕθm\displaystyle\left\|d\phi^{m}_{\theta}\right\| τ11+c2Jη(m)\displaystyle\leq\tau_{1}\sqrt{1+c^{2}}\left\|J_{\eta}(m)\right\|
τ11+c2r(0)1e2cmJξ(m)\displaystyle\leq\tau_{1}\sqrt{1+c^{2}}\cdot r(0)^{-1}\cdot e^{2cm}\left\|J_{\xi}(m)\right\|
(16) τ11+c21+Jξ(m)2Jξ(m)2r(0)1e2cmJξ(m).\displaystyle\leq\tau_{1}\sqrt{1+c^{2}}\sqrt{1+\dfrac{\left\|J^{\prime}_{\xi}(m)\right\|^{2}}{\left\|J_{\xi}(m)\right\|^{2}}}\cdot r(0)^{-1}\cdot e^{2cm}\left\|J_{\xi}(m)\right\|.

From item 1 of Section 2.3 we have that

1=ξ2=dπθ(ξ)2+Kθ(ξ)2(1+c2)dπθ(ξ)2.\displaystyle 1=\left\|\xi\right\|^{2}=\left\|d\pi_{\theta}(\xi)\right\|^{2}+\left\|K_{\theta}(\xi)\right\|^{2}\leq(1+c^{2})\left\|d\pi_{\theta}(\xi)\right\|^{2}.

Since 1=η2=dπθ(η)2+Kθ(η)21=\left\|\eta\right\|^{2}=\left\|d\pi_{\theta}(\eta)\right\|^{2}+\left\|K_{\theta}(\eta)\right\|^{2}, the last inequality implies that

r(0)1=dπθ(η)dπθ(ξ)1dπθ(ξ)1+c2.\displaystyle r(0)^{-1}=\dfrac{\left\|d\pi_{\theta}(\eta)\right\|}{\left\|d\pi_{\theta}(\xi)\right\|}\leq\dfrac{1}{\left\|d\pi_{\theta}(\xi)\right\|}\leq\sqrt{1+c^{2}}.

Therefore, substituting in (4) and using (13)

dϕθmκdϕθm,\displaystyle\left\|d\phi^{m}_{\theta}\right\|\leq\kappa\left\|d\phi^{m}_{\theta}\right\|^{*},

where κ=τ1τ21(1+c2)e2cm>1\kappa=\tau_{1}\cdot\tau_{2}^{-1}\cdot(1+c^{2})\cdot e^{2cm}>1. ∎

On the other hand, since the geodesic flow is of Anosov type, we have that the norm dϕθm\left\|d\phi^{m}_{\theta}\right\| is bounded between two positive constants.

Proposition 4.4.

For mm\in\mathbb{N} large enough, there are constants K1,K2>0K_{1},K_{2}>0, K1K_{1} depending on mm, such that

K2<dϕθm<K1\displaystyle K_{2}<\left\|d\phi^{m}_{\theta}\right\|<K_{1}

for every θSM\theta\in SM.

Proof.

Fix θSM\theta\in SM. Since the geodesic flow is of Anosov type, for ηEu(θ)\eta\in E^{u}(\theta) with η=1\left\|\eta\right\|=1 we have that

dϕθmdϕθm(η)C1λm>C1,\displaystyle\left\|d\phi^{m}_{\theta}\right\|\geq\left\|d\phi^{m}_{\theta}(\eta)\right\|\geq C^{-1}\lambda^{-m}>C^{-1},

then K2=1/CK_{2}=1/C. On the other hand, from (3) we have that

dϕθ1\displaystyle\left\|d\phi^{1}_{\theta}\right\| LCλ+L1+c2((1+cc)sinhc+Cλ1+c2)+1\displaystyle\leq LC\lambda+L\sqrt{1+c^{2}}\left(\left(\dfrac{1+c}{c}\right)\sinh c+C\lambda\sqrt{1+c^{2}}\right)+1
LCλ+L1+c2(1+cc)sinhc+LCλ(1+c2)+1\displaystyle\leq LC\lambda+L\sqrt{1+c^{2}}\left(\dfrac{1+c}{c}\right)\sinh c+LC\lambda(1+c^{2})+1
(17) 2LCλ+LCλc2+L1+c2(1+cc)sinhc+1:=h(c).\displaystyle\leq 2LC\lambda+LC\lambda c^{2}+L\sqrt{1+c^{2}}\left(\dfrac{1+c}{c}\right)\sinh c+1:=h(c).

Then, we can consider K1=h(c)mK_{1}=h(c)^{m}. ∎

A direct consequence of Proposition 4.4 is the following result.

Corollary 4.5.

Given ε>0\varepsilon>0, there is β(0,1)\beta\in(0,1), depending on mm, such that

βdϕθ~m<dϕθm,θ~SM:d(θ,θ~)<ε\displaystyle\beta\left\|d\phi^{m}_{\tilde{\theta}}\right\|<\left\|d\phi^{m}_{\theta}\right\|,\hskip 11.38092pt\forall\hskip 2.84544pt\tilde{\theta}\in SM:d(\theta,\tilde{\theta})<\varepsilon

for every θSM\theta\in SM.

Proof.

By Proposition 4.4 we have that

K2K1<dϕθmdϕθ~m<K1K2\displaystyle\dfrac{K_{2}}{K_{1}}<\dfrac{\left\|d\phi^{m}_{\theta}\right\|}{\left\|d\phi^{m}_{\tilde{\theta}}\right\|}<\dfrac{K_{1}}{K_{2}}

Considering β=K2K1=C1h(c)m\beta=\dfrac{K_{2}}{K_{1}}=\dfrac{C^{-1}}{h(c)^{m}} the conclusion of the corollary follows. ∎

5. Ruelle’s Inequality

In this section, we will prove Theorem 1.1. For this, we will adapt the idea of the proof of Ruelle’s inequality for diffeomorphisms in the compact case exhibited in [1].

Let MM be a complete Riemannian manifold satisfying all the hypotheses of Theorem 1.1 and μ\mu an ϕt\phi^{t}-invariant probability measure on SMSM. By simplicity, we consider μ\mu an ergodic ϕt\phi^{t}-invariant probability measure on SMSM. In this case, we denote by {𝒳i}\left\{\mathcal{X}_{i}\right\} the Lyapunov exponents and {ki}\left\{k_{i}\right\} their respective multiplicities. The proof in the non-ergodic case is a consequence of the ergodic decomposition of such a measure. We can also assume that ϕ=ϕ1\phi=\phi^{1} is an ergodic transformation with respect to μ\mu. If it is not the case, we can choose an ergodic-time τ\tau for μ\mu and prove the theorem for the map ϕτ\phi^{\tau}. The proof of the theorem for the map ϕτ\phi^{\tau} implies the proof for the map ϕ\phi because the entropy of ϕτ\phi^{\tau} and the Lyapunov exponents are τ\tau-multiples of the respective values of ϕ\phi.

Fix ε>0\varepsilon>0 and mm\in\mathbb{N} large enough. There exists a compact set KSMK\subset SM such that μ(K)>1ε\mu(K)>1-\varepsilon. Based on the results in Section 4, we present the following theorem, which constitutes a similar version to the inclusion (10.3) described in [1]. Consider the constants κ>1\kappa>1 and 0<β<10<\beta<1 given by Proposition 4.3 and Corollary 4.5 respectively.

Theorem 5.1.

Let MM be a complete Riemannian manifold without conjugate points and sectional curvature bounded below by c2-c^{2}, for some c>0c>0. If the geodesic flow is of Anosov type, then for every θK\theta\in K there exists ϱ:=ϱ(K)(0,1)\varrho:=\varrho(K)\in(0,1) such that

ϕm(expθ(B(0,βκ1ϱ)))expϕm(θ)(dϕθm(B(0,ϱ)).\displaystyle\phi^{m}(exp_{\theta}(B(0,\beta\kappa^{-1}\varrho)))\subseteq exp_{\phi^{m}(\theta)}(d\phi^{m}_{\theta}(B(0,\varrho)).
Proof.

We will proceed by contradiction. Suppose that for every nn\in\mathbb{N}, there are θnK\theta_{n}\in K and vnTθnSMv_{n}\in T_{\theta_{n}}SM with vn=βκ1n\left\|v_{n}\right\|=\dfrac{\beta\kappa^{-1}}{n} such that

ϕm(expθn(vn))=expϕm(θn)(dϕθnm(wn))\displaystyle\phi^{m}(exp_{\theta_{n}}(v_{n}))=exp_{\phi^{m}(\theta_{n})}(d\phi^{m}_{\theta_{n}}(w_{n}))

where wn=1n\left\|w_{n}\right\|=\dfrac{1}{n}. Since KK is compact and wn0w_{n}\rightarrow 0, then dϕθnm(wn)\left\|d\phi^{m}_{\theta_{n}}(w_{n})\right\| is less than injectivity radius of the exponential map restricted to the compact set KK, for nn large enough by Proposition 4.4. Therefore

dϕθnm(wn)\displaystyle\left\|d\phi^{m}_{\theta_{n}}(w_{n})\right\| =d(ϕm(θn),expϕm(θn)(dϕθnm(wn)))\displaystyle=d(\phi^{m}(\theta_{n}),exp_{\phi^{m}(\theta_{n})}(d\phi^{m}_{\theta_{n}}(w_{n})))
=d(ϕm(θn),ϕm(expθn(vn)))\displaystyle=d(\phi^{m}(\theta_{n}),\phi^{m}(exp_{\theta_{n}}(v_{n})))
01(ϕmcn)(t)𝑑t,\displaystyle\leq\int_{0}^{1}\left\|(\phi^{m}\circ c_{n})^{\prime}(t)\right\|dt,

where cn(t)=expθn(tvn)c_{n}(t)=exp_{\theta_{n}}(tv_{n}). Then

dϕθnm(wn)\displaystyle\left\|d\phi^{m}_{\theta_{n}}(w_{n})\right\| supt[0,1]dϕcn(t)m01cn(t)𝑑t\displaystyle\leq\sup_{t\in[0,1]}\left\|d\phi^{m}_{c_{n}(t)}\right\|\int_{0}^{1}\left\|c^{\prime}_{n}(t)\right\|dt
=supt[0,1]dϕcn(t)mvn.\displaystyle=\sup_{t\in[0,1]}\left\|d\phi^{m}_{c_{n}(t)}\right\|\cdot\left\|v_{n}\right\|.

For nn large enough, by Corollary 4.5 we have that

κβ1dϕθnm(wn)wn\displaystyle\kappa\beta^{-1}\dfrac{\left\|d\phi^{m}_{\theta_{n}}(w_{n})\right\|}{\left\|w_{n}\right\|} =wnvndϕθnm(wn)wn\displaystyle=\dfrac{\left\|w_{n}\right\|}{\left\|v_{n}\right\|}\dfrac{\left\|d\phi^{m}_{\theta_{n}}(w_{n})\right\|}{\left\|w_{n}\right\|}
supt[0,1]dcn(t)ϕm\displaystyle\leq\sup_{t\in[0,1]}\left\|d_{c_{n}(t)}\phi^{m}\right\|
<β1dϕθnm.\displaystyle<\beta^{-1}\left\|d\phi^{m}_{\theta_{n}}\right\|.

Therefore

κβ1dϕθnm<β1dϕθnm\displaystyle\kappa\beta^{-1}\left\|d\phi^{m}_{\theta_{n}}\right\|^{*}<\beta^{-1}\left\|d\phi^{m}_{\theta_{n}}\right\|

which contradicts the Proposition 4.3. ∎

Now, denote by ϱm=βκ1ϱ<1\varrho_{m}=\beta\kappa^{-1}\varrho<1, where the constants β,κ\beta,\kappa and ϱ\varrho come from Theorem 5.1. Using the techniques of separate sets applied in [1] we define a finite partition 𝒫=𝒫K{SMK}\mathcal{P}=\mathcal{P}_{K}\cup\left\{SM\setminus K\right\} of SMSM in the following way:

  • .

    𝒫K\mathcal{P}_{K} is a partition of KK such that for every X𝒫KX\in\mathcal{P}_{K}, there exist balls B(x,r)B(x,r^{\prime}) and B(x,r)B(x,r) such that the constants satisfy 0<r<r<2rϱm20<r^{\prime}<r<2r^{\prime}\leq\dfrac{\varrho_{m}}{2} and

    B(x,r)XB(x,r).B(x,r^{\prime})\subset X\subset B(x,r).
  • .

    There exists a constant ζ>0\zeta>0 such that the cardinal of 𝒫K\mathcal{P}_{K}, denoted by |𝒫K|\left|\mathcal{P}_{K}\right|, satisfies

    |𝒫K|ζ(ϱm)dim(SM).\left|\mathcal{P}_{K}\right|\leq\zeta\cdot(\varrho_{m})^{-\dim(SM)}.
  • .

    hμ(ϕm,𝒫)hμ(ϕm)εh_{\mu}(\phi^{m},\mathcal{P})\geq h_{\mu}(\phi^{m})-\varepsilon.

By definition of entropy,

hμ(ϕm,𝒫)\displaystyle h_{\mu}(\phi^{m},\mathcal{P}) =limk+Hμ(𝒫|ϕm𝒫ϕkm𝒫)\displaystyle=\lim_{k\to+\infty}H_{\mu}\left(\left.\mathcal{P}\right|\phi^{m}\mathcal{P}\vee\ldots\vee\phi^{km}\mathcal{P}\right)
Hμ(𝒫|ϕm𝒫)\displaystyle\leq H_{\mu}\left(\left.\mathcal{P}\right|\phi^{m}\mathcal{P}\right)
(18) Dϕm𝒫μ(D)logcard{X𝒫:XD}.\displaystyle\leq\sum_{D\in\phi^{m}\mathcal{P}}\mu(D)\cdot\log\text{card}\left\{X\in\mathcal{P}:X\cap D\neq\emptyset\right\}.

Denote by φ=supθSMdϕθ>1\varphi=\sup_{\theta\in SM}\left\|d\phi_{\theta}\right\|>1. First, we estimate the number of elements X𝒫X\in\mathcal{P} that intersect a given element Dϕm𝒫D\in\phi^{m}\mathcal{P}.

Lemma 5.2.

There exists a constant L1>0L_{1}>0 such that if Dϕm𝒫D\in\phi^{m}\mathcal{P} then

card{X𝒫:XD}L1max{φmdim(SM),(ϱm)dim(SM)}.\displaystyle\emph{card}\left\{X\in\mathcal{P}:X\cap D\neq\emptyset\right\}\leq L_{1}\cdot\max\left\{\varphi^{m\cdot\dim(SM)},(\varrho_{m})^{-\dim(SM)}\right\}.
Proof.

Consider Dϕm𝒫D\in\phi^{m}\mathcal{P}, then D=ϕm(X)D=\phi^{m}(X^{\prime}) for some X𝒫X^{\prime}\in\mathcal{P}.
Case I: X𝒫KX^{\prime}\in\mathcal{P}_{K}.
By the mean value inequality

diam(D)\displaystyle\text{diam}(D) =diam(ϕm(X))\displaystyle=\text{diam}(\phi^{m}(X^{\prime}))
supθSMdϕθmdiam(X)\displaystyle\leq\sup_{\theta\in SM}\left\|d\phi_{\theta}\right\|^{m}\cdot\text{diam}(X^{\prime})
φm4r,\displaystyle\leq\varphi^{m}\cdot 4r^{\prime},

since XB(x,2r)X^{\prime}\subset B(x,2r^{\prime}). If X𝒫KX\in\mathcal{P}_{K} satisfies XDX\cap D\neq\emptyset, then XX is contained in a 4r4r^{\prime}-neighborhood of DD, denoted by WW. Since φm>1\varphi^{m}>1 we have that

diam(W)\displaystyle\text{diam}(W) φm4r+8r\displaystyle\leq\varphi^{m}\cdot 4r^{\prime}+8r^{\prime}
=4r(φm+2)\displaystyle=4r^{\prime}\cdot\left(\varphi^{m}+2\right)
<12rφm.\displaystyle<12r^{\prime}\cdot\varphi^{m}.

Hence

(19) {X𝒫K:XD}vol(X)vol(W)A1(r)dim(SM)φmdim(SM),\displaystyle\sum_{\left\{X\in\mathcal{P}_{K}:X\cap D\neq\emptyset\right\}}\text{vol}(X)\leq\text{vol}(W)\leq A_{1}\cdot(r^{\prime})^{\dim(SM)}\cdot\varphi^{m\cdot\dim(SM)},

where A1>0A_{1}>0. Since X𝒫KX\in\mathcal{P}_{K} contains a ball of radius rr^{\prime}, the volume of XX is bounded below by

(20) A2(r)dim(SM)vol(X),\displaystyle A_{2}\cdot(r^{\prime})^{\dim(SM)}\leq\text{vol}(X),

where A2>0A_{2}>0. From (19) and (20) we have that

card{X𝒫:XD}\displaystyle\text{card}\left\{X\in\mathcal{P}:X\cap D\neq\emptyset\right\} A1A2φmdim(SM)+1\displaystyle\leq\dfrac{A_{1}}{A_{2}}\cdot\varphi^{m\cdot\dim(SM)}+1
(A1A2+1)φmdim(SM).\displaystyle\leq\left(\dfrac{A_{1}}{A_{2}}+1\right)\cdot\varphi^{m\cdot\dim(SM)}.

Case II: X=SMKX^{\prime}=SM\setminus K.
In this case, we have that

card{X𝒫:XD}\displaystyle\text{card}\left\{X\in\mathcal{P}:X\cap D\neq\emptyset\right\} |𝒫K|+1\displaystyle\leq\left|\mathcal{P}_{K}\right|+1
(ζ+1)(ϱm)dim(SM).\displaystyle\leq(\zeta+1)(\varrho_{m})^{-\dim(SM)}.

Considering L1=max{A1A2+1,ζ+1}L_{1}=\max\left\{\dfrac{A_{1}}{A_{2}}+1,\zeta+1\right\} we obtain the desired result. ∎

Now we will get a finer exponential bound for the number of those sets Dϕm𝒫KD\in\phi^{m}\mathcal{P}_{K} that contain regular points. For this, let Λm\Lambda_{m} be the set of regular points θSM\theta\in SM which satisfy the following condition: for kmk\geq m and ξTθSM\xi\in T_{\theta}SM

ek(𝒳(θ,ξ)ε)ξdϕθk(ξ)ek(𝒳(θ,ξ)+ε)ξ,\displaystyle e^{k\left(\mathcal{X}(\theta,\xi)-\varepsilon\right)}\left\|\xi\right\|\leq\left\|d\phi^{k}_{\theta}(\xi)\right\|\leq e^{k\left(\mathcal{X}(\theta,\xi)+\varepsilon\right)}\left\|\xi\right\|,

where 𝒳(θ,ξ)=limn±1nlogdϕθn(ξ)\mathcal{X}(\theta,\xi)=\displaystyle\lim_{n\rightarrow\pm\infty}\dfrac{1}{n}\log\left\|d\phi^{n}_{\theta}(\xi)\right\|.

Lemma 5.3.

If Dϕm𝒫KD\in\phi^{m}\mathcal{P}_{K} has non-empty intersection with Λm\Lambda_{m}, then there is a constant L2>0L_{2}>0 such that

card{X𝒫:XD}L2emεi:𝒳i>0em(𝒳i+ε)ki.\displaystyle\emph{card}\left\{X\in\mathcal{P}:X\cap D\neq\emptyset\right\}\leq L_{2}\cdot e^{m\varepsilon}\prod_{i:\mathcal{X}_{i}>0}e^{m(\mathcal{X}_{i}+\varepsilon)k_{i}}.
Proof.

Let X𝒫KX^{\prime}\in\mathcal{P}_{K} such that ϕm(X)=D\phi^{m}(X^{\prime})=D and suppose that XΛmX^{\prime}\cap\Lambda_{m}\neq\emptyset. Pick a point θXΛm\theta\in X^{\prime}\cap\Lambda_{m} and consider the ball B=B(0,ϱ)TθSMB=B(0,\varrho)\subset T_{\theta}SM. We claim that

Xexpθ(B(0,ϱm)),X^{\prime}\subseteq exp_{\theta}(B(0,\varrho_{m})),

where expθexp_{\theta} denotes the exponential map defined on the tangent plane TθSMT_{\theta}SM. In fact, let zXz\in X^{\prime}. Since SMSM is complete with the Sasaki metric (see Lemma 2.1) we can choose wTθSMw\in T_{\theta}SM such that γ(t)=expθ(tw)\gamma(t)=exp_{\theta}(tw), where γ\gamma is a geodesic with γ(0)=θ\gamma(0)=\theta and γ(1)=expθ(w)=z\gamma(1)=exp_{\theta}(w)=z. As diam 𝒫K<ϱm\mathcal{P}_{K}<\varrho_{m} then

d(θ,z)=l(γ)<ϱm.d(\theta,z)=l(\gamma)<\varrho_{m}.

Similar to the proof of Proposition 2.2, we obtain that

ϱm>01γ(s)𝑑s=w.\displaystyle\varrho_{m}>\int_{0}^{1}\left\|\gamma^{\prime}(s)\right\|ds=\left\|w\right\|.

Then wB(0,ϱm)w\in B(0,\varrho_{m}) and hence

z=expθ(w)expθ(B(0,ϱm)).z=exp_{\theta}(w)\in exp_{\theta}(B(0,\varrho_{m})).

Since zXz\in X^{\prime} was arbitrary, the claim is proven. Therefore, from Theorem 5.1 we have that

D=ϕm(X)B0:=expϕm(θ)(B~0),\displaystyle D=\phi^{m}(X^{\prime})\subseteq B_{0}:=exp_{\phi^{m}(\theta)}(\tilde{B}_{0}),

where B~0=dϕθm(B)\tilde{B}_{0}=d\phi^{m}_{\theta}(B) is an ellipsoid. Since the curvature tensor and the derivative of the curvature tensor of MM are both uniformly bounded, we have that the Sasaki sectional curvature of SMSM is uniformly bounded (see (2.1)). This implies that the curvature tensor of SMSM is uniformly bounded. Applying Proposition 2.2 to SMSM, there exists t0>0t_{0}>0 such that

(21) d(expϕm(θ))tv52\displaystyle\left\|d(exp_{\phi^{m}(\theta)})_{tv}\right\|\leq\dfrac{5}{2}

for every |t|t0\left|t\right|\leq t_{0} and vTϕm(θ)SMv\in T_{\phi^{m}(\theta)}SM with v=1\left\|v\right\|=1. Then, for mm large enough, we have that

diam(D)\displaystyle\text{diam}(D) h(c)mdiam(X)\displaystyle\leq h(c)^{m}\cdot\text{diam}(X^{\prime})
h(c)mϱm\displaystyle\leq h(c)^{m}\cdot\varrho_{m}
=1Cτ2τ111+c2e2cmϱ\displaystyle=\dfrac{1}{C}\cdot\dfrac{\tau_{2}}{\tau_{1}}\cdot\dfrac{1}{1+c^{2}}\cdot e^{-2cm}\cdot\varrho
<t02,\displaystyle<\dfrac{t_{0}}{2},

where h(c)h(c) is the expression that bounds the derivative of ϕ\phi (see Proposition 4.4). Therefore, we can choose B0B_{0} that satisfies DB0D\subset B_{0} and diam(B0)<t0(B_{0})<t_{0}. We know that diam𝒫K<ϱm<ϱ\mathcal{P}_{K}<\varrho_{m}<\varrho, then if X𝒫KX\in\mathcal{P}_{K} intersects DD, it lies in the set

B1={ΨSM:d(Ψ,B0)<ϱ}.B_{1}=\left\{\Psi\in SM:d(\Psi,B_{0})<\varrho\right\}.

Since XB(x,r)X\subset B(x,r) and 2r<ϱm<ϱ2r<\varrho_{m}<\varrho, then B(x,ϱ/2)B1B(x,\varrho/2)\subset B_{1} and

(22) card{X𝒫K:XD}bvol(B1)ϱdim(SM),\displaystyle\text{card}\left\{X\in\mathcal{P}_{K}:X\cap D\neq\emptyset\right\}\leq b\cdot\text{vol}(B_{1})\cdot\varrho^{-\dim(SM)},

for some b>0b>0, where vol(B1)(B_{1}) denotes the volume of B1B_{1} induced by the Sasaki metric. Consider a subset B~0B~0\tilde{B}^{*}_{0}\subset\tilde{B}_{0} such that expϕm(θ)exp_{\phi^{m}(\theta)} is a diffeomorphism between B~0\tilde{B}^{*}_{0} and B0B_{0}. Since

|detd(expϕm(θ))v|d(expϕm(θ))vdim(SM)\displaystyle\left|\det d(exp_{\phi^{m}(\theta)})_{v}\right|\leq\left\|d(exp_{\phi^{m}(\theta)})_{v}\right\|^{\dim(SM)}

for every vB~0v\in\tilde{B}^{*}_{0}, from (21) we have that

vol(B0)(52)dim(SM)vol(B~0).\displaystyle\text{vol}(B_{0})\leq\left(\dfrac{5}{2}\right)^{\dim(SM)}\cdot\text{vol}(\tilde{B}_{0}).

This implies that the volume of B1B_{1} is bounded, up to a bounded factor, by the product of the lengths of the axes of the ellipsoid B~0\tilde{B}_{0}. Those corresponding to non-positive Lyapunov exponents are at most sub-exponentially large. The remaining ones are of size at most em(𝒳i+ε)e^{m(\mathcal{X}_{i}+\varepsilon)}, up to a bounded factor, for all sufficiently large mm. Thus

vol(B1)\displaystyle\text{vol}(B_{1}) Aemε(diam(B))dim(SM)i:𝒳i>0em(𝒳i+ε)ki\displaystyle\leq A\cdot e^{m\varepsilon}\cdot(\text{diam}(B))^{\dim(SM)}\prod_{i:\mathcal{X}_{i}>0}e^{m(\mathcal{X}_{i}+\varepsilon)k_{i}}
Aemε(2ϱ)dim(SM)i:𝒳i>0em(𝒳i+ε)ki\displaystyle\leq A\cdot e^{m\varepsilon}\cdot(2\varrho)^{\dim(SM)}\prod_{i:\mathcal{X}_{i}>0}e^{m(\mathcal{X}_{i}+\varepsilon)k_{i}}
=A~emεϱdim(SM)i:𝒳i>0em(𝒳i+ε)ki,\displaystyle=\tilde{A}\cdot e^{m\varepsilon}\cdot\varrho^{\dim(SM)}\prod_{i:\mathcal{X}_{i}>0}e^{m(\mathcal{X}_{i}+\varepsilon)k_{i}},

where A~=A2dim(SM)\tilde{A}=A\cdot 2^{\dim(SM)}, for some A>0A>0. Then substituting in (22) we have that

card{X𝒫:XD}\displaystyle\text{card}\left\{X\in\mathcal{P}:X\cap D\neq\emptyset\right\} bvol(B1)ϱdim(SM)+1\displaystyle\leq b\cdot\text{vol}(B_{1})\cdot\varrho^{-\dim(SM)}+1
bA~emεi:𝒳i>0em(𝒳i+ε)ki+1\displaystyle\leq b\cdot\tilde{A}\cdot e^{m\varepsilon}\prod_{i:\mathcal{X}_{i}>0}e^{m(\mathcal{X}_{i}+\varepsilon)k_{i}}+1
(bA~+1)emεi:𝒳i>0em(𝒳i+ε)ki.\displaystyle\leq(b\cdot\tilde{A}+1)\cdot e^{m\varepsilon}\prod_{i:\mathcal{X}_{i}>0}e^{m(\mathcal{X}_{i}+\varepsilon)k_{i}}.

Considering L2=bA~+1L_{2}=b\cdot\tilde{A}+1 we obtain the desired result. ∎

Proof of Theorem 1.1. We have that μ(SMK)<ε\mu(SM\setminus K)<\varepsilon. From (5), Lemmas 5.2 and 5.3 we obtain

mhμ(ϕ)ε\displaystyle mh_{\mu}(\phi)-\varepsilon =hμ(ϕm)ε\displaystyle=h_{\mu}(\phi^{m})-\varepsilon
hμ(ϕm,𝒫)\displaystyle\leq h_{\mu}(\phi^{m},\mathcal{P})
Dϕm𝒫μ(D)logcard{X𝒫:XD}\displaystyle\leq\sum_{D\in\phi^{m}\mathcal{P}}\mu(D)\cdot\log\text{card}\left\{X\in\mathcal{P}:X\cap D\neq\emptyset\right\}
Dϕm𝒫K,DΛm=μ(D)logcard{X𝒫:XD}\displaystyle\leq\sum_{D\in\phi^{m}\mathcal{P}_{K},D\cap\Lambda_{m}=\emptyset}\mu(D)\cdot\log\text{card}\left\{X\in\mathcal{P}:X\cap D\neq\emptyset\right\}
+Dϕm𝒫K,DΛmμ(D)logcard{X𝒫:XD}\displaystyle\hskip 14.22636pt+\sum_{D\in\phi^{m}\mathcal{P}_{K},D\cap\Lambda_{m}\neq\emptyset}\mu(D)\cdot\log\text{card}\left\{X\in\mathcal{P}:X\cap D\neq\emptyset\right\}
+μ(ϕm(SMK))logcard{X𝒫:Xϕm(SMK)}\displaystyle\hskip 14.22636pt+\mu(\phi^{m}(SM\setminus K))\cdot\log\text{card}\left\{X\in\mathcal{P}:X\cap\phi^{m}(SM\setminus K)\neq\emptyset\right\}
Dϕm𝒫K,DΛm=μ(D)(log(L1)+dim(SM)max{mlog(φ),log(ϱm)})\displaystyle\leq\sum_{D\in\phi^{m}\mathcal{P}_{K},D\cap\Lambda_{m}=\emptyset}\mu(D)\left(\log(L_{1})+\dim(SM)\cdot\max\left\{m\log(\varphi),-\log(\varrho_{m})\right\}\right)
+Dϕm𝒫K,DΛmμ(D)(log(L2)+mε+mi:𝒳i>0(𝒳i+ε)ki)\displaystyle\hskip 14.22636pt+\sum_{D\in\phi^{m}\mathcal{P}_{K},D\cap\Lambda_{m}\neq\emptyset}\mu(D)\left(\log(L_{2})+m\varepsilon+m\sum_{i:\mathcal{X}_{i}>0}(\mathcal{X}_{i}+\varepsilon)k_{i}\right)
+μ(SMK)(log(L1)+dim(SM)max{mlog(φ),log(ϱm)})\displaystyle\hskip 14.22636pt+\mu(SM\setminus K)\cdot\left(\log(L_{1})+\dim(SM)\cdot\max\left\{m\log(\varphi),-\log(\varrho_{m})\right\}\right)
(log(L1)+dim(SM)max{mlog(φ),log(ϱm)})μ(SMΛm)\displaystyle\leq\left(\log(L_{1})+\dim(SM)\cdot\max\left\{m\log\left(\varphi\right),-\log(\varrho_{m})\right\}\right)\cdot\mu(SM\setminus\Lambda_{m})
+log(L2)+mε+mi:𝒳i>0(𝒳i+ε)ki\displaystyle\hskip 14.22636pt+\log(L_{2})+m\varepsilon+m\sum_{i:\mathcal{X}_{i}>0}(\mathcal{X}_{i}+\varepsilon)k_{i}
(23) +ε(log(L1)+dim(SM)max{mlog(φ),log(ϱm)}).\displaystyle\hskip 14.22636pt+\varepsilon\cdot\left(\log(L_{1})+\dim(SM)\cdot\max\left\{m\log\left(\varphi\right),-\log(\varrho_{m})\right\}\right).

By Oseledec’s Theorem we have that μ(SMΛm)0\mu(SM\setminus\Lambda_{m})\rightarrow 0 as mm\rightarrow\infty. Moreover,

limm+1mlog(ϱm)=log(h(c))2c,\displaystyle\lim_{m\rightarrow+\infty}\dfrac{1}{m}\log(\varrho_{m})=-\log(h(c))-2c,

where h(c)h(c) is the expression that bounds the derivative of ϕ\phi (see Proposition 4.4). Then, dividing by mm in (5) and taking m+m\rightarrow+\infty we obtain

hμ(ϕ)ε+i:𝒳i>0(𝒳i+ε)ki+εdim(SM)max{log(φ),log(h(c))+2c}.\displaystyle h_{\mu}(\phi)\leq\varepsilon+\sum_{i:\mathcal{X}_{i}>0}(\mathcal{X}_{i}+\varepsilon)k_{i}+\varepsilon\cdot\dim(SM)\cdot\max\left\{\log\left(\varphi\right),\log(h(c))+2c\right\}.

Letting ε0\varepsilon\rightarrow 0 we have

hμ(ϕ)i:𝒳i>0𝒳iki,\displaystyle h_{\mu}(\phi)\leq\sum_{i:\mathcal{X}_{i}>0}\mathcal{X}_{i}k_{i},

which is the desired upper bound. \hfill\square

6. Pesin’s Formula

In this section, we aim to prove Theorem 1.2. To achieve this goal, we will use the techniques applied by Mañé in [11] which don’t use the theory of stable manifolds. Adopting this strategy greatly simplifies our proof since we only need to corroborate that all the technical hypotheses used by Mañé continue to be satisfied under the condition of the geodesic flow being Anosov. To simplify notation, we write

𝒳+(θ)=𝒳i(θ)>0𝒳i(θ)dim(Hi(θ)).\displaystyle\mathcal{X}^{+}(\theta)=\sum_{\mathcal{X}_{i}(\theta)>0}\mathcal{X}_{i}(\theta)\cdot\dim(H_{i}(\theta)).

We start introducing some notations. Set g:SMSMg:SM\rightarrow SM a map and ρ:SM(0,1)\rho:SM\rightarrow(0,1) a function. For θSM\theta\in SM and n0n\geq 0, define

Sn(g,ρ,θ)={ωSM:d(gj(θ),gj(ω))ρ(gj(θ)),0jn}.\displaystyle S_{n}(g,\rho,\theta)=\left\{\omega\in SM:d(g^{j}(\theta),g^{j}(\omega))\leq\rho(g^{j}(\theta)),0\leq j\leq n\right\}.

If μ\mu is a measure on SMSM and gg and ρ\rho are measurable, define

hμ(g,ρ,θ)=lim supn1nlogμ(Sn(g,ρ,θ)).\displaystyle h_{\mu}(g,\rho,\theta)=\limsup_{n\rightarrow\infty}-\dfrac{1}{n}\log\mu(S_{n}(g,\rho,\theta)).

Let EE be a normed space and E=E1E2E=E_{1}\oplus E_{2} a splitting. We say that a subset WEW\subset E is a (E1,E2)(E_{1},E_{2})-graph if there exists an open set UE2U\subset E_{2} and a C1C^{1}-map ψ:UE1\psi:U\rightarrow E_{1} such that W={(ψ(x),x):xU}W=\left\{(\psi(x),x):x\in U\right\}. The number

sup{ψ(x)ψ(y)xy:x,yU,xy}\sup\left\{\dfrac{\left\|\psi(x)-\psi(y)\right\|}{\left\|x-y\right\|}:x,y\in U,x\neq y\right\}

is called the dispersion of WW.

Let MM be a complete Riemannian manifold and μ\mu an ϕt\phi^{t}-invariant probability measure on SMSM satisfying the assumptions of Theorem 1.2. Denote by ν\nu the Lebesgue measure on SMSM. Since the geodesic flow is of Anosov type, consider

Ecs(θ)=G(θ)Es(θ)E^{cs}(\theta)=\left\langle G(\theta)\right\rangle\oplus E^{s}(\theta)

for every θSM\theta\in SM. From Theorem 3.1 there is a set ΛSM\Lambda\subset SM such that μ(SMΛ)=0\mu(SM\setminus\Lambda)=0 and the Lyapunov exponents of ϕ\phi exist for every θΛ\theta\in\Lambda. Fix any ε>0\varepsilon>0. By Egorov’s and Oseledec’s Theorems, there is a compact set KΛK\subset\Lambda with μ(K)1ε\mu(K)\geq 1-\varepsilon such that the splitting TθSM=Ecs(θ)Eu(θ)T_{\theta}SM=E^{cs}(\theta)\oplus E^{u}(\theta) is continuous when θ\theta varies in KK and, for some N>0N>0, there are constants α>β>1\alpha>\beta>1 such that, if g=ϕNg=\phi^{N}, the inequalities

dgθn(η)\displaystyle\left\|dg^{n}_{\theta}(\eta)\right\| αnη\displaystyle\geq\alpha^{n}\left\|\eta\right\|
dgθn|Ecs(θ)\displaystyle\left\|\left.dg^{n}_{\theta}\right|_{E^{cs}(\theta)}\right\| βn\displaystyle\leq\beta^{n}
(24) log|det(dgθn|Eu(θ))|\displaystyle\log\left|\det\left(\left.dg^{n}_{\theta}\right|_{E^{u}(\theta)}\right)\right| Nn(𝒳+(θ)ε)\displaystyle\geq Nn\left(\mathcal{X}^{+}(\theta)-\varepsilon\right)

hold for all θK\theta\in K, n0n\geq 0 and ηEu(θ)\eta\in E^{u}(\theta).

In the same way as in [11], in the remainder of this section, we will treat SMSM as if it were an Euclidean space. The arguments we use can be formalized without any difficulty by the direct use of local coordinates. Since the geodesic flow is C1C^{1}-Hölder, we have the following result proved by Mañe in [11].

Lemma 6.1.

For every σ>0\sigma>0 there is ξ>0\xi>0 such that, if θK\theta\in K and gm(θ)Kg^{m}(\theta)\in K for some m>0m>0, then if a set WSMW\subset SM is contained in the ball Bξm(θ)B_{\xi^{m}}(\theta) and is a (Ecs(θ),Eu(θ))(E^{cs}(\theta),E^{u}(\theta))-graph with dispersion σ\leq\sigma, then gm(W)g^{m}(W) is a (Ecs(gm(θ)),Eu(gm(θ)))(E^{cs}(g^{m}(\theta)),E^{u}(g^{m}(\theta)))-graph with dispersion σ\leq\sigma.

Fix the constant σ>0\sigma>0 of the statement of Lemma 6.1 small enough such that exists a(0,1)a\in(0,1), at0/2a\leq t_{0}/2, where t0t_{0} comes from Proposition 2.2 applied to SMSM, with the following property: if θK\theta\in K, ωSM\omega\in SM and d(θ,ω)<ad(\theta,\omega)<a, then for every subspace ETωSME\subset T_{\omega}SM which is a (Ecs(θ),Eu(θ))(E^{cs}(\theta),E^{u}(\theta))-graph with dispersion σ\leq\sigma we have

(25) |log|det(dgω|E)|log|det(dgθ|Eu(θ))||ε.\displaystyle\left|\log\left|\det\left(\left.dg_{\omega}\right|_{E}\right)\right|-\log|\det(\left.dg_{\theta}\right|_{E^{u}(\theta)})|\right|\leq\varepsilon.

We proved in Theorem 3.1 that the norm of the derivative of ϕ\phi is bounded, then denote

P=sup{log|det(dϕθ|E)|:θSM,ETθSM}.P=\sup\left\{\log\left|\det\left.\left(d\phi_{\theta}\right|_{E}\right)\right|:\theta\in SM,E\subset T_{\theta}SM\right\}.

The following proposition is an adaptation of Mañe’s result in [11] applied to the case of Anosov geodesic flow for non-compact manifolds. To ensure a comprehensive understanding of our arguments, we chose to include the full proof provided by Mañé.

Proposition 6.2.

For every small ε>0\varepsilon>0, there exist a function ρ:SM(0,1)\rho:SM\rightarrow(0,1) with logρL1(SM,μ)\log\rho\in L^{1}(SM,\mu), an integer N>0N>0 and a compact set KSMK^{\prime}\subset SM with μ(SMK)2ε\mu(SM\setminus K^{\prime})\leq 2\sqrt{\varepsilon} such that

hν(ϕN,ρ,θ)N(𝒳+(θ)εεN4Pε)\displaystyle h_{\nu}(\phi^{N},\rho,\theta)\geq N\left(\mathcal{X}^{+}(\theta)-\varepsilon-\dfrac{\varepsilon}{N}-4P\sqrt{\varepsilon}\right)

for every θK\theta\in K^{\prime}.

Proof.

For θK\theta\in K, define L(θ)L(\theta) as the minimum integer 1\geq 1 such that gL(θ)(θ)Kg^{L(\theta)}(\theta)\in K. This function is well defined for μ\mu-almost every θK\theta\in K and it is integrable. Extend LL to SMSM, putting L(θ)=0L(\theta)=0 when θK\theta\notin K and at points of KK that do not return to this set. Define ρ:SM(0,1)\rho:SM\rightarrow(0,1) as

(26) ρ(θ)=min{a,ξL(θ)},\displaystyle\rho(\theta)=\min\left\{a,\xi^{L(\theta)}\right\},

where a(0,t0/2)a\in(0,t_{0}/2) comes from property (25) and ξ>0\xi>0 comes from Lemma 6.1. Since LL is integrable then clearly logρ\log\rho is also integrable. On the other hand, by Birkhoff’s ergodic theorem, the function

Ψ(θ)=limn+1ncard{0j<n:gj(θ)ΛK}\displaystyle\Psi(\theta)=\lim_{n\rightarrow+\infty}\dfrac{1}{n}\text{card}\left\{0\leq j<n:g^{j}(\theta)\in\Lambda\setminus K\right\}

is defined for μ\mu-almost every θΛ\theta\in\Lambda. Then

εμ(ΛK)\displaystyle\varepsilon\geq\mu(\Lambda\setminus K) =ΛΨ𝑑μ\displaystyle=\int_{\Lambda}\Psi d\mu
{θΛ:Ψ(θ)>ε}Ψ𝑑μ\displaystyle\geq\int_{\left\{\theta\in\Lambda:\Psi(\theta)>\sqrt{\varepsilon}\right\}}\Psi d\mu
>εμ({θΛ:Ψ(θ)>ε}).\displaystyle>\sqrt{\varepsilon}\cdot\mu\left(\left\{\theta\in\Lambda:\Psi(\theta)>\sqrt{\varepsilon}\right\}\right).

Therefore,

μ({θΛ:Ψ(θ)ε})1ε.\displaystyle\mu\left(\left\{\theta\in\Lambda:\Psi(\theta)\leq\sqrt{\varepsilon}\right\}\right)\geq 1-\sqrt{\varepsilon}.

By Egorov’s Theorem, there exists a compact set KKK^{\prime}\subset K with μ(K)12ε\mu(K^{\prime})\geq 1-2\sqrt{\varepsilon} and N0>0N_{0}>0 such that, if nN0n\geq N_{0},

(27) card{0j<n:gj(θ)ΛK}2nε\displaystyle\text{card}\left\{0\leq j<n:g^{j}(\theta)\in\Lambda\setminus K\right\}\leq 2n\sqrt{\varepsilon}

for all θK\theta\in K^{\prime}. Since the subspaces Ecs(θ)E^{cs}(\theta) and Eu(θ)E^{u}(\theta) are not necessary orthogonal, there exists B>0B>0 such that

(28) ν(Sn(g,ρ,θ))BEcs(θ)ν((ω+Eu(θ))Sn(g,ρ,θ))𝑑ν(ω)\displaystyle\nu(S_{n}(g,\rho,\theta))\leq B\int_{E^{cs}(\theta)}\nu\left(\left(\omega+E^{u}(\theta)\right)\cap S_{n}(g,\rho,\theta)\right)d\nu(\omega)

for every θK\theta\in K^{\prime} and n0n\geq 0, where ν\nu also denotes the Lebesgue measure in the subspaces Ecs(θ)E^{cs}(\theta) and ω+Eu(θ)\omega+E^{u}(\theta). For ωEcs(θ)\omega\in E^{cs}(\theta), denote by

Ωn(ω)=(ω+Eu(θ))Sn(g,ρ,θ).\Omega_{n}(\omega)=\left(\omega+E^{u}(\theta)\right)\cap S_{n}(g,\rho,\theta).

Take D>0D>0 such that D>D> vol(W)(W) for every (Ecs(θ),Eu(θ))(E^{cs}(\theta),E^{u}(\theta))-graph WW with dispersion σ\leq\sigma contained in Bρ(θ)(θ)B_{\rho(\theta)}(\theta), where θK\theta\in K^{\prime} and ρ\rho is the function defined in (26). This constant exists because the domain of the graphs is contained in a ball of radius <1<1 and the derivatives of the functions defining the graphs are uniformly bounded in norm by σ\sigma. If gn(θ)Kg^{n}(\theta)\in K^{\prime} and ωEcs(θ)\omega\in E^{cs}(\theta), from Lemma 55 of [11] we have that gn(Ωn(ω))g^{n}(\Omega_{n}(\omega)) is a (Ecs(gn(θ)),Eu(gn(θ)))(E^{cs}(g^{n}(\theta)),E^{u}(g^{n}(\theta)))-graph with dispersion σ\leq\sigma and

(29) D>vol(gn(Ωn(ω)))=Ωn(ω)|detdgzn|TzΩn(ω)|dν(z).\displaystyle D>\text{vol}(g^{n}(\Omega_{n}(\omega)))=\int_{\Omega_{n}(\omega)}\left|\det\left.dg^{n}_{z}\right|_{T_{z}\Omega_{n}(\omega)}\right|d\nu(z).

Fix any θK\theta\in K^{\prime} and let Sn={0j<n:gj(θ)K}S_{n}=\left\{0\leq j<n:g^{j}(\theta)\in K^{\prime}\right\}. If nN0n\geq N_{0}, it follows from (6), (25) and (27) that for ωEcs(θ)\omega\in E^{cs}(\theta) we have

log|det(dgzn|TzΩn(ω))|\displaystyle\log\left|\det\left(\left.dg^{n}_{z}\right|_{T_{z}\Omega_{n}(\omega)}\right)\right| =j=0n1log|det(dggj(z)|Tgj(z)gj(Ωn(ω)))|\displaystyle=\sum_{j=0}^{n-1}\log\left|\det\left(\left.dg_{g^{j}(z)}\right|_{T_{g^{j}(z)}g^{j}(\Omega_{n}(\omega))}\right)\right|
jSnlog|det(dggj(z)|Tgj(z)gj(Ωn(ω)))|NP(ncardSn)\displaystyle\geq\sum_{j\in S_{n}}\log\left|\det\left(\left.dg_{g^{j}(z)}\right|_{T_{g^{j}(z)}g^{j}(\Omega_{n}(\omega))}\right)\right|-NP(n-\text{card}S_{n})
jSnlog|det(dggj(θ)|Eu(gj(θ)))|εnNP(ncardSn)\displaystyle\geq\sum_{j\in S_{n}}\log\left|\det\left(\left.dg_{g^{j}(\theta)}\right|_{E^{u}(g^{j}(\theta))}\right)\right|-\varepsilon n-NP(n-\text{card}S_{n})
j=0n1log|det(dggj(θ)|Eu(gj(θ)))|εn2NP(ncardSn)\displaystyle\geq\sum_{j=0}^{n-1}\log\left|\det\left(\left.dg_{g^{j}(\theta)}\right|_{E^{u}(g^{j}(\theta))}\right)\right|-\varepsilon n-2NP(n-\text{card}S_{n})
=log|det(dgθn|Eu(θ))|εn2NP(ncardSn)\displaystyle=\log\left|\det\left(\left.dg^{n}_{\theta}\right|_{E^{u}(\theta)}\right)\right|-\varepsilon n-2NP(n-\text{card}S_{n})
nN(𝒳+(θ)ε)εn2NP(ncardSn)\displaystyle\geq nN(\mathcal{X}^{+}(\theta)-\varepsilon)-\varepsilon n-2NP(n-\text{card}S_{n})
nN(𝒳+(θ)ε)εn4NPnε.\displaystyle\geq nN(\mathcal{X}^{+}(\theta)-\varepsilon)-\varepsilon n-4NPn\sqrt{\varepsilon}.

From (29) we obtain that

D>ν(Ωn(ω))exp(nN(𝒳+(θ)ε)εn4NPnε)\displaystyle D>\nu(\Omega_{n}(\omega))\cdot\exp\left(nN(\mathcal{X}^{+}(\theta)-\varepsilon)-\varepsilon n-4NPn\sqrt{\varepsilon}\right)

for every θK\theta\in K^{\prime} and ωEcs(θ)\omega\in E^{cs}(\theta). It follows from (28) that

ν(Sn(g,ρ,θ))BDexp(nN(𝒳+(θ)ε)+εn+4NPnε).\displaystyle\nu(S_{n}(g,\rho,\theta))\leq B\cdot D\cdot\exp\left(-nN(\mathcal{X}^{+}(\theta)-\varepsilon)+\varepsilon n+4NPn\sqrt{\varepsilon}\right).

Therefore, for every θK\theta\in K^{\prime},

hν(g,ρ,θ)=lim supn1nlogν(Sn(g,ρ,θ))N(𝒳+(θ)εεN4Pε).\displaystyle h_{\nu}(g,\rho,\theta)=\limsup_{n\rightarrow\infty}-\dfrac{1}{n}\log\nu(S_{n}(g,\rho,\theta))\geq N\left(\mathcal{X}^{+}(\theta)-\varepsilon-\dfrac{\varepsilon}{N}-4P\sqrt{\varepsilon}\right).

This completes the proof of the proposition. ∎

We will show that the function ρ\rho of Proposition 6.2 allows us to find a lower bound for the entropy of ϕN\phi^{N}. To prove this, Mañé constructed a partition of the manifold with certain properties using strongly the compactness condition (see Lemma 2 of [11]). Since the manifold SMSM is not necessarily compact in our case, we will use another technique to construct a partition that satisfies the same properties. Consider the constant a(0,1)a\in(0,1), a<t0/2a<t_{0}/2, used in property (25).

Lemma 6.3.

Let MM be a complete Riemannian manifold and suppose that the curvature tensor and the derivative of the curvature tensor are both uniformly bounded. For every θSM\theta\in SM we have that

diamexpθU52diam U,\displaystyle\emph{diam}\,exp_{\theta}U\leq\dfrac{5}{2}\cdot\emph{diam }U,

where UB(0,a)TθSMU\subset B(0,a)\subset T_{\theta}SM.

Proof.

Fix θSM\theta\in SM and consider UB(0,a)TθSMU\subset B(0,a)\subset T_{\theta}SM. We need to prove that

d(expθu,expθv)52uv\displaystyle d(exp_{\theta}u,exp_{\theta}v)\leq\dfrac{5}{2}\left\|u-v\right\|

for every u,vUu,v\in U. Consider the segment q(t)=tu+(1t)vq(t)=tu+(1-t)v and the curve γ(t)=expθq(t)\gamma(t)=exp_{\theta}q(t) that joins expθuexp_{\theta}u with expθvexp_{\theta}v. Then

l(γ)\displaystyle l(\gamma) =01γ(t)\displaystyle=\int_{0}^{1}\left\|\gamma^{\prime}(t)\right\|
(30) =01d(expθ)q(t)(uv)𝑑t.\displaystyle=\int_{0}^{1}\left\|d(exp_{\theta})_{q(t)}(u-v)\right\|dt.

For each t[0,1]t\in[0,1], there are w(t)TθSMw(t)\in T_{\theta}SM with w(t)=1\left\|w(t)\right\|=1 and s(t)s(t)\in\mathbb{R} with |s(t)|t0\left|s(t)\right|\leq t_{0} such that

q(t)=s(t)w(t).\displaystyle q(t)=s(t)w(t).

Since at0/2a\leq t_{0}/2, from Proposition 2.2 we have that

d(expθ)q(t)(uv)\displaystyle\left\|d(exp_{\theta})_{q(t)}(u-v)\right\| =d(expθ)s(t)w(t)(uv)\displaystyle=\left\|d(exp_{\theta})_{s(t)w(t)}(u-v)\right\|
52uv.\displaystyle\leq\dfrac{5}{2}\left\|u-v\right\|.

Therefore in (6)

d(expθu,expθv)l(γ)52uv\displaystyle d(exp_{\theta}u,exp_{\theta}v)\leq l(\gamma)\leq\dfrac{5}{2}\left\|u-v\right\|

completing the proof. ∎

Consider the function ρ:SM(0,1)\rho:SM\rightarrow(0,1) defined in (26).

Lemma 6.4.

There exists a countable partition 𝒫\mathcal{P} of SMSM with finite entropy such that, if 𝒫(θ)\mathcal{P}(\theta) denotes the atom of 𝒫\mathcal{P} containing θ\theta, then

diam 𝒫(θ)ρ(θ)\displaystyle\emph{diam }\mathcal{P}(\theta)\leq\rho(\theta)

for μ\mu-almost every θSM\theta\in SM.

Proof.

For each n0n\geq 0, define

Un={θSM:e(n+1)<ρ(θ)en}.\displaystyle U_{n}=\left\{\theta\in SM:e^{-(n+1)}<\rho(\theta)\leq e^{-n}\right\}.

Since logρL1(SM,μ)\log\rho\in L^{1}(SM,\mu), we have that

n=0nμ(Un)n=0Unlogρ(θ)𝑑μ(θ)=SMlogρ(θ)𝑑μ(θ)<.\displaystyle\sum_{n=0}^{\infty}n\mu(U_{n})\leq-\sum_{n=0}^{\infty}\quad\int_{U_{n}}\log\rho(\theta)d\mu(\theta)=-\int_{SM}\log\rho(\theta)d\mu(\theta)<\infty.

Then, by Lemma 1 of [11] we obtain

(31) n=0μ(Un)logμ(Un)<.\displaystyle\sum_{n=0}^{\infty}\mu(U_{n})\log\mu(U_{n})<\infty.

For θSMK\theta\in SM\setminus K^{\prime} we have that ρ(θ)=a\rho(\theta)=a. Then there exists n00n_{0}\geq 0 such that

e(n0+1)<aen0e^{-(n_{0}+1)}<a\leq e^{-n_{0}}

and Un(SMK)=U_{n}\cap(SM\setminus K^{\prime})=\emptyset for every nn0n\neq n_{0}. This implies that UnKU_{n}\subset K^{\prime} for every nn0n\neq n_{0}. Define

Un0=Un0K.\displaystyle U_{n_{0}}^{*}=U_{n_{0}}\cap K^{\prime}.

Since KK^{\prime} is compact, there exist A>0A>0 and r0>0r_{0}>0 such that for all 0<rr00<r\leq r_{0}, there exists a partition 𝒬r\mathcal{Q}_{r} of KK^{\prime} whose atoms have diameter less than or equal to rr and such that the number of atoms in 𝒬r\mathcal{Q}_{r}, denoted by |𝒬r|\left|\mathcal{Q}_{r}\right|, satisfies

|𝒬r|A(1r)dim(SM).\displaystyle\left|\mathcal{Q}_{r}\right|\leq A\left(\dfrac{1}{r}\right)^{\dim(SM)}.

Define 𝒬\mathcal{Q} as the partition of KK^{\prime} given by

  • .

    Sets XUnX\cap U_{n}, for n0n\geq 0, nn0n\neq n_{0}, where X𝒬rnX\in\mathcal{Q}_{r_{n}} and rn=e(n+1)r_{n}=e^{-(n+1)} such that μ(XUn)>0\mu(X\cap U_{n})>0.

  • .

    Sets XUn0X\cap U_{n_{0}}^{*}, where X𝒬rn0X\in\mathcal{Q}_{r_{n_{0}}} and rn0=e(n0+1)r_{n_{0}}=e^{-(n_{0}+1)} such that μ(XUn0)>0\mu(X\cap U_{n_{0}}^{*})>0.

On the other hand, consider 0<ε<a/100<\varepsilon^{\prime}<a/10 such that, we can choose a measurable set (like a “ring” covering SMKSM\setminus K^{\prime})

V1{θSMK:d(θ,K)ε}:=E1\displaystyle V_{1}\subseteq\left\{\theta\in SM\setminus K^{\prime}:d(\theta,K^{\prime})\leq\varepsilon^{\prime}\right\}:=E_{1}

that satisfies

μ(V1)ε.\mu(V_{1})\leq\sqrt{\varepsilon}.

Define K1=KV1K^{\prime}_{1}=K^{\prime}\cup V_{1} and choose a measurable set (like a “ring” covering SMK1SM\setminus K^{\prime}_{1})

V2{θSMK1:d(θ,K1)ε}:=E2\displaystyle V_{2}\subseteq\left\{\theta\in SM\setminus K^{\prime}_{1}:d(\theta,K^{\prime}_{1})\leq\varepsilon^{\prime}\right\}:=E_{2}

that satisfies

μ(V2)ε2.\mu(V_{2})\leq\dfrac{\sqrt{\varepsilon}}{2}.

Proceeding inductively, we define bounded measurable sets

Vn{θSMKn1:d(θ,Kn1)ε}:=En,\displaystyle V_{n}\subseteq\left\{\theta\in SM\setminus K^{\prime}_{n-1}:d(\theta,K^{\prime}_{n-1})\leq\varepsilon^{\prime}\right\}:=E_{n},

where Kn1=KV1Vn1K^{\prime}_{n-1}=K^{\prime}\cup V_{1}\ldots\cup V_{n-1}, with measure

μ(Vn)ε2n1.\mu(V_{n})\leq\dfrac{\sqrt{\varepsilon}}{2^{n-1}}.

Since

n=1nμ(Vn)n=1n2n1ε<,\displaystyle\sum_{n=1}^{\infty}n\mu(V_{n})\leq\sum_{n=1}^{\infty}\dfrac{n}{2^{n-1}}\cdot\sqrt{\varepsilon}<\infty,

by Lemma 1 of [11] we have that

(32) n=1μ(Vn)logμ(Vn)<.\sum_{n=1}^{\infty}\mu(V_{n})\log\mu(V_{n})<\infty.

Let kk be the number of balls of radius a/10a/10 which cover E1E_{1} and denote by B(θ1,a/10),,B(\theta_{1},a/10),\ldots,B(θk,a/10)B(\theta_{k},a/10) this covering. We claim that

E2i=1kB(θi,a/5).\displaystyle E_{2}\subseteq\bigcup_{i=1}^{k}B(\theta_{i},a/5).

In fact, suppose that exists θE2\theta\in E_{2} such that d(θ,θi)a/5d(\theta,\theta_{i})\geq a/5, for every i=1,,ki=1,\ldots,k. By construction, there is ωE1\omega\in E_{1} such that

(33) d(θ,ω)ε<a10.\displaystyle d(\theta,\omega)\leq\varepsilon^{\prime}<\dfrac{a}{10}.

Since we cover E1E_{1} by balls, ωB(θi0,a/10)\omega\in B(\theta_{i_{0}},a/10) for some i0{1,,k}i_{0}\in\{1,\ldots,k\}. Therefore,

d(θ,ω)\displaystyle d(\theta,\omega) d(θ,θi0)d(θi0,ω)\displaystyle\geq d(\theta,\theta_{i_{0}})-d(\theta_{i_{0}},\omega)
>a5a10\displaystyle>\dfrac{a}{5}-\dfrac{a}{10}
=a10,\displaystyle=\dfrac{a}{10},

which is a contradiction with (33). This proves the claim. Since SMSM is complete (see Lemma 2.1), for each i{1,,k}i\in\{1,\ldots,k\}, there is an open ball Bi(0,a/5)TθiSMB^{i}(0,a/5)\subset T_{\theta_{i}}SM such that

expθi(Bi(0,a/5))=B(θi,a/5).exp_{\theta_{i}}(B^{i}(0,a/5))=B(\theta_{i},a/5).

By [20] there exists N1:=N1(a)>0N_{1}:=N_{1}(a)>0, which depends on the dimension of SMSM and aa, such that the minimal number of balls of radius a/10a/10 which can cover Bi(0,a/5)B^{i}(0,a/5) is bounded by N1N_{1}. Suppose that

B1i,,BN1i\displaystyle B^{i}_{1},\ldots,B^{i}_{N_{1}}

are balls of radius a/10a/10 that cover Bi(0,a/5)B^{i}(0,a/5). From Lemma 6.3, if we project these balls to the manifold SMSM by the exponential map we have that expθiBjiexp_{\theta_{i}}B^{i}_{j} are sets of diameter

diam expθiBji52diam Bji=522a10=a2.\displaystyle\text{diam }exp_{\theta_{i}}B^{i}_{j}\leq\dfrac{5}{2}\text{diam }B^{i}_{j}=\dfrac{5}{2}\cdot\dfrac{2a}{10}=\dfrac{a}{2}.

Then we can cover E2E_{2} by kN1kN_{1} sets of diameter a/2\leq a/2. Since every set of diameter a/2\leq a/2 is contained in a ball of radius a/2a/2, we can cover E2E_{2} by kN1kN_{1} balls B(ω1,a/2),,B(ωkN1,a/2)B(\omega_{1},a/2),\ldots,B(\omega_{kN_{1}},a/2). Analogously, since ε<a/10\varepsilon^{\prime}<a/10 we have that

E3i=1kN1B(ωi,6a/10).\displaystyle E_{3}\subset\bigcup_{i=1}^{kN_{1}}B(\omega_{i},6a/10).

For each i{1,,kN1}i\in\{1,\ldots,kN_{1}\}, there is an open ball Bi(0,6a/10)TωiSMB^{i}(0,6a/10)\subset T_{\omega_{i}}SM such that

expωi(Bi(0,6a/10))=B(ωi,6a/10).exp_{\omega_{i}}(B^{i}(0,6a/10))=B(\omega_{i},6a/10).

By [20] there exists N2:=N2(a)>0N_{2}:=N_{2}(a)>0, which depends on the dimension of SMSM and aa, such that the minimal number of balls of radius a/10a/10 which can cover Bi(0,6a/10)B^{i}(0,6a/10) is bounded by N2N_{2} and repeating the previous process we have that we can cover E3E_{3} by kN1N2kN_{1}N_{2} balls of radius a/2a/2. Continuing inductively, we obtain that EnE_{n} can be covered by kN1N2n2kN_{1}N_{2}^{n-2} balls of radius a/2a/2. Therefore, for every n1n\geq 1, define a partition 𝒫^n\hat{\mathcal{P}}_{n} of VnV_{n} whose atoms have diameter a\leq a and the number of atoms satisfies

|𝒫^1|k,|𝒫^n|kN1N2n2,n2.\displaystyle|\hat{\mathcal{P}}_{1}|\leq k,\hskip 17.07182pt|\hat{\mathcal{P}}_{n}|\leq kN_{1}N_{2}^{n-2},\hskip 8.5359pt\forall n\geq 2.

Finally, define the partition of SMSM as

𝒫=𝒬n1𝒫^n.\mathcal{P}=\mathcal{Q}\cup\bigcup_{n\geq 1}\hat{\mathcal{P}}_{n}.

Recalling the well-known inequality

i=1mxilogxi(i=1mxi)(logmlogi=1mxi)\displaystyle-\sum_{i=1}^{m}x_{i}\log x_{i}\leq\left(\sum_{i=1}^{m}x_{i}\right)\left(\log m-\log\sum_{i=1}^{m}x_{i}\right)

which holds for any set of real numbers 0<xi10<x_{i}\leq 1, i=1,,mi=1,\ldots,m. We claim that H(𝒫)<+H(\mathcal{P})<+\infty. In fact, from (31) and (32) we obtain that

H(𝒫)\displaystyle H(\mathcal{P}) =n0,nn0(P𝒬,PUnμ(P)logμ(P))+(P𝒬,PUn0μ(P)logμ(P))\displaystyle=\sum_{n\geq 0,\ n\neq n_{0}}\left(-\sum_{P\in\mathcal{Q},P\subset U_{n}}\mu(P)\log\mu(P)\right)+\left(-\sum_{P\in\mathcal{Q},P\subset U_{n_{0}}^{*}}\mu(P)\log\mu(P)\right)
+n1(P𝒫^nμ(P)logμ(P))\displaystyle\hskip 14.22636pt+\sum_{n\geq 1}\left(-\sum_{P\in\hat{\mathcal{P}}_{n}}\mu(P)\log\mu(P)\right)
n0,nn0μ(Un)[log|𝒬rn|logμ(Un)]+μ(Un0)[log|𝒬rn0|logμ(Un0)]\displaystyle\leq\sum_{n\geq 0,\ n\neq n_{0}}\mu(U_{n})\left[\log\left|\mathcal{Q}_{r_{n}}\right|-\log\mu(U_{n})\right]+\mu(U_{n_{0}}^{*})\left[\log|\mathcal{Q}_{r_{n_{0}}}|-\log\mu(U_{n_{0}}^{*})\right]
+n1,μ(Vn)[log|𝒫^n|logμ(Vn)]\displaystyle\hskip 14.22636pt+\sum_{n\geq 1,}\mu(V_{n})\left[\log|\hat{\mathcal{P}}_{n}|-\log\mu(V_{n})\right]
n0,nn0μ(Un)[logA+dim(SM)(n+1)logμ(Un)]\displaystyle\leq\sum_{n\geq 0,\ n\neq n_{0}}\mu(U_{n})\left[\log A+\dim(SM)(n+1)-\log\mu(U_{n})\right]
+μ(Un0)[logA+dim(SM)(n0+1)logμ(Un0)]+μ(V1)[logklogμ(V1)]\displaystyle\hskip 17.07182pt+\mu(U_{n_{0}}^{*})\left[\log A+\dim(SM)(n_{0}+1)-\log\mu(U_{n_{0}}^{*})\right]+\mu(V_{1})\left[\log k-\log\mu(V_{1})\right]
+n2μ(Vn)[logk+logN1+(n2)logN2logμ(Vn)]\displaystyle\hskip 17.07182pt+\sum_{n\geq 2}\mu(V_{n})\left[\log k+\log N_{1}+(n-2)\log N_{2}-\log\mu(V_{n})\right]
<.\displaystyle<\infty.

Moreover, if θUn\theta\in U_{n}, for n0,nn0n\geq 0,n\neq n_{0}, then 𝒫(θ)\mathcal{P}(\theta) is contained in an atom of 𝒬rn\mathcal{Q}_{r_{n}} and

diam 𝒫(θ)rn=e(n+1)<ρ(θ).\displaystyle\text{diam }\mathcal{P}(\theta)\leq r_{n}=e^{-(n+1)}<\rho(\theta).

If θUn0\theta\in U_{n_{0}}^{*}, then 𝒫(θ)\mathcal{P}(\theta) is contained in an atom of 𝒬rn0\mathcal{Q}_{r_{n_{0}}} and

diam 𝒫(θ)rn0=e(n0+1)<ρ(θ).\displaystyle\text{diam }\mathcal{P}(\theta)\leq r_{n_{0}}=e^{-(n_{0}+1)}<\rho(\theta).

In another case, if θVn\theta\in V_{n}, for n1n\geq 1, then 𝒫(θ)\mathcal{P}(\theta) is contained in an atom of 𝒫^n\mathcal{\hat{P}}_{n} and diam𝒫(θ)a=ρ(θ)\text{diam}\mathcal{P}(\theta)\leq a=\rho(\theta). ∎

Given that MM has finite volume, it follows that SMSM also has finite volume. Lemma 6.4, together with the Radon-Nikodym Theorem and Shannon-McMillan-Breiman Theorem, allow us to obtain the following result proved in [11].

Proposition 6.5.

If μν\mu\ll\nu, where ν\nu denotes the Lebesgue measure on SMSM, then

hμ(ϕN)SMhν(ϕN,ρ,θ)𝑑μ(θ).\displaystyle h_{\mu}(\phi^{N})\geq\int\limits_{SM}h_{\nu}(\phi^{N},\rho,\theta)d\mu(\theta).

Proof of Theorem 1.2. We just need to prove that

hμ(ϕ)SM𝒳+(θ)𝑑μ(θ).\displaystyle h_{\mu}(\phi)\geq\int_{SM}\mathcal{X}^{+}(\theta)d\mu(\theta).

Consider Υ=supθSM{dϕθ1,dϕθ1}\displaystyle\Upsilon=\sup_{\theta\in SM}\left\{\left\|d\phi^{1}_{\theta}\right\|,\left\|d\phi^{-1}_{\theta}\right\|\right\}. Then

SMK𝒳+(θ)𝑑μ(θ)\displaystyle\int\limits_{SM\setminus{K^{\prime}}}\mathcal{X}^{+}(\theta)d\mu(\theta) μ(SMK)dim(SM)logΥ\displaystyle\leq\mu(SM\setminus K^{\prime})\cdot\dim(SM)\cdot\log\Upsilon
2εdim(SM)logΥ.\displaystyle\leq 2\sqrt{\varepsilon}\cdot\dim(SM)\cdot\log\Upsilon.

From Propositions 6.2 and 6.5 we have that

hμ(ϕN)\displaystyle h_{\mu}(\phi^{N}) SMhν(ϕN,ρ,θ)𝑑μ(θ)\displaystyle\geq\int_{SM}h_{\nu}(\phi^{N},\rho,\theta)d\mu(\theta)
Khν(ϕN,ρ,θ)𝑑μ(θ)\displaystyle\geq\int_{K^{\prime}}h_{\nu}(\phi^{N},\rho,\theta)d\mu(\theta)
NK𝒳+(θ)𝑑μ(θ)Nεε4NPε\displaystyle\geq N\int_{K^{\prime}}\mathcal{X}^{+}(\theta)d\mu(\theta)-N\varepsilon-\varepsilon-4NP\sqrt{\varepsilon}
NSM𝒳+(θ)𝑑μ(θ)2εNdim(SM)logΥNεε4NPε.\displaystyle\geq N\int_{SM}\mathcal{X}^{+}(\theta)d\mu(\theta)-2\sqrt{\varepsilon}N\cdot\dim(SM)\cdot\log\Upsilon-N\varepsilon-\varepsilon-4NP\sqrt{\varepsilon}.

Hence,

hμ(ϕ)SM𝒳+(θ)𝑑μ(θ)2εdim(SM)logΥεεN4Pε.\displaystyle h_{\mu}(\phi)\geq\int_{SM}\mathcal{X}^{+}(\theta)d\mu(\theta)-2\sqrt{\varepsilon}\cdot\dim(SM)\cdot\log\Upsilon-\varepsilon-\dfrac{\varepsilon}{N}-4P\sqrt{\varepsilon}.

Letting ε0\varepsilon\rightarrow 0 we obtain the desired lower bound.\hfill\square

Acknowledgments

Alexander Cantoral thanks FAPERJ for partially supporting the research (Grant E-26/202.303/2022). Sergio Romaña thanks “Bolsa Jovem Cientista do Nosso Estado No. E-26/201.432/2022”, NNSFC 12071202, and NNSFC 12161141002 from China. The second author thanks the Department of Mathematics of the SUSTech- China for its hospitality during the execution of this work.

References

  • [1] L. Barreira and Y. B. Pesin. Nonuniform hyperbolicity: Dynamics of systems with nonzero Lyapunov exponents, volume 115. Cambridge University Press Cambridge, 2007.
  • [2] J. Bolton. Conditions under which a geodesic flow is Anosov. Math. Ann., 240(2):103–113, 1979.
  • [3] K. Burns, H. Masur, and A. Wilkinson. The Weil-Petersson geodesic flow is ergodic. Ann. of Math. (2), 175(2):835–908, 2012.
  • [4] P. Eberlein. When is a geodesic flow of anosov type? i. Journal of Differential Geometry, 8(3):437–463, 1973.
  • [5] I. C. Ipsen and C. D. Meyer. The angle between complementary subspaces. The American mathematical monthly, 102(10):904–911, 1995.
  • [6] W. Klingenberg. Riemannian manifolds with geodesic flow of Anosov type. Ann. of Math. (2), 99:1–13, 1974.
  • [7] O. Kowalski and M. Sekizawa. On tangent sphere bundles with small or large constant radius. volume 18, pages 207–219. 2000. Special issue in memory of Alfred Gray (1939–1998).
  • [8] J. M. Lee. Introduction to Riemannian manifolds, volume 2. Springer, 2018.
  • [9] G. Liao and N. Qiu. Margulis–ruelle inequality for general manifolds. Ergodic Theory and Dynamical Systems, 42(6):2064–2079, 2022.
  • [10] R. Mañé. On a theorem of Klingenberg. Dynamical systems and bifurcation theory (Rio de Janeiro, 1985), 160:319–345, 1987.
  • [11] R. Mañé. A proof of Pesin’s formula. Ergodic Theory and Dynamical Systems, 1(1):95–102, 1981.
  • [12] Í. Melo and S. Romaña. Riemannian manifolds with Anosov geodesic flow do not have conjugate points. arXiv preprint arXiv:2008.12898, 2020.
  • [13] Í. Melo and S. Romaña. Some rigidity theorems for Anosov geodesic flows in manifolds of finite volume. Qualitative Theory of Dynamical Systems, 23(3):114, 2024.
  • [14] V. I. Oseledec. A multiplicative ergodic theorem. Characteristic Ljapunov, exponents of dynamical systems. Trudy Moskov. Mat. Obšč., 19:179–210, 1968.
  • [15] G. P. Paternain. Geodesic flows, volume 180 of Progress in Mathematics. Birkhäuser Boston, Inc., Boston, MA, 1999.
  • [16] Y. B. Pesin. Characteristic Lyapunov exponents and smooth ergodic theory. Russian Mathematical Surveys, 32(4):55, 1977.
  • [17] F. Riquelme. Counterexamples to ruelle’s inequality in the noncompact case. In Annales de l’institut Fourier, volume 67, pages 23–41, 2017.
  • [18] F. Riquelme. Ruelle’s inequality in negative curvature. Discrete Contin. Dyn. Syst., 38(6):2809–2825, 2018.
  • [19] D. Ruelle. An inequality for the entropy of differentiable maps. Boletim da Sociedade Brasileira de Matemática-Bulletin/Brazilian Mathematical Society, 9(1):83–87, 1978.
  • [20] J.-L. Verger-Gaugry. Covering a ball with smaller equal balls in n\mathbb{R}^{n}. Discrete & Computational Geometry, 33:143–155, 2005.