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

Pressure gaps, Geometric potentials, and nonpositively curved manifolds

Dong Chen Indiana University Indianapolis, Indianapolis, IN 46202, USA [email protected] Lien-Yung Kao George Washington University, Washington, D.C. 20052, USA [email protected]  and  Kiho Park Korea Institute for Advanced Study, Seoul 02455, Republic of Korea [email protected] Dedicated to the memory of Todd Fisher
Abstract.

In this paper, we derive a general pressure gap criterion for closed rank 1 manifolds with singular sets characterized by codimension 1 totally geodesic flat subtori. As an application, we demonstrate that under specific curvature constraints, potentials that decay faster than geometric potentials (towards the singular set) exhibit pressure gaps and lack phase transitions. Additionally, we prove that geometric potentials are Hölder continuous near singular sets.

Parts of this work were done during the third author’s visit to the National Center for Theoretical Sciences (NCTS) in Taiwan. He would like to thank NCTS for their support and hospitality.

1. Introduction.

This paper is dedicated to characterizing pressure gaps in nonuniformly hyperbolic dynamical systems originating from geometric contexts. One distinctive feature of these systems is the absence of pressure gaps. Broadly speaking, pressure gaps can be interpreted as indicating that the "magnitude" of "nonuniform hyperbolicity," as measured by the topological pressure from the perspective of a potential, is small. It has been demonstrated that the presence of a pressure gap is crucial for a potential and its equilibrium states to exhibit ergodic properties similar to those in uniformly hyperbolic systems, including uniqueness, equidistribution, and the Bernoulli property.

The work by Burns, Climenhaga, Fisher, and Thompson in [BCFT18] initiated the study of pressure gaps (see Historical remarks in this section). However, the characterization of pressure gaps remains incomplete. This paper aims to further investigate pressure gaps within natural and concrete geometric contexts.

Let MM be a closed, connected, smooth nn-dimensional manifold, and gg be a Cm+2C^{m+2} (m>0m>0) rank 1 Riemannian metric on MM. Let =(ft)t\mathcal{F}=(f_{t})_{t\in\mathbb{R}} be the geodesic flow on the unit tangent bundle of the Riemannian manifold (M,g)(M,g). The topological pressure P(φ)P(\varphi) of a potential φ\varphi is the supremum of the free energy hμ()+φ𝑑μh_{\mu}({\mathcal{F}})+\int\varphi\,d\mu over {\mathcal{F}}-invariant Borel probability measures, where hμ()h_{\mu}({\mathcal{F}}) is the measure-theoretic entropy. A measure that achieves this supremum is called an equilibrium state of φ\varphi.

For the geodesic flow {\mathcal{F}}, "nonuniform hyperbolicity" arises from a geometric object on T1MT^{1}M, namely, the singular set Sing\mathrm{Sing} (see Section 2 for the precise definition). A potential φ\varphi is said to have a pressure gap if P(Sing,φ)<P(φ)P(\mathrm{Sing},\varphi)<P(\varphi), where P(Sing,φ)P(\mathrm{Sing},\varphi) is the pressure restricted to the {\mathcal{F}}-invariant set Sing\mathrm{Sing}. In this paper, we prove that if φ\varphi decays rapidly enough near the singular set, then φ\varphi has a pressure gap.

Setting

We shall introduce notations and the setup to contextualize our results. Throughout the paper, we assume the following conditions on the Riemannian manifold MM:

  1. (C1)

    MM is a closed rank 1 manifold with nonpositive sectional curvature.

  2. (C2)

    Sing\mathrm{Sing} is either T1T0T^{1}T_{0} or the flat strip case T1T0×[1,1]T^{1}T_{0}\times[-1,1], where T0T_{0} is a codimension 1 totally geodesic flat torus.

  3. (C3)

    MM is negatively curved outside T0T_{0} or the flat strip T0×[1,1]T_{0}\times[-1,1].

For these types of manifolds, it is convenient to study the behavior of geodesics near T0T_{0} or T0×[1,1]T_{0}\times[-1,1] using Fermi coordinates (see Section 2.3 for more details). For pMp\in M, let x(p)x(p) be the signed distance from pp to T0T_{0} or T0×[1,1]T_{0}\times[-1,1], and s(p)s(p) the closest point on T0T_{0} or T0×[1,1]T_{0}\times[-1,1] from pp. The map s:MT0s:M\to T_{0} is a projection onto T0T_{0}, which induces a projection ds:TMTT0ds:TM\to TT_{0}.

For vTpMv\in T_{p}M, we consider two components from its Fermi coordinates: xv=x(p)x_{v}=x(p), and the signed angle ϕv\phi_{v} between vv and the hypersurface x=x(p)x=x(p). We define the radial curvature K(v)K_{\perp}(v) at vv as the sectional curvature of the tangent plane spanned by {v,X}\{v,X\} where X=/xX=\partial/\partial x.

We are particularly interested in the following two types of manifolds:

  • MM is type 1 if MM has an order mm uniform curvature bound, i.e., KK_{\perp} vanishes uniformly to order m1m-1 over T0T_{0} or the flat strip (see (4.2) for more details).

  • MM is type 2 if dimM=2\dim M=2 and MM has an order mm nonuniform curvature bound, i.e., the Gaussian curvature vanishes up to order m1m-1 over T0T_{0} or the flat strip (see (5.1) and (5.2)).

In Examples subsection 1, we discuss manifolds satisfying the above hypotheses in more detail. For example, surfaces of genus greater than one with analytic Riemannian metrics are type 2 manifolds. See Figure 1.1 and Figure 1.2 for some illustrations.

Refer to caption
(a) No flat strips
Refer to caption
(b) With a flat strip
Figure 1.1. Type 1 manifolds
Refer to caption
(a) No flat strips
Refer to caption
(b) With a flat strip
Figure 1.2. Type 2 surfaces

In this paper, in addition to conditions (C1)-(C3), we assume the following condition on the continuous potential φ:T1M\varphi:T^{1}M\to\mathbb{R}:

  1. (C4)

    If MM is type 1, then φ\varphi is constant on Sing\mathrm{Sing}. If MM is type 2, then φ\varphi is transversally constant on Sing\mathrm{Sing}, meaning φ\varphi depends only on the image of the projection SingT1T0\mathrm{Sing}\to T^{1}T_{0}.

Note that when MM is type 2, we do not assume that φ\varphi is constant on Sing\mathrm{Sing}.

Any potential function φ:T1M\varphi:T^{1}M\to\mathbb{R} can be extended to φ:TM\varphi:TM\to\mathbb{R} via

(1.1) φ(v)=|v|φ(v|v|).\varphi(v)=|v|\varphi\left(\frac{v}{|v|}\right).

With this extension, the integral of φ\varphi along any curve is independent of parametrization.

General results for pressure gaps:

We are now ready to state our first result on obtaining the pressure gap. We denote by NR(Sing)N_{R}(\mathrm{Sing}) the radius RR neighborhood of Sing\mathrm{Sing} in T1MT^{1}M.

Theorem A (Pressure gap criterion).

Suppose MM and φ\varphi satisfy conditions (C1)-(C4), and there exist R,C,ε1,ε2>0R,C,\varepsilon_{1},\varepsilon_{2}>0 such that

(1.2) φ(v)φ(ds(v))C(|xv|m2+ε1+|ϕv|mm+2+ε2)\varphi(v)-\varphi(ds(v))\geq-C\left(|x_{v}|^{\frac{m}{2}+\varepsilon_{1}}+|\phi_{v}|^{\frac{m}{m+2}+\varepsilon_{2}}\right)

for any vNR(Sing)v\in N_{R}(\mathrm{Sing}). Then P(Sing,φ)<P(φ)P(\mathrm{Sing},\varphi)<P(\varphi).

In particular, any potential that is Hölder continuous with sufficiently large exponents at Sing\mathrm{Sing} satisfies (1.2). We also note that it is sufficient to have a lower bound on φ(v)φ(ds(v))\varphi(v)-\varphi(ds(v)) near Sing\mathrm{Sing}, since when φ(v)\varphi(v) is larger than φ(ds(v))\varphi(ds(v)), φ\varphi accumulates more pressure outside the singular set. When φ\varphi is locally constant, we achieve [BCFT18, Theorem B] in our setup:

Corollary (Locally constant potentials).

Suppose MM is type 1 or type 2, and φ\varphi is locally constant in a neighborhood of Sing\mathrm{Sing}, then P(φ)>P(Sing,φ)P(\varphi)>P(\mathrm{Sing},\varphi).

Remark 1.1.

We briefly discuss the improvements and differences between Theorem A and [BCFT18, Theorem B]:

  1. (1)

    We do not assume that φ\varphi is constant on Sing\mathrm{Sing} when MM is type 2.

  2. (2)

    One key difference is the construction of shadowing orbits. [BCFT18] uses stable and unstable manifolds to construct orbit segments that shadow those contained in the singular set. This method applies to all nonpositively curved manifolds but provides less control. Our approach requires more precise estimates of the shadowing orbits. To achieve this, we use bouncing orbits (see Figure 4.1) to shadow singular orbits, with curvature bounds providing additional control. See Section 3.1 for more details.

  3. (3)

    Another difference is that the topological entropy of our singular set is zero, allowing us to better estimate P(Sing,φ)P(\mathrm{Sing},\varphi). This is the primary reason we can dispense with the locally constant assumption in [BCFT18, Theorem B]. However, when MM has a flat strip, the transversally constant condition in Theorem A is necessary because increasing φ\varphi in the middle of the strip could eliminate the pressure gap. See [BCFT18, Section 10.1] for more details.

Remark 1.2.

For brevity and readability, we assume MM contains only one flat torus T0T_{0} or one flat strip. However, Theorem A holds when Sing\mathrm{Sing} is induced by finitely many codimension 1 totally geodesic flat tori or flat strips. Specifically, Theorem A is valid under the following assumptions:

  1. (1)

    Sing=i=0k𝖲i\mathrm{Sing}=\coprod_{i=0}^{k}\mathsf{S}_{i}, where 𝖲i=T1Ti\mathsf{S}_{i}=T^{1}T_{i} or T1Ti×[1,1]T^{1}T_{i}\times[-1,1] for a codimension 1 totally geodesic flat subtorus TiT_{i}.

  2. (2)

    Without loss of generality, we assume P(φ|𝖲0)=max{P(φ|𝖲i):i=0,,k}=P(Sing,φ)P(\varphi|_{\mathsf{S}_{0}})=\max\{P(\varphi|_{\mathsf{S}_{i}}):i=0,\ldots,k\}=P(\mathrm{Sing},\varphi). We only need φ\varphi to satisfy the assumptions of Theorem A near 𝖲0\mathsf{S}_{0}.

  3. (3)

    MM satisfies the curvature bounds near T0T_{0} or T0×[1,1]T_{0}\times[-1,1] as a type 1 or type 2 manifold.

We say that a potential φ\varphi has a phase transition at q0q_{0} if the pressure map qP(qφ)q\mapsto P(q\varphi) fails to be differentiable at q0q_{0}. It is well known that the uniqueness of equilibrium states implies the differentiability of the pressure map; see [Rue78].

Our second main result is that no phase transition appears if φ\varphi decays rapidly near Sing\mathrm{Sing}:

Theorem B (No phase transition).

Let MM and φ\varphi satisfy conditions (C1)-(C4), and let φ\varphi be a Hölder continuous potential such that

(1.3) |φ(v)φ(ds(v))|C(|xv|m2+ε1+|ϕv|mm+2+ε2),vNR(Sing).\left|\varphi(v)-\varphi(ds(v))\right|\leq C\left(|x_{v}|^{\frac{m}{2}+\varepsilon_{1}}+|\phi_{v}|^{\frac{m}{m+2}+\varepsilon_{2}}\right),\quad\forall\,v\in N_{R}(\mathrm{Sing}).

Then qφq\varphi has a unique equilibrium state for each qq\in\mathbb{R}; thus, φ\varphi does not have phase transitions.

Remark 1.3.

The argument for Theorem B remains valid even if φ\varphi is only controlled from below, i.e., satisfying (1.2). In this case, the conclusion is that qφq\varphi has a unique equilibrium state for q>0q>0.

The proof of Theorem A relies on an abstract pressure criterion (Proposition 7.2). For readability, we defer the detailed exposition of this abstract result to Section 7. In essence, Proposition 7.2 demonstrates that if the geodesic flow \mathcal{F} satisfies the following conditions: (1) a "strong" specification property, (2) the singular set can be shadowed by nearby vectors, and (3) the potential φ\varphi decays rapidly enough, then φ\varphi exhibits a pressure gap.

Results for Geometric Potentials

The second theme of this paper is dedicated to studying the behavior of the geometric potential φu:T1M\varphi^{u}:T^{1}M\to\mathbb{R} near T0T_{0}. Recall that the geometric potential φu\varphi^{u} is defined as

φu(v):=limt01tlogdet(dftEu(v))\varphi^{u}(v):=-\lim_{t\to 0}\frac{1}{t}\log\det(df_{t}\mid_{E^{u}(v)})

where Eu(v)E^{u}(v) is the unstable subspace (see Section 2 for details).

For type 2 surfaces without flat strips, Gerber and Niţică [GN99, Theorem 3.1] and Gerber and Wilkinson [GW99, Lemma 3.3] provided Hölder continuity estimates for the geometric potential φu\varphi^{u} at Sing\mathrm{Sing}. The following result shows that, under a natural Ricci curvature constraint, similar Hölder continuity estimates can be extended to higher-dimensional cases.

In what follows, we denote by a(v)b(v)a(v)\approx b(v) near SS, if there exists a neighborhood NN of SS and C>1C>1 such that C1b(v)a(v)Cb(v)C^{-1}b(v)\leq a(v)\leq Cb(v) for vNv\in N. We say MM has order mm Ricci curvature bounds if the Ricci curvature Ric(v)\mathrm{Ric}(v) vanishes uniformly to order m1m-1 over T0T_{0} (see (6.3)).

Theorem C (Geometric Potentials).

Let MM be a type 1 manifold without flat strips. Suppose MM has an order mm Ricci curvature bound, then near Sing\mathrm{Sing} we have

φu(v)|xv|m2+|ϕv|mm+2.-\varphi^{u}(v)\approx|x_{v}|^{\frac{m}{2}}+|\phi_{v}|^{\frac{m}{m+2}}.

The no-flat-strip condition is necessary for the Hölder continuity. See Remark 6.5 for more details.

Remark 1.4.

In general, radial curvature and Ricci curvature have no strong relationship. Only in the surface case are these two curvatures the same. Nevertheless, in the appendix, we show that if the Riemannian metric is a warped product, then the Ricci curvature bound and the radial curvature bound hypotheses are equivalent.

In most cases, the Hölder continuity of geometric potentials is unknown for nonuniformly hyperbolic systems, especially in higher dimensions. As an immediate consequence of Theorem C, we have the following partial result for higher-dimensional manifolds:

Theorem D (Local Hölder Continuity).

Under the same assumptions as Theorem C, the geometric potential φu\varphi^{u} is Hölder continuous at Sing\mathrm{Sing}.

We note that our method, inspired by [GW99], currently only establishes the Hölder continuity of the geometric potential φu\varphi^{u} at Sing\mathrm{Sing}. Achieving global Hölder continuity of φu\varphi^{u} may require the Hölder continuity of the unstable Jacobian tensor UuU^{u}, which is still unclear in higher-dimensional cases.

On the other hand, as a consequence of Theorem C, we know that φu\varphi^{u} is a borderline case of the pressure gap criterion given in Theorem A (i.e., ε1=ε2=0\varepsilon_{1}=\varepsilon_{2}=0). Moreover, it is known that for manifolds (including surfaces) whose singular sets are unit tangent bundles of flat, totally geodesic codimension 1 tori, the geometric potential φu\varphi^{u} exhibits a phase transition at q=1q=1 (see [BBFS21, p. 530] and [BCFT18, Theorem C]).

In other words, Theorem C shows that the pressure gap criterion given in Theorem A is optimal in the sense that there are examples at the boundary of our criterion that do not have pressure gaps (see Figure 1.3).

Refer to caption
Figure 1.3. Region where potentials |φ(v)||xv|a+|ϕv|b|\varphi(v)|\approx|x_{v}|^{a}+|\phi_{v}|^{b} exhibit a pressure gap and no phase transitions.

In Figure 1.3, each point corresponds to potentials φ\varphi satisfying φ(v)|xv|a+|ϕv|b-\varphi(v)\approx|x_{v}|^{a}+|\phi_{v}|^{b} near Sing\mathrm{Sing}. The shaded region represents potentials that have a pressure gap and no phase transitions by Theorem A. The geometric potential φu\varphi^{u} lies at the vertex (m2,mm+2)(\frac{m}{2},\frac{m}{m+2}) of the shaded region.

We conclude this subsection by posing an open question:

Question.

Suppose (a,b)(a,b) lies on the boundary of the shaded region in Figure 1.3. Is there a potential φ\varphi satisfying |φ(v)||xv|a+|ϕv|b|\varphi(v)|\approx|x_{v}|^{a}+|\phi_{v}|^{b} near Sing\mathrm{Sing} such that φ\varphi has a phase transition?

Examples:

The prototype of type 1 manifolds is the surface of revolution with profile f(x)=1+|x|rf(x)=1+|x|^{r}, which is the main example discussed in Lima, Matheus, and Melbourne [LMM]. For higher dimensions, an important example is the Heintze example (see Ballman, Brin, and Eberlein [BBE85] or [BCFT18, Section 10.2]). The simplest version of the Heintze example starts with a finite volume hyperbolic 3-manifold with one cusp, then removes the cusp and flattens the region near the cross-section. Recall that the cross-section of the cusp is a codimension 1 totally geodesic flat torus. The Heintze example is obtained by gluing two identical copies of the above 3-manifold along the cross-section (see Figure 1.1).

Type 2 surfaces (see Figure 1.2) were introduced in Gerber and Niţică [GN99] and Gerber and Wilkinson [GW99]. The archetype is a rank 1 nonpositively curved surface with an analytic metric. In such cases, it is well-known that Sing\mathrm{Sing} consists of unit tangent bundles of finitely many closed geodesics (see [BCFT18, Section 10.1] for a sketched proof).

We remark that for nonpositively curved surfaces, Coudène and Schapira [CS14, Theorem 3.2] (inspired by the unpublished work of Cao and Xavier [CX08]) showed that flat strips for nonpositively curved manifolds close up. However, in general, it is unknown if the singular geodesics or the (higher dimensional) zero curvature strips close up. Nevertheless, in all known examples, to the best of the authors’ knowledge, the singular sets do close up, leading to our hypothesis on the existence of T0T_{0}.

Historical Remarks

There is no singular set when the dynamical system is uniformly hyperbolic. Hence, the pressure gap persists for a broad class of potentials, allowing one to derive ergodic properties for associated equilibrium states. The origin of this fact traces back to the work of Bowen [Bow74] for maps and Franco [Fra77] for flows. For nonuniformly hyperbolic systems, Climenhaga and Thompson [CT16] proposed using the pressure gap as a condition to obtain ergodic properties of equilibrium states, particularly uniqueness, similar to uniform hyperbolic cases.

Burns et al. [BCFT18] applied the argument from [CT16] and derived a necessary condition for the pressure gap. Specifically, they showed that for closed rank 1 nonpositively curved manifolds, if φ\varphi is locally constant near Sing\mathrm{Sing}, then φ\varphi has a pressure gap. This work was inspired by Knieper [Kni98], where the entropy gap htop(Sing)<htop()h_{\mathrm{top}}(\mathrm{Sing})<h_{\mathrm{top}}(\mathcal{F}) was established as a consequence of the uniqueness of the measure of maximal entropy. Recall that the topological entropy htop()h_{\mathrm{top}}(\mathcal{F}) is the pressure of the zero potential.

Gelfert and Schapira [GS14] compared different notions of pressure for closed rank 1 nonpositively curved manifolds, such as topological pressure, Gurevich pressure (or periodic orbit pressure), and their restrictions on singular and regular sets. They pointed out that under certain conditions, these different notions of pressure are identical. Similar discussions can be found in [BCFT18, Propositions 2.8 and 6.4].

In geometry, the Liouville measure is an equilibrium state for the geometric potential. Ergodic properties of equilibrium states have been extensively studied. Several recent contributions have employed the Climenhaga-Thompson strategy (see [CT21] for a survey on the strategy). For example, Chen, Kao, and Park [CKP20, CKP21] worked on no focal points settings; Climenhaga, Knieper, and War focused on no conjugate points manifolds [CKW21]; Call, Constantine, Erchenko, and Sawyer [CCE+] discussed flat surfaces with singularities.

There are many other relevant discussions on the ergodic properties of equilibrium states. For example, uniqueness is discussed in [GR19], the Kolmogorov property in [CT22], the Bernoulli property in [Pes77, BG89, OW98, LLS16, CT22, ALP], and the central limit theorem in [TW21, LMM].

For geometric potentials of rank 1 surfaces, Burns and Gelfert [BG14] pointed out the existence of a phase transition. This was also confirmed in [BCFT18] using a different approach. A recent work by Burns, Buzzi, Fisher, and Sawyer [BBFS21] further investigated the edge case of φ=qφu\varphi=q\varphi^{u} for q=1q=1. They showed that the Liouville measure is the only equilibrium state not supported on Sing\mathrm{Sing}. However, the Hölder continuity of φu\varphi^{u} is even less known. It is only assured by Gerber and Wilkinson [GW99, Theorem I] for type 2 surfaces. Much less is known about φu\varphi^{u} in higher dimensional cases.

Outline of the Paper

In Section 2, we recall some relevant background material from geometry and dynamics. In Section 3, we introduce the main shadowing technique and a key inequality to prove Theorem A. Sections 4 and 5 are devoted to technical estimates in type 1 and type 2 settings by analyzing the relevant Riccati equations. Section 6 focuses on the geometric potential and the proof of Theorem C. The proof of Theorem A is presented in Section 7 as a consequence of a more general pressure gap criterion, Proposition 7.2. The proof draws inspiration from [BCFT18, Theorem B]. However, our specific setup allows us to circumvent several technicalities and arrive at a more straightforward proof than that presented in [BCFT18]. In the appendix, we show that for warped product metrics, radial curvature bounds and Ricci curvature bounds are equivalent, and we provide a proof of Peres’ lemma for flows.

Acknowledgments

The authors are grateful to Amie Wilkinson for proposing this question, to Jairo Bochi for bringing Peres’ lemma to our attention, and to Keith Burns, Vaughn Climenhaga, Todd Fisher, and Dan Thompson for enlightening discussions. We also thank the referees for their helpful comments in improving this work. Lastly, the authors dedicate this work to Todd Fisher for the inspiration he brought us in his too-short but luminous life.

2. Preliminary

2.1. Geometry of nonpositively curved manifolds

This subsection will survey relevant geometric features of nonpositively curved manifolds. A more comprehensive survey of these results can be found in [Bal95, Ebe01].

Let MM be a closed nonpositively curved manifold, and {ft}t\{f_{t}\}_{t\in\mathbb{R}} the geodesic flow on the unit tangent bundle T1MT^{1}M. The tangent bundle TT1MTT^{1}M contains three dftdf_{t}-invariant continuous bundles Es,EuE^{s},E^{u}, and EcE^{c}. The bundle EcE^{c} is one-dimensional along the flow direction, and the other two bundles Es/uE^{s/u}, which are orthogonal to EcE^{c} with respect to the Sasaki metric, can be described using Jacobi fields. If MM is negatively curved, these three bundles form a splitting of TT1MTT^{1}M.

A Jacobi field JJ along a geodesic γ\gamma is a vector field along γ\gamma satisfying the Jacobi equation

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

where RR is a Riemannian curvature tensor. A Jacobi field JJ is orthogonal if there exists t0t_{0}\in\mathbb{R} such that J(t0)J(t_{0}) and J(t)J^{\prime}(t) are perpendicular to γ(t0)\gamma^{\prime}(t_{0}). It is well known that when this orthogonal property holds at some t0t_{0}\in\mathbb{R}, then it holds for all tt\in\mathbb{R}. A Jacobi field JJ is parallel if J(t)=0J^{\prime}(t)=0 for all tt\in\mathbb{R}.

Denoting the space of orthogonal Jacobi fields along γ\gamma by 𝒥(γ)\mathcal{J}^{\perp}(\gamma), we can identify TvT1MT_{v}T^{1}M with 𝒥(γv)\mathcal{J}^{\perp}(\gamma_{v}) as follows. Consider a vector vTpMv\in T_{p}M. Using the Levi-Civita connection the tangent space TvTMT_{v}TM at vv may be identified with the direct sum HvVvH_{v}\oplus V_{v} of horizontal and vertical subspace, respectively, each isomorphic to TpMT_{p}M equipped with the norm induced from the Riemannian metric on MM. The Sasaki metric on TMTM is defined by declaring HvH_{v} and VvV_{v} to be orthogonal. Restricted to T1MT^{1}M, the tangent space TvT1MT_{v}T^{1}M corresponds to HvvH_{v}\oplus v^{\perp} under this identification. Then any vector ξTvT1M\xi\in T_{v}T^{1}M for vT1Mv\in T^{1}M may be written as (ξh,ξv)(\xi_{h},\xi_{v}) and can be identified with an orthogonal Jacobi field Jξ𝒥(γv)J_{\xi}\in\mathcal{J}^{\perp}(\gamma_{v}) along γv\gamma_{v} with the initial conditions (Jξ(0),Jξ(0))=(ξh,ξv)(J_{\xi}(0),J^{\prime}_{\xi}(0))=(\xi_{h},\xi_{v}). Moreover, the Sasaki norm of dft(ξ)df_{t}(\xi) satisfies

dft(ξ)2=Jξ(t)2+Jξ(t)2.\|df_{t}(\xi)\|^{2}=\|J_{\xi}(t)\|^{2}+\|J^{\prime}_{\xi}(t)\|^{2}.

The stable subspace Es(v)E^{s}(v) at vT1Mv\in T^{1}M is then defined as

Es(v):={ξTvT1M:Jξ(t) is bounded for t0}.E^{s}(v):=\{\xi\in T_{v}T^{1}M\colon\|J_{\xi}(t)\|\text{ is bounded for }t\geq 0\}.

Similarly, the unstable subspace Eu(v)E^{u}(v) consists of vectors ξTvT1M\xi\in T_{v}T^{1}M where Jξ(t)\|J_{\xi}(t)\| is bounded for t0t\leq 0. Notice that the subbundles Eu(v)E^{u}(v) and Es(v)E^{s}(v) are integrable to the respective foliations Wu(v)T1MW^{u}(v)\subset T^{1}M and Ws(v)T1MW^{s}(v)\subset T^{1}M. The footprints of Wu(v)W^{u}(v) and Ws(v)W^{s}(v) on MM are called the unstable and stable horospheres, which are denoted by Hu(v)H^{u}(v) and Hs(v)H^{s}(v), respectively.

The rank of a vector vT1Mv\in T^{1}M is the dimension of the space of parallel Jacobi fields along γv\gamma_{v}, which coincides with the number 1+dim(EsEu)1+\dim(E^{s}\cap E^{u}). We say the manifold is rank 1 if it has at least one rank 1 vector. This paper will focus mainly on closed rank 1 nonpositively curved manifolds.

The singular set is a set of vectors on which the geodesic flow fails to display uniform hyperbolicity, and it is defined by

Sing:={vT1M:Es(v)Eu(v)0}.\mathrm{Sing}:=\{v\in T^{1}M\colon E^{s}(v)\cap E^{u}(v)\neq 0\}.

The singular set is closed and \mathcal{F}-invariant, and in the case where MM is a surface the singular set can be characterized as the set of vectors vv where the Gaussian curvature K(γv(t))K(\gamma_{v}(t)) vanishes for all tt\in\mathbb{R}. The complement of the singular set is the regular set

Reg:=T1MSing.\mathrm{Reg}:=T^{1}M\setminus\mathrm{Sing}.

The geodesic flow restricted to the regular set is hyperbolic, but the degree of hyperbolicity, which can be measured by the function

λ(v):=min(λu(v),λs(v))\lambda(v):=\min(\lambda^{u}(v),\lambda^{s}(v))

where λu(v)\lambda^{u}(v) is the minimum eigenvalue of the shape operator on the unstable horosphere Hu(v)H^{u}(v) at vv. Using λ\lambda we can define nested compact subsets {Reg(η)}η>0\{\mathrm{Reg}(\eta)\}_{\eta>0} of Reg\mathrm{Reg} where

Reg(η):={vReg:λ(v)η}.\mathrm{Reg}(\eta):=\{v\in\mathrm{Reg}\colon\lambda(v)\geq\eta\}.

These subsets may be viewed as uniformity blocks in the sense of Pesin’s theory, where the hyperbolicity is uniform. More details and properties of the function λ\lambda can be found in [BCFT18].

The geometric potential φu:T1M\varphi^{u}:T^{1}M\to\mathbb{R} is an important potential that measures the infinitesimal volume growth in the unstable direction:

φu(v):=limt01tlogdet(dftEu(v))=ddt|t=0logdet(dftEu(v)).\varphi^{u}(v):=-\lim_{t\to 0}\frac{1}{t}\log\det(df_{t}\mid_{E^{u}(v)})=-\frac{d}{dt}\Big{|}_{t=0}\log\det(df_{t}\mid_{E^{u}(v)}).

In order to study the geometric potential, it is convenient to study Riccati equations. Interestingly, the shape operator of unstable (and unstable) horosphere is a solution of a Riccati equation. To see this, we start by introducing terminologies.

Let HMH\subset M be a hypersurface orthogonal to γv\gamma_{v} at πv\pi v where π:T1MM\pi:T^{1}M\to M is the canonical projection. An orthogonal Jacobi field J𝒥(γv)J\in\mathcal{J}^{\perp}(\gamma_{v}) is called a HH-Jacobi field along γv\gamma_{v}, if JJ comes from varying γv\gamma_{v} through unit speed geodesics orthogonal to HH. We denote the set of HH- Jacobi fields by 𝒥H(γv)\mathcal{J}_{H}(\gamma_{v}) . The shape operator on HH is the symmetric linear operator U:TπvHTπvHU:T_{\pi v}H\to T_{\pi v}H defined by U(v)=vNU(v)=\nabla_{v}N, where NN is the unit normal vector field toward the same side as vv.

We are particularly interested in the unstable horosphere H=Hu(v)H=H^{u}(v) at vv. In this case, JHu(γv)J_{H^{u}}(\gamma_{v}) coincides with the space of unstable Jacobi field Ju(γv)J^{u}(\gamma_{v}). For tt\in\mathbb{R}, let Uvu(t):TπvHu(ftv)TπvHu(ftv)U_{v}^{u}(t):T_{\pi v}H^{u}(f_{t}v)\to T_{\pi v}H^{u}(f_{t}v) be the shape operator of the unstable horosphere Hu(ftv)H^{u}(f_{t}v). We know Uvu(t)U_{v}^{u}(t) is a a positive semidefinite symmetric linear operator on (ftv)(f_{t}v)^{\perp}, and for any unstable Jacobi field J(t)J(t) it satisfies J(t)=Uvu(t)J(t)J^{\prime}(t)=U_{v}^{u}(t)J(t); see [BCFT18, Lemma 2.9].

For any vector vT1Mv\in T^{1}M, let 𝒦(v):vv\mathcal{K}(v):v^{\perp}\to v^{\perp} is the symmetric linear map defined via 𝒦(v)X,Y:=R(X,v)v,Y\langle\mathcal{K}(v)X,Y\rangle:=\langle R(X,v)v,Y\rangle for X,YvX,Y\in v^{\perp}. Using the Jacobi equation, for an unstable Jacobi field J(t)J(t) we know J′′(t)+K(ftv)J(t)=0J^{\prime\prime}(t)+K(f_{t}v)J(t)=0 and J(t)=Uvu(t)J(t)J^{\prime}(t)=U_{v}^{u}(t)J(t). Thus, we get the operator-valued Riccati equation:

(2.1) (Uvu)(t)+Uvu(t)2+𝒦(γ˙v(t))=0;(U_{v}^{u})^{\prime}(t)+U_{v}^{u}(t)^{2}+{\mathcal{K}}(\dot{\gamma}_{v}(t))=0;

see [BCFT18, (7.6)]. Using the above notation, the Ricci curvature Ric(v){\rm Ric}(v) at vv is defined as the trace of the map 𝒦(v){\mathcal{K}}(v).

2.2. Thermodynamic formalism

We now briefly survey relevant results in thermodynamic formalism. The general notion of topological entropy and pressure described in the following can be defined for an arbitrary flow ={ft}t\mathcal{F}=\{f_{t}\}_{t\in\mathbb{R}} in a compact metric space (X,d)(X,d).

For any t>0t>0, we define a metric dtd_{t} on XX via

dt(x,y):=max0τtd(ftx,fty),d_{t}(x,y):=\max\limits_{0\leq\tau\leq t}d(f_{t}x,f_{t}y),

and the corresponding δ\delta-ball around xXx\in X in dtd_{t}-metric will be denoted by Bt(x,δ)B_{t}(x,\delta). We say a subset EE of XX is (t,δ)(t,\delta)-separated if dt(x,y)δd_{t}(x,y)\geq\delta for distinct x,yEx,y\in E. Moreover, we will identify (x,t)X×[0,+)(x,t)\in X\times[0,\mathbb{R}^{+}) with the orbit segment of length tt starting at xx.

Let φ:X\varphi\colon X\to\mathbb{R} be a continuous function on XX, which we often call a potential. We define Φ(x,t):=0tφ(fτx)𝑑τ{\displaystyle\Phi(x,t):=\int_{0}^{t}\varphi(f_{\tau}x)\,d\tau} to be the integral of φ\varphi along an orbit segment (x,t)(x,t). For any subset 𝒞X×[0,+)\mathcal{C}\in X\times[0,\mathbb{R}^{+}), we let 𝒞t\mathcal{C}_{t} be the subset of 𝒞\mathcal{C} consisting of orbit segments of length tt. We define

Λ(𝒞,φ,δ,t)=sup{xEeΦ(x,t):E𝒞t is (t,δ)separated}.\Lambda(\mathcal{C},\varphi,\delta,t)=\sup\Big{\{}\sum\limits_{x\in E}e^{\Phi(x,t)}\colon E\subset\mathcal{C}_{t}\text{ is }(t,\delta)-\text{separated}\Big{\}}.

The topological pressure of φ\varphi on 𝒞\mathcal{C} is then defined by

P(𝒞,φ):=limδ0lim supt1tlogΛ(𝒞,φ,δ,t).P(\mathcal{C},\varphi):=\lim\limits_{\delta\to 0}\limsup\limits_{t\to\infty}\frac{1}{t}\log\Lambda(\mathcal{C},\varphi,\delta,t).

When 𝒞\mathcal{C} is the entire orbit space X×[0,+)X\times[0,\mathbb{R}^{+}), then we denote it by P(φ)P(\varphi) and call it the topological pressure of φ\varphi. In the case where φ0\varphi\equiv 0, the resulting pressure P(0)P(0) is called the topological entropy of the flow \mathcal{F} denoted by htop()h_{top}(\mathcal{F}).

Denoting by ()\mathcal{M}(\mathcal{F}) the set of all \mathcal{F}-invariant measures on XX, the pressure P(φ)P(\varphi) satisfies the variational principle

P(φ)=supμ(){hμ()+φ𝑑μ}P(\varphi)=\sup\limits_{\mu\in\mathcal{M}(\mathcal{F})}\Big{\{}h_{\mu}(\mathcal{F})+\int\varphi\,d\mu\Big{\}}

where hμ()h_{\mu}(\mathcal{F}) is the measure-theoretic entropy of μ\mu. Any invariant measure μ()\mu\in\mathcal{M}(\mathcal{F}) attaining the supremum is called an equilibrium state for φ\varphi. Likewise, any invariant measure attaining the supremum when φ0\varphi\equiv 0 is called a measure of maximal entropy.

2.3. Codimension 1 totally geodesic flat torus and Fermi coordinates

Let MM be an nn-dimensional closed rank 1 nonpositively curved manifold and T0T_{0} a totally geodesic (n1)(n-1)-torus in MM with K0K\equiv 0 on any xT0x\in T_{0}. We further suppose that the complement of T0T_{0} is negatively curved and that curvature away from a small neighborhood of T0T_{0} admits a uniform upper bound strictly smaller than 0. A more precise control of the curvature of the neighborhood will be specified later, depending on the setting under consideration.

In what follows, we fix a fundamental domain in M~\widetilde{M} the universal covering of MM and (abusing the notation) continue denoting the lifts of pMp\in M and vTpMv\in T_{p}M by pp and vv, respectively. Recall that the Fermi coordinate of pp is given by (s,x)(s,x) where ss is an (n1)(n-1)-dimensional coordinate on T~0\widetilde{T}_{0} and xx measures the signed distance on M~\widetilde{M} to T~0\widetilde{T}_{0}.

For pM~p\in\widetilde{M} near T~0\widetilde{T}_{0}, by x(p)x(p) we mean the xx-coordinate of pp. For any vTpM~v\in T_{p}\widetilde{M} with pp near T~0\widetilde{T}_{0}, we define xv:=x(p)x_{v}:=x(p) and denote by ϕv\phi_{v} the signed angle between vv and the hypersurface x=x(p)x=x(p); we adopt the convention that ϕX=π/2\phi_{X}=\pi/2 when X=/xX=\partial/\partial x. We also define

(2.2) xv(τ):=x(γv(τ)) and ϕv(τ):=ϕ(γv(τ)).x_{v}(\tau):=x(\gamma_{v}(\tau))\text{ and }\phi_{v}(\tau):=\phi(\gamma_{v}^{\prime}(\tau)).

When there is no confusion, we may write xvx_{v} and ϕv\phi_{v} for xv(0)x_{v}(0) and ϕv(0)\phi_{v}(0), respectively.

Remark 2.1.

With respect to Fermi coordinates (s,x)(s,x), the curve s=const.s=\text{const.} is always a geodesic perpendicular to T~0\widetilde{T}_{0}, while (s(t),x(t))(s(t),x(t)) with x(t)x0x(t)\equiv x_{0} for some x0x_{0} is not a geodesic unless x0=0x_{0}=0 and s(t)s(t) is linear.

For ε>0\varepsilon>0 small, the Riemannian metric near T~0\widetilde{T}_{0} can be written as

(2.3) g=dx2+gx,|x|εg=dx^{2}+g_{x},\quad|x|\leq\varepsilon

where gxg_{x} is the Riemannian metric on T~x:=T~0×{x}\widetilde{T}_{x}:=\widetilde{T}_{0}\times\{x\}. In particular, g0g_{0} is the Euclidean metric on T~0\widetilde{T}_{0}.

Denoting by X:=/xX:=\partial/\partial x the vertical vector field, the second fundamental form on T~x\widetilde{T}_{x} is defined via

II(v,w):=vX,w,\text{II}(v,w):=\langle\nabla_{v}X,w\rangle,

for any v,wT(s,x)T~xv,w\in T_{(s,x)}\widetilde{T}_{x}. The shape operator U(s,x):T(s,x)T~xT(s,x)T~xU(s,x):T_{(s,x)}\widetilde{T}_{x}\to T_{(s,x)}\widetilde{T}_{x} is defined via

II(v,w)=U(s,x)v,w.\text{II}(v,w)=\langle U(s,x)v,w\rangle.

As II is bilinear and symmetric, the shape operator U(s,x)U(s,x) is diagonalizable. Its eigenvalues

λ1(s,x)λ2(s,x)λn1(s,x)\lambda_{1}(s,x)\leq\lambda_{2}(s,x)\leq\cdots\leq\lambda_{n-1}(s,x)

are called principal curvatures at (s,x)(s,x).

For any geodesic γ(t)=(s(t),x(t))\gamma(t)=(s(t),x(t)) near T~0\widetilde{T}_{0}, by the first variation formula, we have

(2.4) x=sinϕ,x^{\prime}=\sin\phi,

where ϕ(t):=ϕ(γ(t))\phi(t):=\phi(\gamma^{\prime}(t)), which then gives x′′=ϕcosϕx^{\prime\prime}=\phi^{\prime}\cos\phi.

We denote by

γ(t):=γ(t)γ(t),XX\gamma^{\prime}_{\perp}(t):=\gamma^{\prime}(t)-\langle\gamma^{\prime}(t),X\rangle X

the component of γ(t)\gamma^{\prime}(t) that is orthogonal to XX. Then |γ|=cosϕ|\gamma^{\prime}_{\perp}|=\cos\phi.

Lemma 2.2.

If γ(t)=(s(t),x(t))\gamma(t)=(s(t),x(t)) is a geodesic on M~\widetilde{M} near T~0\widetilde{T}_{0}, we have

x′′=II(γ,γ)[λ1(s,x)cos2ϕ,λn1(s,x)cos2ϕ].x^{\prime\prime}=\text{II}(\gamma^{\prime}_{\perp},\gamma^{\prime}_{\perp})\in[\lambda_{1}(s,x)\cos^{2}\phi,\lambda_{n-1}(s,x)\cos^{2}\phi].
Proof.

By (2.4), we have x=sinϕ=X,γ(t)x^{\prime}=\sin\phi=\langle X,\gamma^{\prime}(t)\rangle. Thus

x′′=ddtX,γ(t)=γ(t)X,γ(t)=γ(t)X,γ(t)=II(γ,γ).x^{\prime\prime}=\frac{d}{dt}\langle X,\gamma^{\prime}(t)\rangle=\langle\nabla_{\gamma^{\prime}(t)}X,\gamma^{\prime}(t)\rangle=\langle\nabla_{\gamma^{\prime}_{\perp}(t)}X,\gamma^{\prime}_{\perp}(t)\rangle=\text{II}(\gamma^{\prime}_{\perp},\gamma^{\prime}_{\perp}).

Note that the third equality uses the fact XX=0\nabla_{X}X=0; see Remark 2.1. Since |γ|=cosϕ|\gamma^{\prime}_{\perp}|=\cos\phi, we have

II(γ,γ)λ1(s,x)|γ|2=λ1(s,x)cos2ϕ\text{II}(\gamma^{\prime}_{\perp},\gamma^{\prime}_{\perp})\geq\lambda_{1}(s,x)|\gamma^{\prime}_{\perp}|^{2}=\lambda_{1}(s,x)\cos^{2}\phi

and

II(γ,γ)λn1(s,x)cos2ϕ.\text{II}(\gamma^{\prime}_{\perp},\gamma^{\prime}_{\perp})\leq\lambda_{n-1}(s,x)\cos^{2}\phi.

This completes the proof. ∎

Remark 2.3.

When MM is a surface, then T0T_{0} is a closed geodesic. In this case the Riemannian metric (2.3) near T~0\widetilde{T}_{0} may be written as

(2.5) g=dx2+G(s,x)2ds2g=dx^{2}+G(s,x)^{2}ds^{2}

with G(s,0)1G(s,0)\equiv 1. The Gaussian curvature is given by K=Gxx/GK=-G_{xx}/G, and the second fundamental form at T~x\widetilde{T}_{x} is Gx/GG_{x}/G. In particular, by Lemma 2.2, x′′x^{\prime\prime} admits the following expression

(2.6) x′′=λ(s,x)cos2ϕ=GxGcos2ϕ.x^{\prime\prime}=\lambda(s,x)\cos^{2}\phi=\frac{G_{x}}{G}\cos^{2}\phi.

3. Preparation and outline for pressure gaps results

3.1. Shadowing map

To distinguish vectors in Sing\mathrm{Sing} and generic vectors in T1MT^{1}M, we will use different fonts to denote them; more precisely, we will write 𝗏Sing\mathsf{v}\in\mathrm{Sing} and vT1Mv\in T^{1}M. Given an orbit segment (𝗏,t)T1M×+(\mathsf{v},t)\in T^{1}M\times\mathbb{R}^{+} with 𝗏Sing\mathsf{v}\in\mathrm{Sing}, we now describe a method for constructing a new orbit segment that shadows (𝗏,t)(\mathsf{v},t). Though simple, this construction will be crucial in proving Theorem A and Theorem D.

For any 𝗏Sing\mathsf{v}\in\text{Sing}, suppose π(𝗏)=(s0,0)\pi(\mathsf{v})=(s_{0},0) for some s0s_{0}. For any t>0t>0 and any R>0R>0 such that the Fermi coordinates (s,x)(s,x) are well-defined for |x|<R|x|<R, there exists s1s_{1}\in\mathbb{R} such that the distance on M~\widetilde{M} between (s0,R)(s_{0},R) and ((sπ)(fs1𝗏),R)((s\circ\pi)(f_{s_{1}}\mathsf{v}),R) is equal to tt; see Figure 3.1. From the triangle inequality we know that |ts1|<2R|t-s_{1}|<2R.

Denoting by γ\gamma the geodesic connecting these two points, we define

(3.1) Πt,R(𝗏):=γ(0).\Pi_{t,R}(\mathsf{v}):=\gamma^{\prime}(0).

Throughout the paper, we will often write w𝗏w_{\mathsf{v}}, or simply ww, to denote Πt,R(𝗏)\Pi_{t,R}(\mathsf{v}) whenever the context is clear.

Refer to caption
Figure 3.1. Construction of Πt,R\Pi_{t,R}

The next few lemmas establish a few properties on the map Πt,R\Pi_{t,R}.

Lemma 3.1.

For any 𝗏Sing\mathsf{v}\in\mathrm{Sing} and t>0t>0, the following statements hold:

  1. (1)

    For any R>0R>0, the function xw(τ)x_{w}(\tau) is convex for any τ[0,t]\tau\in[0,t].

  2. (2)

    There exists R0>0R_{0}>0 such that for any R(0,R0)R\in(0,R_{0}),

    |ϕw(τ)|<π4|\phi_{w}(\tau)|<\frac{\pi}{4}

    for all τ[0,t]\tau\in[0,t].

  3. (3)

    For any R>0R>0, there exists η=η(R)>0\eta=\eta(R)>0 such that w,ftwReg(η)w,f_{t}w\in\text{Reg}(\eta).

Proof.

For (1), it is not hard to see from the construction of Πt,R\Pi_{t,R} that |ϕw(τ)|π/2|\phi_{w}(\tau)|\leq\pi/2 and ϕw(τ)0\phi_{w}^{\prime}(\tau)\geq 0 for all τ[0,t]\tau\in[0,t]. The statement then is a consequence of xw′′(τ)=ϕw(τ)cosϕw(τ)x_{w}^{\prime\prime}(\tau)=\phi_{w}^{\prime}(\tau)\cos\phi_{w}(\tau) from (2.4).

For (2), it is clear that

|ϕw(τ)|max{|ϕw(0)|,|ϕw(t)|}|\phi_{w}(\tau)|\leq\max\{|\phi_{w}(0)|,|\phi_{w}(t)|\}

for all τ[0,t]\tau\in[0,t]. We then observe that |ϕw(0)||ϕ(γw,s(0))||\phi_{w}(0)|\leq|\phi(\gamma^{\prime}_{w,s}(0))| where γw,s\gamma_{w,s} is the geodesic with the same initial point as γw\gamma_{w} that is forward asymptotic to T~0\widetilde{T}_{0}. The statement then follows as the function

Rsup{|ϕ(u)|:uWs(𝗏) for some 𝗏Sing and x(πu)=R}R\mapsto\sup\{|\phi(u)|\colon u\in W^{s}(\mathsf{v})\text{ for some }\mathsf{v}\in\mathrm{Sing}\text{ and }x(\pi u)=R\}

vanishes when R=0R=0 and varies continuously in RR. By repeating the same argument with the unstable manifolds Wu(𝗏)W^{u}(\mathsf{v}) for 𝗏Sing\mathsf{v}\in\mathrm{Sing}, we can find R0>0R_{0}>0 with the desired property.

For (3), any unit vector vT1Mv\in T^{1}M satisfies x(π(v))0x(\pi(v))\neq 0 belongs to Reg\mathrm{Reg}. Since Reg\mathrm{Reg} is exhausted by compact subsets Reg(η)\mathrm{Reg}(\eta) and the flat torus T0T_{0} in MM is compact, the statement follows. ∎

Remark 3.2.

From here on, we will assume that RR belongs to [0,min(R0,R1)][0,\min(R_{0},R_{1})] with R0,R1R_{0},R_{1} defined as in above lemmas. In particular, we will often evoke Lemma 3.1 to use the inequality

cosϕw(τ)[22,1]\cos\phi_{w}(\tau)\leq\Big{[}\frac{\sqrt{2}}{2},1\Big{]}

for any w=Πt,R(𝗏)w=\Pi_{t,R}(\mathsf{v}) and τ[0,t]\tau\in[0,t].

3.2. Key inequality and the outline of Theorem A

In this subsection, we provide a brief outline of what consists of the remaining sections. Recall from Lemma 3.1 that xw(τ)x_{w}(\tau) is a convex function on τ[0,t]\tau\in[0,t] for any w=Πt,R(𝗏)w=\Pi_{t,R}(\mathsf{v}), and hence there exists a well-defined number t~[0,t]\tilde{t}\in[0,t] such that

(3.2) xw(t~)=minτ[0,t]xw(τ)x_{w}(\tilde{t})=\min_{\tau\in[0,t]}x_{w}(\tau)

and that t~\tilde{t} is the smallest among all such numbers. In the case where xw(τ)x_{w}(\tau) is strictly convex, there is a unique t~\tilde{t} which attains the minimum of xw(τ)x_{w}(\tau). Note that xw(τ)0x^{\prime}_{w}(\tau)\leq 0 and ϕw(τ)0\phi_{w}(\tau)\leq 0 for τ[0,t~]\tau\in[0,\tilde{t}\,].

The goal of the next few sections is to establish bounds on the distance xw(τ)x_{w}(\tau) and the angle ϕw(τ)\phi_{w}(\tau) under the assumption that the radial curvature KK_{\perp} vanishes to the order of m1m-1 at T0T_{0} and that the curvature near T0T_{0} is controlled; see Section 4 and 5 for the precise description of the setting. In particular, we will show that for suitable R>0R>0, there exists Q=Q(R)>1Q=Q(R)>1 independent of tt such that the shadowing vector w=Πt,R(𝗏)w=\Pi_{t,R}(\mathsf{v}) for any 𝗏Sing\mathsf{v}\in\mathrm{Sing} and any t>0t>0 satisfies

(3.3) Q1(τ+1)2mxw(τ)Q(τ+1)2m,|ϕw(τ)|Q[(τ+1)m+2m(t~+1)m+2m]\displaystyle\begin{split}Q^{-1}(\tau+1)^{-\frac{2}{m}}\leq&x_{w}(\tau)\leq Q(\tau+1)^{-\frac{2}{m}},\\ &|\phi_{w}(\tau)|\leq Q[(\tau+1)^{-\frac{m+2}{m}}-(\tilde{t}+1)^{-\frac{m+2}{m}}]\end{split}

for any τ[0,t~]\tau\in[0,\tilde{t}\,]. From its derivation in Proposition 4.2 and 5.3, it will be clear that the analogous inequality holds for τ[t~,t]\tau\in[\tilde{t},t] by simply applying the symmetric argument starting from τ=t\tau=t instead of τ=0\tau=0:

(3.4) Q1(tτ+1)2mxw(τ)Q(tτ+1)2m,|ϕw(τ)|Q[(tτ+1)m+2m(t~+1)m+2m].\displaystyle\begin{split}Q^{-1}(t-\tau+1)^{-\frac{2}{m}}\leq&x_{w}(\tau)\leq Q(t-\tau+1)^{-\frac{2}{m}},\\ &|\phi_{w}(\tau)|\leq Q[(t-\tau+1)^{-\frac{m+2}{m}}-(\tilde{t}+1)^{-\frac{m+2}{m}}].\end{split}

In the setting considered in Section 4 where a uniform control on the curvature KK_{\perp} is assumed on the entire T~0\widetilde{T}_{0}, the angle |ϕw(τ)||\phi_{w}(\tau)| from (3.3) and (3.4) admits the corresponding lower bound also; see Proposition 4.2.

In the context considered by Gerber and Wilkinson [GW99] where MM is a surface (see Section 5 for details) fits into the assumption of KK_{\perp} described in the above paragraph, and φ\varphi satisfying (1.2) is related to the geometric potential, as elaborated in more detail in the next subsection.

Assuming the estimates (3.3) and (3.4), we now derive useful consequences from them when considering potentials φ\varphi satisfying (1.2). We will see in Section 7 that, together with certain properties of the geodesic flow, these results serve as sufficient criteria for the potential φ\varphi to have the pressure gap. From direct integration using the estimates (3.3) and (3.4) we immediately get

Proposition 3.3 (Key inequality).

For some R>0R>0, suppose that the shadowing vector w=Πt,R(𝗏)w=\Pi_{t,R}(\mathsf{v}) satisfies (3.3) and (3.4) and φ\varphi satisfies the assumption of Theorem A. Then there exists Q=Q(R,C,ε1,ε2,δ)>0Q=Q(R,C,\varepsilon_{1},\varepsilon_{2},\delta)>0 such that

0tφ(fτu)𝑑τ0s1φ(fτ𝗏)𝑑τQ\int_{0}^{t}\varphi(f_{\tau}u)d\tau\geq\int_{0}^{s_{1}}\varphi(f_{\tau}\mathsf{v})d\tau-Q

for any uBt(w,δ)u\in B_{t}(w,\delta) where w=Πt,R(𝗏)w=\Pi_{t,R}(\mathsf{v}) for some 𝗏Sing\mathsf{v}\in\mathrm{Sing}.

Proof.

We firstly prove the inequality for u=wu=w. For type 1 manifold is easier as φc\varphi\equiv c on Sing. By (1.2), (3.3) and (3.4), we have

0tφ(fτu)𝑑τctQcs1Q\int_{0}^{t}\varphi(f_{\tau}u)d\tau\geq ct-Q\geq cs_{1}-Q

as |ts1|2R.|t-s_{1}|\leq 2R.
For Type 2 manifolds, use (1.2), (3.3) and (3.4) again we have

0tφ(fτw)𝑑τ0tφ(fτds(w))𝑑τQ=0s1φ(fτ𝗏)𝑑τQ.\int_{0}^{t}\varphi(f_{\tau}w)d\tau\geq\int_{0}^{t}\varphi(f_{\tau}ds(w))d\tau-Q=\int_{0}^{s_{1}}\varphi(f_{\tau}\mathsf{v})d\tau-Q.

For general uBt(w,δ)u\in B_{t}(w,\delta), we begin by intersecting the geodesic γu(τ)\gamma_{u}(\tau) with the hyperplane x=Rx=R so that there are exactly two intersection points, each near π(u)\pi(u) and π(ftu)\pi(f_{t}u); see Figure 3.2. We denote by (u0,t0)(u_{0},t_{0}) the orbit segment connecting such intersection points. Since two orbit segments (u,t)(u,t) and (u0,t0)(u_{0},t_{0}) differ only at either ends by length at most 2δ2\delta, there exists a constant CδC_{\delta} depending only on δ\delta and φ\|\varphi\| such that

|0tφ(fτu)𝑑τ0t0φ(fτu0)𝑑τ|<Cδ.\Big{|}\int_{0}^{t}\varphi(f_{\tau}u)d\tau-\int_{0}^{t_{0}}\varphi(f_{\tau}u_{0})d\tau\Big{|}<C_{\delta}.
Refer to caption
Figure 3.2. Figure for Proposition 3.3

Notice from its construction that u0u_{0} is equal to Πt0,R(v0)\Pi_{t_{0},R}(v_{0}) for some v0Singv_{0}\in\mathrm{Sing} near vv, and hence the integral 0t0φ(fτu0)𝑑τ{\displaystyle\int_{0}^{t_{0}}\varphi(f_{\tau}u_{0})d\tau} admits a uniform lower bound independent of t0t_{0} of the above proposition. Therefore, the same is true for 0tφ(fτu)𝑑τ{\displaystyle\int_{0}^{t}\varphi(f_{\tau}u)d\tau} for uBt(w,δ)u\in B_{t}(w,\delta). ∎

4. Estimates for type 1 manifolds

Recall that X=/xX=\partial/\partial x is the vertical vector field. For any vTM~v\in T\widetilde{M} that is not collinear with XX we define the radial curvature of vv by

(4.1) K(v):=Kσ,K_{\perp}(v):=K_{\sigma},

that is, the sectional curvature of the plane σ:=span{v,X}.\sigma:=\text{span}\{v,X\}.

In this section, we will consider the first of the two settings in which KK_{\perp} vanishes uniformly to the order m1m-1. Namely, if (s,x)(s,x) are the Fermi coordinates along T~0\widetilde{T}_{0}, there exists C1,C2,ε>0C_{1},C_{2},\varepsilon>0 such that

(4.2) C1|x(v)|mK(v)C2|x(v)|m-C_{1}|x(v)|^{m}\leq K_{\perp}(v)\leq-C_{2}|x(v)|^{m}

for any vXv\perp X with |x(v)|<ε|x(v)|<\varepsilon.

As outlined in Subsection 3.2, the main goal of this section is to prove that under the above assumption on KK_{\perp}, the shadowing vector w=Πt,R(𝗏)w=\Pi_{t,R}(\mathsf{v}) for any 𝗏Sing\mathsf{v}\in\mathrm{Sing} satisfies the estimates on xwx_{w} and ϕw\phi_{w} claimed in (3.3).

Indeed, in this subsection, we will derive estimates on xv(t)x_{v}(t) for generic vectors vT1Mv\in T^{1}M near T~0\tilde{T}_{0}; namely, bouncing, asymptotic, and crossing vectors (see Definition 4.1). In Section 6, we will discuss behaviors of geometric potentials φu\varphi^{u} with respect to bouncing, asymptotic, and crossing vectors. Notice that by Definition 4.1, shadowing vectors w=Πt,R(𝗏)w=\Pi_{t,R}(\mathsf{v}) are bouncing vectors.

Let NR(T~0):={(s,x):sT~0|x|<R}N_{R}(\widetilde{T}_{0}):=\left\{(s,x):s\in\widetilde{T}_{0}\ |x|<R\right\} be a neighborhood of T~0\widetilde{T}_{0}, vT1Mv\in T^{1}M and γv(t)=(sv(t),xv(t))\gamma_{v}(t)=(s_{v}(t),x_{v}(t)) in the Fermi coordinates.

Definition 4.1.

Suppose vT1Mv\in T^{1}M such that γv(0)NR(T~0)\gamma_{v}(0)\in N_{R}(\tilde{T}_{0}). Let T1=inf{t:γv(t)NR(T~0)}-T_{1}=\inf\{t:\ \gamma_{v}(t)\in N_{R}(\widetilde{T}_{0})\} and T2=sup{t:γv(t)NR(T~0)}T_{2}=\sup\{t:\ \gamma_{v}(t)\in N_{R}(\widetilde{T}_{0})\}. We say that vv (relative to NR(T~0)N_{R}(\widetilde{T}_{0})) is

  1. (1)

    bouncing if T1,T2<T_{1},T_{2}<\infty and xv(t)>0x_{v}(t)>0 for t(T1,T2)t\in(-T_{1},T_{2}),

  2. (2)

    asymptotic if T1=T_{1}=\infty or T2=T_{2}=\infty,

  3. (3)

    crossing if T1,T2<T_{1},T_{2}<\infty and xv(t0)=0x_{v}(t_{0})=0 for t0(T1,T2)t_{0}\in(-T_{1},T_{2}).

Please see Fig 4.1 for examples of these vectors. The definition and study of these vectors are inspired by [LMM]. Notice that according to Lemma 5.5 the above definition is well-defined when RR is sufficiently small. Moreover, by definition, all shadowing vectors are bouncing vectors and asymptotic vectors are limiting cases of the bouncing vectors. More precisely, one can regard asymptotic vectors as bouncing vectors vv with the minimal of xvx_{v} (when ϕ(v)<0\phi(v)<0) occurring at t~=\tilde{t}=\infty where, recalling from (3.2), t~\tilde{t} is the the time (unique in this case) in which xw(t)x_{w}(t) attains its minimum.

Refer to caption
Figure 4.1. Bouncing, asymptotic and crossing vectors

For type 1 manifolds, the condition (3.3) is established in the proposition below (as for bouncing vectors):

Proposition 4.2.

For any R>0R>0 sufficiently small (see Lemma 4.5 for the domain RR can take), there exists Q=Q(R)>1Q=Q(R)>1 independent of tt such that for any vT1Mv\in T^{1}M with xv(0)=Rx_{v}(0)=R and ϕv(0)<0\phi_{v}(0)<0,

  1. (1)

    if vv is a bouncing vector relative to NR(T~0)N_{R}(\widetilde{T}_{0}), then for τ[0,t~]\tau\in[0,\tilde{t}\,] we have

    Q1(τ+1)2mxv(τ)Q(τ+1)2m;Q^{-1}(\tau+1)^{-\frac{2}{m}}\leq x_{v}(\tau)\leq Q(\tau+1)^{-\frac{2}{m}};
    Q1[(τ+1)m+2m(t~+1)m+2m]|ϕv(τ)|Q[(τ+1)m+2m(t~+1)m+2m];Q^{-1}[(\tau+1)^{-\frac{m+2}{m}}-(\tilde{t}+1)^{-\frac{m+2}{m}}]\leq|\phi_{v}(\tau)|\leq Q[(\tau+1)^{-\frac{m+2}{m}}-(\tilde{t}+1)^{-\frac{m+2}{m}}];
  2. (2)

    if vv is an asymptotic vector relative to NR(T~0)N_{R}(\widetilde{T}_{0}), then for τ[0,)\tau\in[0,\infty) we have

    Q1(τ+1)2mxv(τ)Q(τ+1)2m;Q^{-1}(\tau+1)^{-\frac{2}{m}}\leq x_{v}(\tau)\leq Q(\tau+1)^{-\frac{2}{m}};
    Q1[(τ+1)m+2m]|ϕv(τ)|Q[(τ+1)m+2m];Q^{-1}[(\tau+1)^{-\frac{m+2}{m}}]\leq|\phi_{v}(\tau)|\leq Q[(\tau+1)^{-\frac{m+2}{m}}];
  3. (3)

    if vv is a crossing vector relative to NR(T~0)N_{R}(\widetilde{T}_{0}), then for τ[0,t0]\tau\in[0,t_{0}], we have

    Q1[(τ+1)2m(t0+1)2m]xv(τ)Q[(τ+1)2m(t0+1)2m];Q^{-1}[(\tau+1)^{-\frac{2}{m}}-(t_{0}+1)^{-\frac{2}{m}}]\leq x_{v}(\tau)\leq Q[(\tau+1)^{-\frac{2}{m}}-(t_{0}+1)^{-\frac{2}{m}}];
    Q1(τ+1)m+2m|ϕv(τ)|Q(τ+1)m+2m.Q^{-1}(\tau+1)^{-\frac{m+2}{m}}\leq|\phi_{v}(\tau)|\leq Q(\tau+1)^{-\frac{m+2}{m}}.
Remark 4.3.

Proposition 4.2 is more general than (3.3). The lower bound of |ϕv(τ)||\phi_{v}(\tau)| is not necessary for deriving Proposition 3.3.

We begin by collecting relevant lemmas to prove this proposition, the first of which concerns the general property of Riccati solutions.

Lemma 4.4.

There exists R0=R0(C,m)R_{0}=R_{0}(C,m) such that for any R[0,R0]R\in[0,R_{0}], and any solution U:[0,R]Mn1()U:[0,R]\to M_{n-1}(\mathbb{R}) of the following Riccati equation

(4.3) U+U2CxmIn1=0,U(0)=0,U^{\prime}+U^{2}-Cx^{m}I_{n-1}=0,\quad U(0)=0,

we have

Cxm+12(m+1)In1U(x)Cxm+1m+1In1.\frac{Cx^{m+1}}{2(m+1)}I_{n-1}\leq U(x)\leq\frac{Cx^{m+1}}{m+1}I_{n-1}.
Proof.

Let λ:[0,R]\lambda:[0,R]\to\mathbb{R} be the solution of

λ+λ2Cxm=0,λ(0)=0.\lambda^{\prime}+\lambda^{2}-Cx^{m}=0,\quad\lambda(0)=0.

Then λIn1\lambda I_{n-1} satisfies (4.3). Since the solution of any first-order ODE is unique; we have U=λIn1U=\lambda I_{n-1}.

Now we estimate λ\lambda. Since λ=Cxmλ2Cxm\lambda^{\prime}=Cx^{m}-\lambda^{2}\leq Cx^{m}, we have

λ(x)=0xλ(s)𝑑s0xCsm𝑑s=Cm+1xm+1.\lambda(x)=\int_{0}^{x}\lambda^{\prime}(s)ds\leq\int_{0}^{x}Cs^{m}ds=\frac{C}{m+1}x^{m+1}.

establishing the required upper bound on U(x)U(x).

For the lower bound, let R0:=((m+1)22C)1m+2{\displaystyle R_{0}:=\Big{(}\frac{(m+1)^{2}}{2C}\Big{)}^{\frac{1}{m+2}}}. Then for any x[0,R0]x\in[0,R_{0}] the upper bound for λ(x)\lambda(x) gives

λ=Cxmλ2CxmC2(m+1)2x2m+2C2xm\lambda^{\prime}=Cx^{m}-\lambda^{2}\geq Cx^{m}-\frac{C^{2}}{(m+1)^{2}}x^{2m+2}\geq\frac{C}{2}x^{m}

which then gives λ(x)=0xλ(s)𝑑sC2(m+1)xm+1{\displaystyle\lambda(x)=\int_{0}^{x}\lambda^{\prime}(s)ds\geq\frac{C}{2(m+1)}x^{m+1}} as required. ∎

Recalling the notations xv(τ):=x(γv(τ))x_{v}(\tau):=x(\gamma_{v}(\tau)) and ϕv(τ):=ϕ(γv(τ))\phi_{v}(\tau):=\phi(\gamma_{v}^{\prime}(\tau)) from (2.2), the following lemma uses the curvature bound (4.2) to compare xv′′x^{\prime\prime}_{v} with xvm+1x_{v}^{m+1} for any shadowing vector vv.

Lemma 4.5.

There exists R1=R1(C1,C2,m)>0R_{1}=R_{1}(C_{1},C_{2},m)>0 such that for any R[0,R1]R\in[0,R_{1}] and any vv with xv(0)<Rx_{v}(0)<R, we have

C24m+4xvm+1xv′′C1m+1xvm+1,\frac{C_{2}}{4m+4}x_{v}^{m+1}\leq x_{v}^{\prime\prime}\leq\frac{C_{1}}{m+1}x_{v}^{m+1},

as long as |xv|<R|x_{v}|<R. In particular, xvx_{v} is strictly convex and positive.

Proof.

Since C1x(v)mK(v)C2x(v)m,-C_{1}x(v)^{m}\leq K_{\perp}(v)\leq-C_{2}x(v)^{m}, we have

C1xmIn1𝒦(X)C2xmIn1-C_{1}x^{m}I_{n-1}\leq\mathcal{K}(X)\leq-C_{2}x^{m}I_{n-1}

where 𝒦:vv\mathcal{K}:v^{\perp}\to v^{\perp} is the symmetric linear map such that 𝒦(v)X,Y=R(X,v)v,Y\langle\mathcal{K}(v)X,Y\rangle=\langle R(X,v)v,Y\rangle for X,YvX,Y\in v^{\perp}. Using R0R_{0} from the previous lemma, we claim that we can take

R1:=min{R0(C1,m),R0(C2,m)}.R_{1}:=\min\{R_{0}(C_{1},m),R_{0}(C_{2},m)\}.

For i=1,2i=1,2 and R(0,R1)R\in(0,R_{1}), let Ui:[0,R]Mn1()U_{i}:[0,R]\to M_{n-1}(\mathbb{R}) be the solution of

Ui+Ui2CixmIn1=0,Ui(0)=0.U_{i}^{\prime}+U_{i}^{2}-C_{i}x^{m}I_{n-1}=0,\quad U_{i}(0)=0.

By the main theorem in [EH90], the solutions satisfy

U2UU1U_{2}\leq U\leq U_{1}

on [0,R][0,R]. By Lemmas 2.2 and 4.4 and Remark 3.2, we have

xv′′(x)λ1(s,xv)2C24m+4xvm+1x^{\prime\prime}_{v}(x)\geq\frac{\lambda_{1}(s,x_{v})}{2}\geq\frac{C_{2}}{4m+4}x_{v}^{m+1}

and

xv′′(x)λn1(s,xv)C1m+1xvm+1.x^{\prime\prime}_{v}(x)\leq\lambda_{n-1}(s,x_{v})\leq\frac{C_{1}}{m+1}x_{v}^{m+1}.

If xvx_{v} vanishes somewhere in (0,t)(0,t), and assume a~\tilde{a} is the smallest zero of xwx_{w} in (0,t)(0,t). It is clear that xw(a~)0x_{w}^{\prime}(\tilde{a})\leq 0 because if xv(a~)=0x_{v}^{\prime}(\tilde{a})=0, then it would imply xvT~0x_{v}\subseteq\widetilde{T}_{0} which is impossible. Thus xv(a~)<0x_{v}^{\prime}(\tilde{a})<0, and denote by b~\tilde{b} the next zero of xvx_{v}. Then xvx_{v} is a geodesic connecting two distinct points on a totally geodesic submanifold T~0\widetilde{T}_{0}, which would imply xvT~0x_{v}\subseteq\widetilde{T}_{0}, again resulting in a contradiction. Therefore, xv(τ)x_{v}(\tau) is positive for all τ[0,t]\tau\in[0,t]. ∎

We also need the following auxiliary lemma.

Lemma 4.6.

Let f:[a,b]f:[a,b]\to\mathbb{R} be a piecewise smooth, strictly decreasing function with finitely many discontinuities. Assume that f(b)>0f(b)>0 and that there exists 0<Q1<Q2,α>0,β(0,1)0<Q_{1}<Q_{2},\alpha>0,\beta\in(0,1) with αβ>1\alpha\beta>1 such that

Q2(f(τ)αf(b)α)βf(τ)Q1(f(τ)αf(b)α)β-Q_{2}(f(\tau)^{\alpha}-f(b)^{\alpha})^{\beta}\leq f^{\prime}(\tau)\leq-Q_{1}(f(\tau)^{\alpha}-f(b)^{\alpha})^{\beta}

when ff is smooth at τ\tau. Then exists a constant Q0=Q0(α,β)>0Q_{0}=Q_{0}(\alpha,\beta)>0 independent of a,b,fa,b,f such that

(f(a)1αβ+Q2(αβ1)(τa))11αβf(τ)Q0(f(a)1αβ+Q1(αβ1)(τa))11αβ(f(a)^{1-\alpha\beta}+Q_{2}(\alpha\beta-1)(\tau-a))^{\frac{1}{1-\alpha\beta}}\leq f(\tau)\leq Q_{0}(f(a)^{1-\alpha\beta}+Q_{1}(\alpha\beta-1)(\tau-a))^{\frac{1}{1-\alpha\beta}}

for all τ[a,b]\tau\in[a,b].

Proof.

We first consider the lower bound of ff. Firstly we have

(4.4) ddτf(τ)1αβ=(αβ1)(f)f(τ)αβQ2(αβ1),\frac{d}{d\tau}f(\tau)^{1-\alpha\beta}=(\alpha\beta-1)(-f^{\prime})f(\tau)^{-\alpha\beta}\leq Q_{2}(\alpha\beta-1),

whenever ff is smooth. Thus,

f(τ)1αβf(a)1αβ+Q2(αβ1)(τa).f(\tau)^{1-\alpha\beta}\leq f(a)^{1-\alpha\beta}+Q_{2}(\alpha\beta-1)(\tau-a).

Hence,

f(τ)(f(a)1αβ+Q2(αβ1)(τa))11αβ.f(\tau)\geq(f(a)^{1-\alpha\beta}+Q_{2}(\alpha\beta-1)(\tau-a))^{\frac{1}{1-\alpha\beta}}.

Now, we compute the upper bound. Similar to the lower bound, we get

ddτf(τ)1αβ=(αβ1)(f)f(τ)αβQ1(αβ1)(1(f(b)f(τ))α)β,\frac{d}{d\tau}f(\tau)^{1-\alpha\beta}=(\alpha\beta-1)(-f^{\prime})f(\tau)^{-\alpha\beta}\geq Q_{1}(\alpha\beta-1)\left(1-\left(\frac{f(b)}{f(\tau)}\right)^{\alpha}\right)^{\beta},

whenever ff is smooth.

We then define an auxiliary piecewise smooth function

g(τ):=(f(τ)f(b))1αβg(\tau):=\left(\frac{f(\tau)}{f(b)}\right)^{1-\alpha\beta}

which is strictly increasing on [a,b][a,b] with g(a)(0,1)g(a)\in(0,1) and g(b)=1g(b)=1. Moreover,

(4.5) dgdτQ1(αβ1)f(b)αβ1(1gααβ1)β.\frac{dg}{d\tau}\geq Q_{1}(\alpha\beta-1)f(b)^{\alpha\beta-1}\left(1-g^{\frac{\alpha}{\alpha\beta-1}}\right)^{\beta}.

Let g~\tilde{g} be the solution of the ODE

(4.6) dg~dτ=Q1(αβ1)f(b)αβ1(1g~1/β)β,g~(a)=g(a).\frac{d\tilde{g}}{d\tau}=Q_{1}(\alpha\beta-1)f(b)^{\alpha\beta-1}\left(1-\tilde{g}^{1/\beta}\right)^{\beta},\quad\quad\tilde{g}(a)=g(a).

Since ααβ1>1β\frac{\alpha}{\alpha\beta-1}>\frac{1}{\beta}, from (4.5) and (4.6) we know that g~(a)<g(a)\tilde{g}^{\prime}(a)<g^{\prime}(a), thus g~(τ)<g(τ)\tilde{g}(\tau)<g(\tau) when τ\tau is slightly larger than aa. In fact, we have g~(τ)<g(τ)\tilde{g}(\tau)<g(\tau) for all τ(a,b)\tau\in(a,b). This is because if g~(τ)g(τ)\tilde{g}(\tau)\geq g(\tau) for some τ\tau, then we can define τ0(a,b)\tau_{0}\in(a,b) to be the smallest τ\tau with g~(τ)=g(τ)\tilde{g}(\tau)=g(\tau). Since g~(τ)<g(τ)\tilde{g}(\tau)<g(\tau) on (a,τ0)(a,\tau_{0}), the condition g~(τ0)=g(τ0)\tilde{g}(\tau_{0})=g(\tau_{0}) implies g~(τ0)g(τ0)\tilde{g}^{\prime}(\tau_{0})\geq g^{\prime}(\tau_{0}). On the other hand, the condition g~(τ0)=g(τ0)\tilde{g}(\tau_{0})=g(\tau_{0}) considered with (4.5) and (4.6) implies g~(τ0)<g(τ0)\tilde{g}^{\prime}(\tau_{0})<g^{\prime}(\tau_{0}), deriving a contradiction. Thus g~<g\tilde{g}<g on (a,b)(a,b).

By (4.6), we have

(1g~1/β)βdg~dτ=Q1(αβ1)f(b)αβ1.\left(1-\tilde{g}^{1/\beta}\right)^{-\beta}\frac{d\tilde{g}}{d\tau}=Q_{1}(\alpha\beta-1)f(b)^{\alpha\beta-1}.

Thus, for any τ[a,b]\tau\in[a,b],

(4.7) Fβ(g~(τ))Fβ(g~(a))=Q1(αβ1)f(b)αβ1(τa),F_{\beta}(\tilde{g}(\tau))-F_{\beta}(\tilde{g}(a))=Q_{1}(\alpha\beta-1)f(b)^{\alpha\beta-1}(\tau-a),

where

Fβ(x):=0x(1y1/β)β𝑑y.F_{\beta}(x):=\int_{0}^{x}(1-y^{1/\beta})^{-\beta}\,dy.

It is clear that FβF_{\beta} is convex on [0,1][0,1], thus Fβ(x)Fβ(0)=1F^{\prime}_{\beta}(x)\geq F^{\prime}_{\beta}(0)=1 for x[0,1]x\in[0,1]. Hence, for any x[0,1]x\in[0,1],

(4.8) Fβ(x)x.F_{\beta}(x)\geq x.

On the other hand, since FβF_{\beta} is convex and increasing, Fβ(x)/xF_{\beta}(x)/x is also increasing on [0,1][0,1]. Thus, for any x(0,1)x\in(0,1)

(4.9) Fβ(x)xFβ(1)=01tβ1(1t)β𝑑t=(β,1β)<,\frac{F_{\beta}(x)}{x}\leq F_{\beta}(1)=\int_{0}^{1}t^{\beta-1}(1-t)^{-\beta}dt=\mathcal{B}(\beta,1-\beta)<\infty,

where \mathcal{B} is the beta function. By combining (4.7), (4.8), and (4.9), we get

(β,1β)g~(τ)g~(a)Q1(αβ1)f(b)αβ1(τa).\mathcal{B}(\beta,1-\beta)\tilde{g}(\tau)-\tilde{g}(a)\geq Q_{1}(\alpha\beta-1)f(b)^{\alpha\beta-1}(\tau-a).

Hence

(β,1β)f(τ)1αβf(a)1αβQ1(αβ1)(τa).\mathcal{B}(\beta,1-\beta)f(\tau)^{1-\alpha\beta}-f(a)^{1-\alpha\beta}\geq Q_{1}(\alpha\beta-1)(\tau-a).

Setting Q0:=(β,1β)1αβ1Q_{0}:=\mathcal{B}(\beta,1-\beta)^{\frac{1}{\alpha\beta-1}}, we have

f(τ)Q0(f(a)1αβ+Q1(αβ1)(τa))11αβ.f(\tau)\leq Q_{0}(f(a)^{1-\alpha\beta}+Q_{1}(\alpha\beta-1)(\tau-a))^{\frac{1}{1-\alpha\beta}}.

This completes the proof. ∎

We are ready to prove Proposition 4.2.

Proof of Proposition 4.2.

We will use xx to denote xvx_{v} for simplicity. We will also use QQ to denote a generic constant that may need to be updated; this will be made clearer as they show up in the proof.

Case 1: vv is a bouncing vector. We will first prove the lower bound for x(τ)x(\tau) when τ[0,t~]\tau\in[0,\tilde{t}\,]. Noting that x(t~)=0x^{\prime}(\tilde{t})=0, by Lemma 4.5, we have

x(τ)2=τt~2xx′′𝑑s2C1m+1τt~xxm+1ds2C1(m+1)2(x(τ)m+2x(t~)m+2)x^{\prime}(\tau)^{2}=-\int_{\tau}^{\tilde{t}}2x^{\prime}x^{\prime\prime}ds\leq\frac{2C_{1}}{m+1}\int_{\tau}^{\tilde{t}}-x^{\prime}x^{m+1}ds\leq\frac{2C_{1}}{(m+1)^{2}}(x(\tau)^{m+2}-x(\tilde{t})^{m+2})

and

x(τ)2=τt~2xx′′𝑑sC22m+2τt~xxm+1dsC22(m+2)2(x(τ)m+2x(t~)m+2).x^{\prime}(\tau)^{2}=-\int_{\tau}^{\tilde{t}}2x^{\prime}x^{\prime\prime}ds\geq\frac{C_{2}}{2m+2}\int_{\tau}^{\tilde{t}}-x^{\prime}x^{m+1}ds\geq\frac{C_{2}}{2(m+2)^{2}}(x(\tau)^{m+2}-x(\tilde{t})^{m+2}).

By taking α=m+2,β=1/2,f(a)=R\alpha=m+2,\beta=1/2,f(a)=R in Lemma 4.6, we know that there exists QQ independent of v,tv,t such that for any τ[0,t~]\tau\in[0,\tilde{t}],

Q1(τ+1)2mx(τ)Q(τ+1)2m.Q^{-1}(\tau+1)^{-\frac{2}{m}}\leq x(\tau)\leq Q(\tau+1)^{-\frac{2}{m}}.

For ϕ\phi, recall that x=sinϕx^{\prime}=\sin\phi. Thus (sinϕ)=x′′(\sin\phi)^{\prime}=x^{\prime\prime}. By Lemma 4.5 we have

Q1(τ+1)2(m+1)mC24m+4xm+1(sinϕ)C1m+1xm+1Q(τ+1)2(m+1)m.Q^{-1}(\tau+1)^{-\frac{2(m+1)}{m}}\leq\frac{C_{2}}{4m+4}x^{m+1}\leq(\sin\phi)^{\prime}\leq\frac{C_{1}}{m+1}x^{m+1}\leq Q(\tau+1)^{-\frac{2(m+1)}{m}}.

Taking the integral on [τ,t~][\tau,\tilde{t}], we get

Q1[(τ+1)m+2m(t~+1)m+2m]sinϕ(τ)Q[(τ+1)m+2m(t~+1)m+2m].Q^{-1}[(\tau+1)^{-\frac{m+2}{m}}-(\tilde{t}+1)^{-\frac{m+2}{m}}]\leq\sin\phi(\tau)\leq Q[(\tau+1)^{-\frac{m+2}{m}}-(\tilde{t}+1)^{-\frac{m+2}{m}}].

Since sinϕ/ϕ[2/π,1]\sin\phi/\phi\in[2/\pi,1], we have the same bounds for ϕ\phi.

Case 2: vv is an asymptotic vector. It is not hard to see that the above argument is valid when t~=.\tilde{t}=\infty.

Case 3: vv is a crossing vector. Denote by a(t):=sinϕ(t)a(t):=-\sin\phi(t). By Lemma 4.5, we know that whenever |x|R|x|\leq R,

C24m+4xm+1x′′C1m+1xm+1.\frac{C_{2}}{4m+4}x^{m+1}\leq x^{\prime\prime}\leq\frac{C_{1}}{m+1}x^{m+1}.

Thus

C24m+4(xxm+1)xx′′C1m+1(xxm+1).\frac{C_{2}}{4m+4}(-x^{\prime}x^{m+1})\leq-x^{\prime}x^{\prime\prime}\leq\frac{C_{1}}{m+1}(-x^{\prime}x^{m+1}).

Taking integral, we get

(4.10) C2(2m+4)2xm+2a2a(t0)2C1(m+1)2xm+2.\frac{C_{2}}{(2m+4)^{2}}x^{m+2}\leq a^{2}-a(t_{0})^{2}\leq\frac{C_{1}}{(m+1)^{2}}x^{m+2}.

Hence

(m+1)2C1(a2a(t0)2)xm+2(2m+4)2C2(a2a(t0)2)\frac{(m+1)^{2}}{C_{1}}(a^{2}-a(t_{0})^{2})\leq x^{m+2}\leq\frac{(2m+4)^{2}}{C_{2}}(a^{2}-a(t_{0})^{2})

Since x=ax^{\prime}=-a, we have a=x′′a^{\prime}=-x^{\prime\prime}. Thus there exists C>c>0C>c>0 independent of v,tv,t such that

C(a2a(t0)2)m+1m+2ac(a2a(t0)2)m+1m+2.-C(a^{2}-a(t_{0})^{2})^{\frac{m+1}{m+2}}\leq a^{\prime}\leq-c(a^{2}-a(t_{0})^{2})^{\frac{m+1}{m+2}}.

Setting α=2,β=m+1m+2\alpha=2,\beta=\frac{m+1}{m+2} in Lemma 4.6, we have

(4.11) (a(0)mm+2+C(τa))m+2ma(τ)Q0(a(0)mm+2+c(τa))m+2m(a(0)^{-\frac{m}{m+2}}+C(\tau-a))^{-\frac{m+2}{m}}\leq a(\tau)\leq Q_{0}(a(0)^{-\frac{m}{m+2}}+c(\tau-a))^{-\frac{m+2}{m}}

Since xv(0)=Rx_{v}(0)=R and a(0)>0a(0)>0, by compactness, a(0)a(0) has a uniform lower bound depending on RR. Thus, we have

Q1(τ+1)m+2m|ϕ(τ)|Q(τ+1)m+2m.Q^{-1}(\tau+1)^{-\frac{m+2}{m}}\leq|\phi(\tau)|\leq Q(\tau+1)^{-\frac{m+2}{m}}.

Similar to Case 1, the bound of xx comes from x(τ)=τt0a(s)𝑑sx(\tau)=\int_{\tau}^{t_{0}}a(s)ds and (4.11). ∎

Remark 4.7.

Note that we did not use the full strength of Lemma 4.6 in the above proof; that is, x(τ)x(\tau) had no discontinuities. The next section will make similar use of Lemma 4.6 applied to a function with discontinuities.

5. Estimates for type 2 surfaces

In this section, we consider a different setting considered by Gerber and Niţică [GN99] as well as Gerber and Wilkinson [GW99] where M=SM=S is a complete nonpositively curved surface, and T0T_{0} is a closed geodesic of some length γ0\gamma_{0} on which the Gaussian curvature KK vanishes to order m1m-1. Namely, if (s,x)(s,x) are the Fermi coordinates along T~0\widetilde{T}_{0}, there exists C1,C2,ε>0C_{1},C_{2},\varepsilon>0 and an interval L=[0,γ1]L=[0,\gamma_{1}] for some γ1(0,γ0)\gamma_{1}\in(0,\gamma_{0}) such that

(5.1) C1|x|mK(s,x)0-C_{1}|x|^{m}\leq K(s,x)\leq 0

for all |x|<ε|x|<\varepsilon and for all ss\in\mathbb{R} and

(5.2) C1|x|mK(s,x)C2|x|m-C_{1}|x|^{m}\leq K(s,x)\leq-C_{2}|x|^{m}

for all |x|<ε|x|<\varepsilon and sL~s\in\tilde{L}. To simplify the argument, whenever applicable, we will adopt the notation for the Riemannian metric gg specified for a surface introduced in Remark 2.3.

Remark 5.1.

Compared to the assumption in the previous section, the underlying manifold considered in this section is 2-dimensional, and the curvature assumption near T0T_{0} is weakened: the neighborhood of only a small subset LL of T0T_{0} is assumed to satisfy the uniform curvature bound as in (4.2).

On the complement of LL in T0T_{0} and its neighborhood, only the trivial upper bound (i.e., zero) is imposed on the curvature. Despite the weaker assumption on the curvature, the low dimensionality of the manifold enables us to do a finer analysis to prove the similar estimates (3.3) on xwx_{w} and ϕw\phi_{w} for w=Πt,R(𝗏)w=\Pi_{t,R}(\mathsf{v}). Furthermore, unlike in the previous section where RR from the definition (3.1) of the shadowing map Πt,R\Pi_{t,R} had to be carefully chosen, this setting is less sensitive to the choice of RR.

Remark 5.2.

One can continue using techniques developed in the previous section to study bouncing, asymptotic, and crossing vectors. However, the geometric potential estimates were well studied in the surface setting in [GW99]. Without deviating from the main goal and to simplify the argument in this section, we will only focus on shadowing vectors w=Πt,R(𝗏)w=\Pi_{t,R}(\mathsf{v}).

The goal of this section is to show that under this different set of assumptions, the shadowing vector w=Πt,R(𝗏)w=\Pi_{t,R}(\mathsf{v}) satisfies the estimates in xwx_{w} and ϕw\phi_{w} as claimed in (3.3). Recalling that γ0\gamma_{0} is the length of T0T_{0}, we state it as a proposition below, which is the analog of Proposition 4.2.

Proposition 5.3.

There exists Q=Q(R)>1Q=Q(R)>1 independent of tt such that for any shadowing vector w=Πt,R(𝗏)w=\Pi_{t,R}(\mathsf{v}) and τ[0,t~]\tau\in[0,\tilde{t}] we have

Q1(τ+1)2mxw(τ)Q(τ+1)2m,Q^{-1}(\tau+1)^{-\frac{2}{m}}\leq x_{w}(\tau)\leq Q(\tau+1)^{-\frac{2}{m}},

and

|ϕw(τ)|Q[(τ+1)m+2m(t~+1)m+2m],|\phi_{w}(\tau)|\leq Q[(\tau+1)^{-\frac{m+2}{m}}-(\tilde{t}+1)^{-\frac{m+2}{m}}],

and for any τ[0,t~22γ0]\tau\in[0,\tilde{t}-2\sqrt{2}\gamma_{0}],

|ϕw(τ)|Q1[(τ+1)m+2m(t~+1)m+2m].|\phi_{w}(\tau)|\geq Q^{-1}[(\tau+1)^{-\frac{m+2}{m}}-(\tilde{t}+1)^{-\frac{m+2}{m}}].
Remark 5.4.

We do not expect the lower bound of |ϕw(τ)||\phi_{w}(\tau)| to hold for all τ[0,t~]\tau\in[0,\tilde{t}\,] since we have little control of the metric near γ(t~)\gamma(\tilde{t}). For instance, when the metric near γ(t~)\gamma(\tilde{t}) is isometric to the surface of revolution of 1+x2m1+x^{2m}, by Proposition 3.1 we know that ϕw(τ)\phi_{w}(\tau) is of the same scale as (τ+1)m+1m(t~+1)m+1m(\tau+1)^{-\frac{m+1}{m}}-(\tilde{t}+1)^{-\frac{m+1}{m}} when τ\tau is near t~\tilde{t}.

To prove the proposition, we need to exploit the assumptions on K(s,x)K(s,x) and establish a few auxiliary lemmas. Consider any w=Πt,R(𝗏)w=\Pi_{t,R}(\mathsf{v}) for some t,R>0t,R>0 and vSingv\in\mathrm{Sing}. Recall that t~[0,t]\tilde{t}\in[0,t] is the smallest number in which xw(τ)x_{w}(\tau) attains the minimum. We can decompose [0,t~][0,\tilde{t}] into subintervals Ii=[ti,ti]I_{i}=[t_{i},t^{\prime}_{i}] and Ii=(ti,ti+1)I^{\prime}_{i}=(t^{\prime}_{i},t_{i+1}) for i=0,1,,ni=0,1,\cdots,n so that

s(γw(τ))L~ when τIis(\gamma_{w}(\tau))\in\tilde{L}\text{ when }\tau\in I_{i}

and

s(γw(τ))L~ when τIi;s(\gamma_{w}(\tau))\notin\tilde{L}\text{ when }\tau\in I^{\prime}_{i};

see Figure 5.1. Notice that t~tn\tilde{t}\geq t_{n}, thus InI_{n} is not empty (though it may be arbitrarily short), but InI^{\prime}_{n} may be empty.

Refer to caption
Figure 5.1. Division into IiI_{i} and IiI_{i}^{\prime}.

Since the angle ϕw(τ)\phi_{w}(\tau) satisfies |ϕw(τ)|[0,π/4]|\phi_{w}(\tau)|\in[0,\pi/4] for any τ[0,t]\tau\in[0,t] from Lemma 3.1, there exists NN\in\mathbb{N} so that

mini|Ii|1Nmaxi|Ii|\min\limits_{i}|I_{i}|\geq\frac{1}{N}\max\limits_{i}|I^{\prime}_{i}|

where the maximum is taken over all possible ii and the minimum is taken over ii in {1,,n}\{1,\ldots,n\} if t0Int_{0}\in I_{n}^{\prime} (i.e. InI_{n}^{\prime} is nonempty) or else (i.e. t~In\tilde{t}\in I_{n} and InI_{n}^{\prime} is empty) in {1,,n1}\{1,\ldots,n-1\} in order to exclude |In||I_{n}| which could be arbitrarily small.

The following lemma shows xw′′(τ)x_{w}^{\prime\prime}(\tau) admits a similar bound as in Lemma 4.5 when τ\tau belongs to IiI_{i} for some ii.

Lemma 5.5.

There exists C0>0C_{0}>0 independent of ww such that

0xw′′C0xwm+10\leq x_{w}^{\prime\prime}\leq C_{0}x_{w}^{m+1}

for all τ[0,t~]\tau\in[0,\tilde{t}\,]. Moreover, there exists c0(0,C0)c_{0}\in(0,C_{0}) such that whenever τIi\tau\in I_{i} for some ii, we also have the lower bound

xw′′c0xwm+1.x_{w}^{\prime\prime}\geq c_{0}x_{w}^{m+1}.
Proof.

Since Gxx=KG2C1xwmG_{xx}=-KG\leq 2C_{1}x_{w}^{m} with Gx(s,0)=0G_{x}(s,0)=0 for all ss, thus Gx(s,xw)2C1(m+1)1xwm+1G_{x}(s,x_{w})\leq 2C_{1}(m+1)^{-1}x_{w}^{m+1}. Let C0:=2C1(m+1)1C_{0}:=2C_{1}(m+1)^{-1}. By (2.6),

xw′′=GxGcos2ϕwGxGC0xwm+1.x^{\prime\prime}_{w}=\frac{G_{x}}{G}\cos^{2}\phi_{w}\leq\frac{G_{x}}{G}\leq C_{0}x_{w}^{m+1}.

For tIit\in I_{i}, we have Gxx=KGC2xwmG_{xx}=-KG\geq C_{2}x_{w}^{m}, thus Gx(s,xw)C2(m+1)1xwm+1G_{x}(s,x_{w})\geq C_{2}(m+1)^{-1}x_{w}^{m+1}. Let c0:=C2(2m+2)1c_{0}:=C_{2}(2m+2)^{-1}. By Lemma 3.1 (1), we have

xw′′=GxGcos2ϕwGx2Gc0xwm+1.x^{\prime\prime}_{w}=\frac{G_{x}}{G}\cos^{2}\phi_{w}\geq\frac{G_{x}}{\sqrt{2}G}\geq c_{0}x_{w}^{m+1}.

This completes the proof. ∎

The following lemma we let ti′′:=ti+ti2{\displaystyle t^{\prime\prime}_{i}:=\frac{t_{i}+t^{\prime}_{i}}{2}}. For simplicity, in the remaining part of this section, we abbreviate x=xwx=x_{w} and ϕ=ϕw\phi=\phi_{w} where w=Πt,R(𝗏)w=\Pi_{t,R}(\mathsf{v}) for any 𝗏Sing\mathsf{v}\in\mathrm{Sing} and t,R>0t,R>0.

Lemma 5.6.

For any ii with |Ii|γ1|I_{i}|\geq\gamma_{1} (namely, those IiI_{i} not containing 0 or t0t_{0}), we have

  1. (1)

    ti′′tixxm+1ds12N+1ti′′ti+1xxm+1ds.{\displaystyle\int_{t^{\prime\prime}_{i}}^{t^{\prime}_{i}}-x^{\prime}x^{m+1}ds\geq\frac{1}{2N+1}\int_{t^{\prime\prime}_{i}}^{t_{i+1}}-x^{\prime}x^{m+1}ds.}

  2. (2)

    τtixxm+1ds12N+1τti+1xxm+1ds{\displaystyle\int_{\tau}^{t^{\prime}_{i}}-x^{\prime}x^{m+1}ds\geq\frac{1}{2N+1}\int_{\tau}^{t_{i+1}}-x^{\prime}x^{m+1}ds} for any τ[ti,ti′′]\tau\in[t_{i},t^{\prime\prime}_{i}].

Proof.

For (1) consider any s[ti′′,ti]s\in[t^{\prime\prime}_{i},t^{\prime}_{i}] and s[ti,ti+1]s^{\prime}\in[t^{\prime}_{i},t_{i+1}]. Since xx is convex and decreasing, we have x(s)x(s)-x^{\prime}(s)\geq-x^{\prime}(s^{\prime}) and x(s)x(s)x(s)\geq x(s^{\prime}). In particular,

x(s)x(s)m+1x(s)x(s)m+1.-x^{\prime}(s)x(s)^{m+1}\geq-x^{\prime}(s^{\prime})x(s^{\prime})^{m+1}.

Since mini|Ii|1Nmaxi|Ii|{\displaystyle\min_{i}|I_{i}|\geq\frac{1}{N}\max_{i}|I^{\prime}_{i}|}, we can divide IiI^{\prime}_{i} into 2N2N subintervals of the same length. The integral on each subinterval, whose length is at most |Ii|/2|I_{i}|/2, is no more than that on [ti′′,ti][t^{\prime\prime}_{i},t^{\prime}_{i}]. Hence

ti′′ti+1xxm+1ds\displaystyle\int_{t^{\prime\prime}_{i}}^{t_{i+1}}-x^{\prime}x^{m+1}ds =\displaystyle= ti′′tixxm+1ds+titi+1xxm+1ds\displaystyle\int_{t^{\prime\prime}_{i}}^{t^{\prime}_{i}}-x^{\prime}x^{m+1}ds+\int_{t^{\prime}_{i}}^{t_{i+1}}-x^{\prime}x^{m+1}ds
\displaystyle\leq ti′′tixxm+1ds+2Nti′′tixxm+1ds\displaystyle\int_{t^{\prime\prime}_{i}}^{t^{\prime}_{i}}-x^{\prime}x^{m+1}ds+2N\int_{t^{\prime\prime}_{i}}^{t^{\prime}_{i}}-x^{\prime}x^{m+1}ds
=\displaystyle= (1+2N)ti′′tixxm+1ds.\displaystyle(1+2N)\int_{t^{\prime\prime}_{i}}^{t^{\prime}_{i}}-x^{\prime}x^{m+1}ds.

For (2), by (1), we have

τtixxm+1ds\displaystyle\int_{\tau}^{t^{\prime}_{i}}-x^{\prime}x^{m+1}ds =\displaystyle= τti′′xxm+1ds+ti′′tixxm+1ds\displaystyle\int_{\tau}^{t^{\prime\prime}_{i}}-x^{\prime}x^{m+1}ds+\int_{t^{\prime\prime}_{i}}^{t^{\prime}_{i}}-x^{\prime}x^{m+1}ds
\displaystyle\geq τti′′xxm+1ds+12N+1ti′′ti+1xxm+1ds\displaystyle\int_{\tau}^{t^{\prime\prime}_{i}}-x^{\prime}x^{m+1}ds+\frac{1}{2N+1}\int_{t^{\prime\prime}_{i}}^{t_{i+1}}-x^{\prime}x^{m+1}ds
\displaystyle\geq 12N+1τti+1xxm+1ds.\displaystyle\frac{1}{2N+1}\int_{\tau}^{t_{i+1}}-x^{\prime}x^{m+1}ds.

Lemma 5.7.

There exists C3>0C_{3}>0 such that for any ii and τ[ti,ti′′]\tau\in[t_{i},t^{\prime\prime}_{i}],

x(τ)2C3x(τ)m+2x(t~)m+2.x^{\prime}(\tau)\leq-2C_{3}\sqrt{x(\tau)^{m+2}-x(\,\tilde{t}\,)^{m+2}}.
Proof.

If t~In\tilde{t}\in I^{\prime}_{n}, by Lemmas 5.5 and 5.6, we have

x(τ)2\displaystyle x^{\prime}(\tau)^{2} =\displaystyle= τt~2xx′′ds2c0(τtixxm+1ds+k=i+1ntktkxxm+1ds)\displaystyle\int_{\tau}^{\tilde{t}}-2x^{\prime}x^{\prime\prime}ds\geq 2c_{0}\left(\int_{\tau}^{t^{\prime}_{i}}-x^{\prime}x^{m+1}ds+\sum_{k=i+1}^{n}\int_{t_{k}}^{t^{\prime}_{k}}-x^{\prime}x^{m+1}ds\right)
\displaystyle\geq 2c02N+1(τti+1xxm+1ds+k=i+1n1tktk+1xxm+1ds+tnt~xxm+1ds)\displaystyle\frac{2c_{0}}{2N+1}\left(\int_{\tau}^{t_{i+1}}-x^{\prime}x^{m+1}ds+\sum_{k=i+1}^{n-1}\int_{t_{k}}^{t_{k+1}}-x^{\prime}x^{m+1}ds+\int_{t_{n}}^{\tilde{t}}-x^{\prime}x^{m+1}ds\right)
\displaystyle\geq 2c02N+1τt0xxm+1ds2c0(m+2)(2N+1)(x(τ)m+2x(t~)m+2).\displaystyle\frac{2c_{0}}{2N+1}\int_{\tau}^{t_{0}}-x^{\prime}x^{m+1}ds\geq\frac{2c_{0}}{(m+2)(2N+1)}(x(\tau)^{m+2}-x(\,\tilde{t}\,)^{m+2}).

Similarly, if t~In\tilde{t}\in I_{n}, we have

x(τ)2\displaystyle x^{\prime}(\tau)^{2} =\displaystyle= τt~2xx′′ds\displaystyle\int_{\tau}^{\tilde{t}}-2x^{\prime}x^{\prime\prime}ds
\displaystyle\geq 2c0(τtixxm+1ds+k=i+1n1tktkxxm+1ds+tnt~xxm+1ds)\displaystyle 2c_{0}\left(\int_{\tau}^{t^{\prime}_{i}}-x^{\prime}x^{m+1}ds+\sum_{k=i+1}^{n-1}\int_{t_{k}}^{t^{\prime}_{k}}-x^{\prime}x^{m+1}ds+\int_{t_{n}}^{\tilde{t}}-x^{\prime}x^{m+1}ds\right)
\displaystyle\geq 2c02N+1(τti+1xxm+1ds+k=i+1n1tktk+1xxm+1ds+tnt~xxm+1ds)\displaystyle\frac{2c_{0}}{2N+1}\left(\int_{\tau}^{t_{i+1}}-x^{\prime}x^{m+1}ds+\sum_{k=i+1}^{n-1}\int_{t_{k}}^{t_{k+1}}-x^{\prime}x^{m+1}ds+\int_{t_{n}}^{\tilde{t}}-x^{\prime}x^{m+1}ds\right)
\displaystyle\geq 2c02N+1τt~xxm+1ds2c0(m+2)(2N+1)(x(τ)m+2x(t~)m+2).\displaystyle\frac{2c_{0}}{2N+1}\int_{\tau}^{\tilde{t}}-x^{\prime}x^{m+1}ds\geq\frac{2c_{0}}{(m+2)(2N+1)}(x(\tau)^{m+2}-x(\,\tilde{t}\,)^{m+2}).

We finish the proof by taking C3:=21/2c01/2(m+2)1/2(2N+1)1/2C_{3}:=2^{-1/2}c_{0}^{1/2}(m+2)^{-1/2}(2N+1)^{-1/2}. ∎

Proof of Proposition 5.3.

As did in Proposition 4.2, we will use QQ to denote a generic constant. The desired lower bound for x(τ)x(\tau) can be established just as done in Proposition 4.2. This is because the upper bound 0xw′′C0xwm+10\leq x_{w}^{\prime\prime}\leq C_{0}x_{w}^{m+1} from Lemma 5.5 holds for all τ[0,t~]\tau\in[0,\tilde{t}\,], and this is the only ingredient needed for the lower bound on x(τ)x(\tau) in Proposition 4.2. In particular, we have x(τ)(Rm2+C0τ)2mx(\tau)\geq(R^{-\frac{m}{2}}+\sqrt{C_{0}}\tau)^{\frac{2}{m}} for all τ[0,t~]\tau\in[0,\tilde{t}\,].

On the other hand, the desired upper bound for x(τ)x(\tau) is more difficult to obtain. The reason for introducing ti′′t_{i}^{\prime\prime} and establishing Lemma 5.7 was to obtain the upper bound. We will prove the case where 0I00\in I^{\prime}_{0} and t~In\tilde{t}\in I_{n}. Other cases are similar, and we will comment on them at the end of the proof. We let Rt1>0R_{t_{1}}>0 be the lower bound on x(t1)x(t_{1}) from the above paragraph.

First, define a sequence

Sn:=(k=1n1|tk′′tk|)+|t~tn|,S0=0, and Si=k=1i|tk′′tk|.S_{n}:=\Big{(}\sum_{k=1}^{n-1}|t^{\prime\prime}_{k}-t_{k}|\Big{)}+|\tilde{t}-t_{n}|,\leavevmode\nobreak\ S_{0}=0,\text{ and }S_{i}=\sum_{k=1}^{i}|t^{\prime\prime}_{k}-t_{k}|.

Then we have 0=S0<S1<<Sn1<Sn0=S_{0}<S_{1}<\cdots<S_{n-1}<S_{n}. Define a function f:[0,Sn]f:[0,S_{n}]\to\mathbb{R} via

f(τ):=x(τSi+ti+1) when τ(Si,Si+1].f(\tau):=x(\tau-S_{i}+t_{i+1})\text{ when }\tau\in(S_{i},S_{i+1}].

Then ff is a piecewise smooth function with discontinuities at each SiS_{i}. Moreover, Lemma 5.7 shows that ff is strictly decreasing function satisfying the assumption of Lemma 4.6 with Q=C3Q=C_{3}. Thus by Lemma 4.6, there exists Q0>1Q_{0}>1 such that

f(τ)Q0(f(0)m2+C3mτ)2mQ0(Rt1m2+C3mτ)2mf(\tau)\leq Q_{0}(f(0)^{-\frac{m}{2}}+C_{3}m\tau)^{-\frac{2}{m}}\leq Q_{0}(R_{t_{1}}^{-\frac{m}{2}}+C_{3}m\tau)^{-\frac{2}{m}}

for all τ[0,Sn]\tau\in[0,S_{n}]. In particular, this inequality provides an upper bound for x(τ)x(\tau) for τ[ti,ti′′]\tau\in[t_{i},t_{i}^{\prime\prime}] for 1in1\leq i\leq n; here tn′′t_{n}^{\prime\prime} should be replaced by t~\tilde{t} when i=ni=n.

For τ[ti′′,ti+1]\tau\in[t_{i}^{\prime\prime},t_{i+1}], we have Siτ12N+2{\displaystyle\frac{S_{i}}{\tau}\geq\frac{1}{2N+2}} from the choice of NN. Thus for τ[ti′′,ti+1]\tau\in[t_{i}^{\prime\prime},t_{i+1}],

x(τ)\displaystyle x(\tau) \displaystyle\leq x(ti′′)=f(Si)Q0(f(0)m2+C3mSi)2m\displaystyle x(t_{i}^{\prime\prime})=f(S_{i})\leq Q_{0}(f(0)^{-\frac{m}{2}}+C_{3}mS_{i})^{-\frac{2}{m}}
\displaystyle\leq Q0(Rt1m/2+C3m2N+2τ)2m.\displaystyle Q_{0}\left(R_{t_{1}}^{-m/2}+\frac{C_{3}m}{2N+2}\tau\right)^{-\frac{2}{m}}.

For the last remaining subset of the domain when τ[0,t1]\tau\in[0,t_{1}], which is due to the assumption that 0I00\in I^{\prime}_{0}, we have t12|γ0|t_{1}\leq\sqrt{2}|\gamma_{0}|. Therefore,

x(τ)x(0)=RR(2|γ0|+1)2m(τ+1)2m.x(\tau)\leq x(0)=R\leq R(\sqrt{2}|\gamma_{0}|+1)^{\frac{2}{m}}(\tau+1)^{-\frac{2}{m}}.

In sum, we can find Q>0Q>0 such that

x(τ)Q(τ+1)2mx(\tau)\leq Q(\tau+1)^{-\frac{2}{m}}

for all τ[0,t~]\tau\in[0,\tilde{t}\,].

For ϕ\phi, using (2.4) and Lemmas 3.1 (1) and 5.5, there exists Q>0Q>0 such that

ϕ2ϕcosϕ=2x′′2C0xm+1Q(τ+1)2(m+1)m.\phi^{\prime}\leq\sqrt{2}\phi^{\prime}\cos\phi=\sqrt{2}x^{\prime\prime}\leq\sqrt{2}C_{0}x^{m+1}\leq Q(\tau+1)^{-\frac{2(m+1)}{m}}.

Thus we get the required upper bound for ϕ\phi:

|ϕ(τ)|=|τt~ϕ𝑑s|τt~|ϕ|𝑑sQτt~(s+1)2(m+1)m𝑑sQ[(τ+1)m+2m(t~+1)m+2m].|\phi(\tau)|=\left|\int_{\tau}^{\tilde{t}}\phi^{\prime}ds\right|\leq\int_{\tau}^{\tilde{t}}|\phi^{\prime}|ds\leq Q\int_{\tau}^{\tilde{t}}(s+1)^{-\frac{2(m+1)}{m}}ds\leq Q[(\tau+1)^{-\frac{m+2}{m}}-(\tilde{t}+1)^{-\frac{m+2}{m}}].

This completes the proof when 0I00\in I^{\prime}_{0} and t~In\tilde{t}\in I_{n}.

Other remaining cases can be dealt with similarly. When 0I00\in I_{0} and t~In\tilde{t}\in I_{n}, then exactly the same proof works; in fact, there is no need to separately consider τ[0,t1]\tau\in[0,t_{1}] like we did above. In the case where t~In\tilde{t}\in I_{n}^{\prime}, we can use proceed just as we did above by bounding x(τ)x(\tau) above by x(tn′′)x(t_{n}^{\prime\prime}) for τ[tn′′,t~]\tau\in[t_{n}^{\prime\prime},\tilde{t}\,].

Now, we consider the lower bound of |ϕ||\phi|. For any τ[0,t~22γ0]\tau\in[0,\tilde{t}-2\sqrt{2}\gamma_{0}], the interval [τ,t~][\tau,\tilde{t}\,] contains at least one [ti,ti][t_{i},t^{\prime}_{i}]. Let ll (resp. LL) be the minimal (resp. maximal) ii with [ti,ti][τ,t~][t_{i},t^{\prime}_{i}]\subset[\tau,\tilde{t}\,]. We firstly compare the integrals of x(s)m+1x(s)^{m+1} on [τ,tl][\tau,t_{l}] and [tl,tl+1][t_{l},t_{l+1}]. Since |ϕ|<π/4|\phi|<\pi/4,

(5.3) tl+1tlγ0tlτ2.t_{l+1}-t_{l}\geq\gamma_{0}\geq\frac{t_{l}-\tau}{\sqrt{2}}.

Moreover, we have

tl+1+1τ+1=1+tl+1ττ+11+22γ0.\frac{t_{l+1}+1}{\tau+1}=1+\frac{t_{l+1}-\tau}{\tau+1}\leq 1+2\sqrt{2}\gamma_{0}.

Hence

x(tl+1)Q(tl+1+1)2/mQ(τ+1)2/mQx(τ).x(t_{l+1})\geq Q(t_{l+1}+1)^{-2/m}\geq Q(\tau+1)^{-2/m}\geq Qx(\tau).

Together with (5.3), since xx is non-increasing, we get

(5.4) tltl+1x(s)m+1𝑑s(tl+1tl)x(tl+1)m+1Qtlτ2x(τ)m+1Qτtlx(s)m+1𝑑s.\int_{t_{l}}^{t_{l+1}}x(s)^{m+1}ds\geq(t_{l+1}-t_{l})x(t_{l+1})^{m+1}\geq Q\frac{t_{l}-\tau}{\sqrt{2}}x(\tau)^{m+1}\geq Q\int_{\tau}^{t_{l}}x(s)^{m+1}ds.

Notice that ϕ=x′′/cosϕx′′\phi^{\prime}=x^{\prime\prime}/\cos\phi\geq x^{\prime\prime}, and |Ii|/|Ij|[1/2,2]|I_{i}|/|I_{j}|\in[1/\sqrt{2},\sqrt{2}] for any i,ji,j. Thus, by (5.4),

|ϕ(τ)|\displaystyle|\phi(\tau)| =\displaystyle= τt~ϕ(s)𝑑sτt~x′′(s)𝑑sc0i=lLtitix(s)m+1𝑑s\displaystyle\int_{\tau}^{\tilde{t}}\phi^{\prime}(s)ds\geq\int_{\tau}^{\tilde{t}}x^{\prime\prime}(s)ds\geq c_{0}\sum_{i=l}^{L}\int_{t_{i}}^{t^{\prime}_{i}}x(s)^{m+1}\,ds
\displaystyle\geq c0N+1i=lL1titi+1x(s)m+1𝑑s+c0N+1+2tLt~x(s)m+1𝑑s\displaystyle\frac{c_{0}}{N+1}\sum_{i=l}^{L-1}\int_{t_{i}}^{t_{i+1}}x(s)^{m+1}ds+\frac{c_{0}}{N+1+\sqrt{2}}\int_{t_{L}}^{\tilde{t}}x(s)^{m+1}ds
\displaystyle\geq Qτt~x(s)m+1𝑑sQτt~(s+1)2(m+1)/m𝑑s\displaystyle Q\int_{\tau}^{\tilde{t}}x(s)^{m+1}ds\geq Q\int_{\tau}^{\tilde{t}}(s+1)^{-2(m+1)/m}ds
\displaystyle\geq Q[(τ+1)m+2m(t~+1)m+2m].\displaystyle Q[(\tau+1)^{-\frac{m+2}{m}}-(\tilde{t}+1)^{-\frac{m+2}{m}}].

6. Geometric potentials

This section aims to prove Theorem C. Let MM be a closed rank 1 nonpositively curved manifold, and =(ft)t{\mathcal{F}=}(f_{t})_{t\in\mathbb{R}} denote the geodesic flow on T1M.T^{1}M. Recall that the geodesic potential is defined via

φu(v):=limt01tlogdet(dft|Eu(v))=ddt|t=0logdet(dft|Eu(v)).\varphi^{u}(v):=-\lim_{t\to 0}\frac{1}{t}\log\det(df_{t}|_{E^{u}(v)})=-\frac{d}{dt}\Big{|}_{t=0}\log\det(df_{t}|_{E^{u}(v)}).

As indicated in [BCFT18, Section 7.2], it is convenient to consider the following auxiliary function whose time evolution is governed by a Riccati equation:

ψu(v):=limt01tlogdet(Jv,tu)=ddt|t=0logdet(Jv,tu),\psi^{u}(v):=-\lim_{t\to 0}\frac{1}{t}\log\det(J_{v,t}^{u})=-\frac{d}{dt}\Big{|}_{t=0}\log\det(J_{v,t}^{u}),

where Jv,tu:wvJ(t)(ftv)J_{v,t}^{u}:w\in v^{\perp}\mapsto J(t)\in(f_{t}v)^{\perp} and J(t)J(t) is a unstable Jacobi field along γv\gamma_{v} such that J(0)=wJ(0)=w. We also have ψu(v)=tr Uvu(0)\psi^{u}(v)=-\text{tr }U_{v}^{u}(0) where Uvu(t)U_{v}^{u}(t) is the shape operator of the unstable horoshpere Hu(ftv)H^{u}(f_{t}v).

Let π:T1MM\pi:T^{1}M\to M be the canonical projection. Its derivative dπv:TvT1MTπvMd\pi_{v}:T_{v}T^{1}M\to T_{\pi v}M sends Eu(v)E^{u}(v) onto v.v^{\perp}. We have dft=dπftv1Jv,tudπvdf_{t}=d\pi_{f_{t}v}^{-1}\circ J_{v,t}^{u}\circ d\pi_{v}, and thus

det(dft|Eu(v))=det(dπftv)1det(Jv,tu)det(dπv).\det(df_{t}|_{E^{u}(v)})=\det(d\pi_{f_{t}v})^{-1}\det(J_{v,t}^{u})\det(d\pi_{v}).

Thus

(6.1) φu(v)ψu(v)=ddt|t=0logdet(dπftv).\varphi^{u}(v)-\psi^{u}(v)=\frac{d}{dt}\Big{|}_{t=0}\log\det(d\pi_{f_{t}v}).

For any tt, since Uvu(t)U_{v}^{u}(t) is symmetric, we can take an orthonormal basis {ei(t)}i=1n1\{e_{i}(t)\}_{i=1}^{n-1} of vv^{\perp} so that Uvu(t)ei(t)=λi(t)ei(t)U_{v}^{u}(t)e_{i}(t)=\lambda_{i}(t)e_{i}(t) with λi(t)0\lambda_{i}(t)\geq 0. Since dft(ξ)S=Jξ(t)2+Jξ(t)2||df_{t}(\xi)||_{S}=\|J_{\xi}(t)\|^{2}+\|J^{\prime}_{\xi}(t)\|^{2} for ξEvu\xi\in E_{v}^{u}, for any tt, we have an orthonormal basis {ξi(t)}i=1n1\{\xi_{i}(t)\}_{i=1}^{n-1} of EftvuE_{f_{t}v}^{u}, where ξi(t)\xi_{i}(t) is determined by

Jξi(t)=ei(t)1+λi(t)2,Jξi(t)=λi(t)ei(t)1+λi(t)2.J_{\xi_{i}}(t)=\frac{e_{i}(t)}{\sqrt{1+\lambda_{i}(t)^{2}}},\quad J^{\prime}_{\xi_{i}}(t)=\frac{\lambda_{i}(t)e_{i}(t)}{\sqrt{1+\lambda_{i}(t)^{2}}}.

Thus dπftv(ξi(t))=ei(t)/1+λi(t)2d\pi_{f_{t}v}(\xi_{i}(t))=e_{i}(t)/\sqrt{1+\lambda_{i}(t)^{2}} and the matrix of dπftvd\pi_{f_{t}v} with respect to these two orthonormal basis is diag((1+λi(t)2)1/2,,(1+λi(t)2)1/2)\text{diag}((1+\lambda_{i}(t)^{2})^{-1/2},\cdots,(1+\lambda_{i}(t)^{2})^{-1/2}). Hence

(6.2) logdet(dπftv)=logi=1n1(1+λi(t)2)1/2=12logdet(In1+Uvu(t)2).\log\det(d\pi_{f_{t}v})=\log\prod_{i=1}^{n-1}(1+\lambda_{i}(t)^{2})^{-1/2}=-\frac{1}{2}\log\det(I_{n-1}+U_{v}^{u}(t)^{2}).

Now we use the following Jacobi formula for A:Mn1A:\mathbb{R}\to M_{n-1}:

ddtdetA(t)=detA(t) tr(A(t)1dA(t)dt).\frac{d}{dt}\det A(t)=\det A(t)\text{ tr}\left(A(t)^{-1}\frac{dA(t)}{dt}\right).

For simplicity, denote by U=UvuU=U_{v}^{u} and I=In1I=I_{n-1}. By (6.1) and (6.2), we have

φu(v)ψu(v)\displaystyle\varphi^{u}(v)-\psi^{u}(v) =\displaystyle= 12 tr[(I+U(0)2)1(U(0)U(0)+U(0)U(0))]\displaystyle-\frac{1}{2}\text{ tr}\left[(I+U(0)^{2})^{-1}(U^{\prime}(0)U(0)+U(0)U^{\prime}(0))\right]
=\displaystyle= 12 tr[(U(0)(I+U(0)2)1+(I+U(0)2)1U(0))U(0)]\displaystyle-\frac{1}{2}\text{ tr}\left[(U(0)(I+U(0)^{2})^{-1}+(I+U(0)^{2})^{-1}U(0))U^{\prime}(0)\right]
=\displaystyle=  tr[U(0)(I+U(0)2)1U(0)]\displaystyle-\text{ tr}\left[U(0)(I+U(0)^{2})^{-1}U^{\prime}(0)\right]

Since U+U2+𝒦=0U^{\prime}+U^{2}+\mathcal{K}=0, UU is positive semidefinite, and |tr(AB)||tr(A)||tr(B)||\text{tr}(AB)|\leq|\text{tr}(A)||\text{tr}(B)| if A,BA,B are positive semidefinite, we have

|φu(v)ψu(v)|\displaystyle|\varphi^{u}(v)-\psi^{u}(v)| \displaystyle\leq tr[U3(0)(I+U(0)2)1]+ tr[U(0)(I+U(0)2)1(𝒦)]\displaystyle\text{ tr}\left[U^{3}(0)(I+U(0)^{2})^{-1}\right]+\text{ tr}\left[U(0)(I+U(0)^{2})^{-1}(-\mathcal{K})\right]
\displaystyle\leq (tr U(0))3+ tr[U(0)(I+U(0)2)1] tr(𝒦)]\displaystyle(\text{tr }U(0))^{3}+\text{ tr}\left[U(0)(I+U(0)^{2})^{-1}]\text{ tr}(-\mathcal{K})\right]
=\displaystyle= ψu(v)3ψu(v)(Ric(v))ψu(v)(ψu(v)2Ric(v)),\displaystyle-\psi^{u}(v)^{3}-\psi^{u}(v)(-\text{Ric}(v))\leq-\psi^{u}(v)(\psi^{u}(v)^{2}-\text{Ric}(v)),

When vv is sufficiently close to Sing, ψu(v)-\psi^{u}(v) and Ric(v)-\text{Ric}(v) are small nonnegative numbers, thus we have φu(v)ψu(v)\varphi^{u}(v)\approx\psi^{u}(v) near Sing\mathrm{Sing}. We summarize the above discussion below:

Proposition 6.1.

Suppose MM is a closed rank 1 nonpositively curved manifold. Then we have

|φu(v)ψu(v)|ψu(v)(ψu(v)2Ric(v)).|\varphi^{u}(v)-\psi^{u}(v)|\leq-\psi^{u}(v)(\psi^{u}(v)^{2}-\text{Ric}(v)).

In particular, we have φu(v)ψu(v)\varphi^{u}(v)\approx\psi^{u}(v) near Sing\mathrm{Sing}.

6.1. The proof of Theorem C

The strategy of the proof is to study the auxiliary function ψu\psi^{u} through the associated Riccati equation. We establish a version of Theorem C for ψu\psi^{u}. Then Theorem C follows Proposition 6.1.

We remark that the additional Ricci curvature constraint is essential in our argument. In the higher dimension scenario, only having radial curvature controlled is insufficient. Nevertheless, for some special Riemannian metrics, namely, warped products, the radial curvature K(v)K_{\perp}(v) and Ricci curvature Ric(v){\rm Ric}(v) are comparable. Since this observation is not in the mainstream of the current paper, we leave the proof in Appendix A.

Let MM be a type 1 manifold with order m1m-1 Ricci curvature bounds, that is, there exists k0,K0>0k_{0},K_{0}>0 such that

(6.3) K0|xv|mRic(v)k0|xv|m-K_{0}|x_{v}|^{m}\leq\text{Ric}(v)\leq-k_{0}|x_{v}|^{m}

for all vv with |xv|ε|x_{v}|\leq\varepsilon.

Proposition 6.2.

Assume MM satisfies the assumption of Theorem C, then

ψu(v)|xv|m/2+|ϕv|m/(m+2),-\psi^{u}(v)\approx|x_{v}|^{m/2}+|\phi_{v}|^{m/(m+2)},

for any vv near Sing.

In particular, we have the same scale estimation for φu-\varphi^{u}, and Theorem C follows. To prove Proposition 6.2, we need the the following lemma.

Lemma 6.3.

Assume there exist K1>k1>0K_{1}>k_{1}>0 so that

  1. (1)

    K12T2Ric(γv(t))k12T2-K_{1}^{2}T^{-2}\leq\text{Ric}(\gamma^{\prime}_{v}(t))\leq-k_{1}^{2}T^{-2} for all t[T,0]t\in[-T,0], then there exists K2>k2>0K_{2}>k_{2}>0 depending on k1,K1k_{1},K_{1} so that

    k2T1ψu(v)K2T1.k_{2}T^{-1}\leq-\psi^{u}(v)\leq K_{2}T^{-1}.
  2. (2)

    K12T2Ric(γv(t))0-K_{1}^{2}T^{-2}\leq\text{Ric}(\gamma^{\prime}_{v}(t))\leq 0 for all t[0,k1T]t\in[0,k_{1}T], and k3T1ψu(v)K3T1k_{3}T^{-1}\leq-\psi^{u}(v)\leq K_{3}T^{-1}, then there exist K4>k4>0K_{4}>k_{4}>0 depending on k1,K1,k3,K3k_{1},K_{1},k_{3},K_{3} so that

    k4T1ψu(γv(t))K4T1,k_{4}T^{-1}\leq-\psi^{u}(\gamma^{\prime}_{v}(t))\leq K_{4}T^{-1},

    for all t[0,k1T].t\in[0,k_{1}T].

Proof.

Denote by u(t)=1n1tr(Uvu(t))u(t)=\frac{1}{n-1}\text{tr}(U_{v}^{u}(t)). Since UvuU_{v}^{u} is diagonalizable and all eigenvalues λi(t)\lambda_{i}(t) are nonnegative, by Cauchy-Schwartz we have

tr((Uvu)2)(n1)2u2(n1)tr((Uvu)2).\text{tr}((U_{v}^{u})^{2})\leq(n-1)^{2}u^{2}\leq(n-1)\text{tr}((U_{v}^{u})^{2}).

Thus, by the Riccati equation,

u=tr((Uvu))n1=tr((Uvu)2)n1Ric(γ˙)n1u2Ric(γ˙)n1.u^{\prime}=\frac{\text{tr}((U_{v}^{u})^{\prime})}{n-1}=-\frac{\text{tr}((U_{v}^{u})^{2})}{n-1}-\frac{\text{Ric}(\dot{\gamma})}{n-1}\leq-u^{2}-\frac{\text{Ric}(\dot{\gamma})}{n-1}.

On the other hand, denote by w(t)=(n1)u(t)=tr(Uvu(t)).w(t)=(n-1)u(t)=\text{tr}(U_{v}^{u}(t)). We have

w=tr((Uvu)2)Ric(γ˙)w2Ric(γ˙).w^{\prime}=-\text{tr}((U_{v}^{u})^{2})-\text{Ric}(\dot{\gamma})\geq-w^{2}-\text{Ric}(\dot{\gamma}).
  1. (1)

    Compare uu with the solution of

    u¯+u¯2K12T2=0,u¯(T)=+u¯(0)=K1T1cothK1.\bar{u}^{\prime}+\bar{u}^{2}-K_{1}^{2}T^{-2}=0,\,\,\bar{u}(-T)=+\infty\quad\Rightarrow\quad\bar{u}(0)=K_{1}T^{-1}\coth K_{1}.

    By the main theorem in [EH90], we have

    ψu(v)=(n1)u(0)(n1)u¯(0)=:K2T1.-\psi^{u}(v)=(n-1)u(0)\leq(n-1)\bar{u}(0)=:K_{2}T^{-1}.

    Compare ww with the solution of

    w¯+w¯2k12T2=0,w¯(T)=0w¯(0)=k1T1tanhk1.\bar{w}^{\prime}+\bar{w}^{2}-k_{1}^{2}T^{-2}=0,\,\,\bar{w}(-T)=0\quad\Rightarrow\quad\bar{w}(0)=k_{1}T^{-1}\tanh k_{1}.

    We have

    ψu(v)=w(0)w¯(0)=:k2T1.-\psi^{u}(v)=w(0)\geq\bar{w}(0)=:k_{2}T^{-1}.
  2. (2)

    Compare uu with the solution of

    u¯+u¯2K12T2=0,u¯(0)=K3T1.\bar{u}^{\prime}+\bar{u}^{2}-K_{1}^{2}T^{-2}=0,\,\,\bar{u}(0)=K_{3}T^{-1}.

    We have

    u¯(t)=K1T1coth(K1T1t+coth1(K3/K1)).\bar{u}(t)=K_{1}T^{-1}\coth(K_{1}T^{-1}t+\coth^{-1}(K_{3}/K_{1})).

    Since coth\coth is decreasing, so does u¯\bar{u}. For t[0,k1T]t\in[0,k_{1}T], we get

    ψu(γv(t))=(n1)u(t)(n1)u¯(t)(n1)u¯(0)=:K4T1.-\psi^{u}(\gamma^{\prime}_{v}(t))=(n-1)u(t)\leq(n-1)\bar{u}(t)\leq(n-1)\bar{u}(0)=:K_{4}T^{-1}.

    Compare ww with the solution of

    w¯+w¯2=0,w¯(0)=k3T1w¯(t)=(t+k31T)1.\bar{w}^{\prime}+\bar{w}^{2}=0,\,\,\bar{w}(0)=k_{3}T^{-1}\quad\Rightarrow\quad\bar{w}(t)=(t+k_{3}^{-1}T)^{-1}.

    For t[0,k1T]t\in[0,k_{1}T],

    ψu(γv(t))=w(t)w¯(t)(k1T+k31T)1=:k4T1.-\psi^{u}(\gamma^{\prime}_{v}(t))=w(t)\geq\bar{w}(t)\geq(k_{1}T+k_{3}^{-1}T)^{-1}=:k_{4}T^{-1}.

Refer to caption
(a) Case 1
Refer to caption
(b) Case 2
Figure 6.1. Proof of Proposition 6.2
Proof of Proposition 6.2.

We follow the main steps in the proof of [GW99, Lemma 3.3]. We use x,ϕx,\phi instead of xv,ϕvx_{v},\phi_{v} for simplicity.
Case 1: vv is bouncing or asymptotic. See Figure 6.1a. Since the asymptotic case is the bouncing case with t~=\tilde{t}=\infty, we only have to consider the bouncing vv. We may assume 0<x(0)ε/20<x(0)\leq\varepsilon/2. Denote by

Tv:=max{T>0:x(t)[x(0)/2,2x(0)],Tt0}.T_{v}:=\max\{T>0:x(t)\in[x(0)/2,2x(0)],\forall-T\leq t\leq 0\}.

For any t[Tv,0]t\in[-T_{v},0], we have K5x(0)mRic(γ˙(t))k5x(0)m-K_{5}x(0)^{m}\leq\text{Ric}(\dot{\gamma}(t))\leq-k_{5}x(0)^{m}. Moreover we have

Lemma 6.4.

Tvk6x(0)m/2T_{v}\geq k_{6}x(0)^{-m/2} for some k6>0k_{6}>0 independent of vv.

Proof.

Case 1: x(Tv)=2x(0)x(-T_{v})=2x(0), and x(t)x(t) is decreasing on [Tv,0][-T_{v},0]. By (4.4) with α=m+2,β=1/2\alpha=m+2,\beta=1/2 , we have

x(0)m/2(2x(0))m/2+K6Tv.x(0)^{-m/2}\leq(2x(0))^{-m/2}+K_{6}T_{v}.

Thus Tvk6x(0)m/2T_{v}\geq k_{6}x(0)^{-m/2}.

Case 2: x(Tv)=x(0)/2x(-T_{v})=x(0)/2, similar to Case 1.

Case 3: x(Tv)=2x(0)x(-T_{v})=2x(0), and x(t)x(t) first decreases, then increases on [Tv,0][-T_{v},0]. Assume T3(0,Tv)T_{3}\in(0,T_{v}) satisfies x(T3)=x(0)x(-T_{3})=x(0). By Case 1 we know that Tv>TvT3k6x(0)m/2T_{v}>T_{v}-T_{3}\geq k_{6}x(0)^{-m/2}. ∎

Take T=k6x(0)m/2T=k_{6}x(0)^{-m/2} in Lemma 6.3(1), we get

k7x(0)m/2ψu(v)K7x(0)m/2.k_{7}x(0)^{m/2}\leq-\psi^{u}(v)\leq K_{7}x(0)^{m/2}.

By (3.6), we know that

|ϕ(0)|m/(m+2)|2sinϕ(0)|m/(m+2)=(2x(0))m/(m+2)Qx(0)m/2.|\phi(0)|^{m/(m+2)}\leq|2\sin\phi(0)|^{m/(m+2)}=(-2x^{\prime}(0))^{m/(m+2)}\leq Qx(0)^{m/2}.

Thus, we finish the proof of Proposition 6.2 in this case.
Case 2: vv is crossing. See Figure 6.1b. Recall that x(t0)=0x(t_{0})=0, and a(t)=sinϕ(t)a(t)=-\sin\phi(t). Denote by A:=a(t0)A:=a(t_{0}). Since vv is close to Sing\mathrm{Sing}, we may assume that 2A<ε(m+2)/22A<\varepsilon^{(m+2)/2} and 0x(0)<A2/(m+2)0\leq x(0)<A^{2/(m+2)}. Since x=sinϕx^{\prime}=\sin\phi, by (4.10) we have

x(0)m/2Am/(m+2)|x(0)|m/(m+2)|ϕ(0)|m/(m+2),andx(0)^{m/2}\leq A^{m/(m+2)}\leq|x^{\prime}(0)|^{m/(m+2)}\leq|\phi(0)|^{m/(m+2)},\text{and}
A|x(0)|=|sinϕ(0)|A2+K8x(0)m+2A1+K8.A\leq|x^{\prime}(0)|=|\sin\phi(0)|\leq\sqrt{A^{2}+K_{8}x(0)^{m+2}}\leq A\sqrt{1+K_{8}}.

Thus, it suffices to prove

(6.4) k9Am/(m+2)ψu(v)K9Am/(m+2).k_{9}A^{m/(m+2)}\leq-\psi^{u}(v)\leq K_{9}A^{m/(m+2)}.

We prove (6.4) in the following two cases:

  • Case 2a: ϕ(0)<0\phi(0)<0. In this case, t0>0t_{0}>0 and x<0x^{\prime}<0 for t[0,t0]t\in[0,t_{0}]. Let Tv,tv>0T_{v},t_{v}>0 be the minimal solutions of

    x(Tv)=(2A)2/(m+2),x(tv)=A2/(m+2).x(-T_{v})=(2A)^{2/(m+2)},\quad x(-t_{v})=A^{2/(m+2)}.

    By the choice of AA, we have x(tv)<x(Tv)<εx(-t_{v})<x(-T_{v})<\varepsilon. For t[Tv,tv]t\in[-T_{v},-t_{v}], by (4.10) and (4.4) with α=2,β=m+1m+2\alpha=2,\beta=\frac{m+1}{m+2} , we have

    Amm+2(2A)mm+2+K10(Tvtv).A^{-\frac{m}{m+2}}\leq(2A)^{-\frac{m}{m+2}}+K_{10}(T_{v}-t_{v}).

    Thus Tvtv>k10Am/(m+2)T_{v}-t_{v}>k_{10}A^{-m/(m+2)}. Take T=k10Am/(m+2)T=k_{10}A^{-m/(m+2)} in Lemma 6.3(1), we have

    (6.5) k11Am/(m+2)ψu(γv(tv))K11Am/(m+2).k_{11}A^{m/(m+2)}\leq-\psi^{u}(\gamma^{\prime}_{v}(-t_{v}))\leq K_{11}A^{m/(m+2)}.

    Since x=sinϕx^{\prime}=\sin\phi, we have ϕ=x′′secϕ0\phi^{\prime}=x^{\prime\prime}\sec\phi\geq 0, thus |x|=|sinϕ||sinϕ(t0)|=A|x^{\prime}|=|\sin\phi|\geq|\sin\phi(t_{0})|=A for all t[tv,t0]t\in[-t_{v},t_{0}]. Thus,

    (6.6) (tv+t0)Atvt0x(s)ds=x(t0)+x(tv)=A2/(m+2).(t_{v}+t_{0})A\leq\int_{-t_{v}}^{t_{0}}-x^{\prime}(s)ds=-x(t_{0})+x(-t_{v})=A^{2/(m+2)}.

    Thus tvAm/(m+2)t_{v}\leq A^{-m/(m+2)}. By (6.5), take T=Am/(m+2)T=A^{-m/(m+2)} in Lemma 6.3(2), we get

    (6.7) k12Am/(m+2)ψu(v)K12Am/(m+2).k_{12}A^{m/(m+2)}\leq-\psi^{u}(v)\leq K_{12}A^{m/(m+2)}.
  • Case 2b: ϕ(0)>0\phi(0)>0. In this case, we have t0<0t_{0}<0. Since γv(t0)\gamma^{\prime}_{v}(t_{0}) crosses T0T_{0} and we do not have flat strips, by the result of Case 2a, we know that

    (6.8) k12Am/(m+2)ψu(γv(t0))K12Am/(m+2),k_{12}A^{m/(m+2)}\leq-\psi^{u}(\gamma^{\prime}_{v}(t_{0}))\leq K_{12}A^{m/(m+2)},

    and |t0|Am/(m+2)|t_{0}|\leq A^{-m/(m+2)} by (6.6). Again take T=Am/(m+2)T=A^{-m/(m+2)} in Lemma 6.3(2), we have

    k13Am/(m+2)ψu(v)K13Am/(m+2).k_{13}A^{m/(m+2)}\leq-\psi^{u}(v)\leq K_{13}A^{m/(m+2)}.

Remark 6.5.

The absence of flat strips is crucial in Proposition 6.2 and Theorem C; otherwise, the Hölder continuity does not hold. Here is a counterexample: consider a surface of revolution generated by

f(x)={|x+0.5|m+2+1,x<0.5,1,0.5x0.5,(x0.5)m+2+1,x>0.5.f(x)=\begin{cases}|x+0.5|^{m+2}+1,&x<-0.5,\\ 1,&-0.5\leq x\leq 0.5,\\ (x-0.5)^{m+2}+1,&x>0.5.\end{cases}

The flat strip is the part with 0.5x0.5-0.5\leq x\leq 0.5, and the metric satisfies both curvature conditions by Lemma A.1. Let vv be a unit vector with x=0.5x=0.5 and angle ϕ>0\phi>0, meaning that vv is a vector exiting the flat strip. At time t=cscϕt=-\csc\phi, the geodesic γv\gamma_{v} enters the strip with vector v=γv(cscϕ)v^{\prime}=\gamma^{\prime}_{v}(-\csc\phi). By symmetry, the angle of vv^{\prime} is ϕ-\phi. Assume that the Hölder continuity in Proposition 6.2 holds for both vv and vv^{\prime}, namely, ψu(v),ψu(v)ϕm/(m+2)\psi^{u}(v),\psi^{u}(v^{\prime})\approx-\phi^{m/(m+2)}. As γv(t)\gamma_{v}(t) is in the flat strip for t[cscϕ,0]t\in[-\csc\phi,0], ψu(ftv)\psi^{u}(f_{t}v) satisfies the Riccati equation u+u2=0u^{\prime}+u^{2}=0, and the solution is

ψu(ftv)=(t+ψu(v)1)1.\psi^{u}(f_{t}v)=(t+\psi^{u}(v)^{-1})^{-1}.

Plug in ψu(v)ϕm/(m+2)\psi^{u}(v)\approx-\phi^{m/(m+2)} and t=cscϕt=-\csc\phi, we have

ψu(v)=(cscϕ+ψu(v)1)1ϕ.\psi^{u}(v^{\prime})=(-\csc\phi+\psi^{u}(v)^{-1})^{-1}\approx-\phi.

Contradictory to ψu(v)ϕm/(m+2)\psi^{u}(v^{\prime})\approx-\phi^{m/(m+2)}.

The Hölder continuity of φu\varphi^{u} is an important, yet still open, question in nonpositively curved geometry. Only some partial results are known for surfaces under certain conditions, including [GW99, Lemma 3.3] where Gerber and Wilkinson show the Hölder continuity of φu\varphi^{u} for type 2 surfaces. Since Ricci curvature and Gaussian curvature are the same thing for surfaces, using Theorem C we obtain a partial generalization of [GW99, Lemma 3.3]:

Corollary 6.6.

Under the same assumptions as Theorem C, ψu\psi^{u} and φu\varphi^{u} are Hölder continuous in a small neighborhood of Sing\mathrm{Sing}.

7. Sufficient criteria for the pressure gap

Let MM be a closed Riemannian manifold and {ft}t\{f_{t}\}_{t\in\mathbb{R}} the geodesic flow on T1MT^{1}M. In this section, we will describe an abstract result to establish the pressure gap for a given potential φ:T1M\varphi\colon T^{1}M\to\mathbb{R}. But first, we need to introduce the notion of specification in the following subsection.

7.1. Specification

While there are various definitions for it in the literature, roughly speaking specification is a property that allows one to find an orbit segment that shadows any given finite number of orbit segments at a desired scale with controlled transition time. It was introduced by Bowen [Bow74] as one of the conditions to establish the uniqueness of the equilibrium states for potentials over uniformly hyperbolic maps. The specification still plays a vital role in many generalizations of this result [BCFT18, CKP20, CKP21]. The following version of the specification is from [BCFT18, Theorem 4.1].

Definition 7.1 (Specification).

We say a set of orbit segments 𝒞\mathcal{C} satisfies the specification at scale ρ>0\rho>0 if there exists 𝒯>0\mathcal{T}>0 such that given finite orbit segments (v1,t1),,(vk,tk)𝒞(v_{1},t_{1}),\ldots,(v_{k},t_{k})\in\mathcal{C} and T1,,TkT_{1},\ldots,T_{k}\in\mathbb{R} with Tj+1Tjtj+𝒯T_{j+1}-T_{j}\geq t_{j}+\mathcal{T} for all 1jk11\leq j\leq k-1, there is wT1Mw\in T^{1}M such that fTjwBtj(vj,ρ)f_{T_{j}}w\in B_{t_{j}}(v_{j},\rho) for all 1jk11\leq j\leq k-1.

This is a stronger version of the specification that appears in [BCFT18] providing flexibility in the transition time. However, in practice, we will always take TjT_{j}’s such that Tj+1Tj=tj+𝒯T_{j+1}-T_{j}=t_{j}+\mathcal{T}; that is, the transition time is exactly equal to 𝒯\mathcal{T}.

7.2. Abstract result on the pressure gap

We now list the conditions together which establish the pressure gap. Let MM be a closed rank 1 manifold with a codimension 1 flat subtorus, and Sing\mathrm{Sing} are induced by the subtorus. As mentioned in the introduction, one can easily extend results in this section to multiple subtori scenarios. However, for brevity, we stick to this simpler assumption.

By setting

𝒞(η)={(v,t)T1M×+:v,ftvReg(η)}\mathcal{C}(\eta)=\{(v,t)\in T^{1}M\times\mathbb{R}^{+}\colon v,f_{t}v\in\mathrm{Reg}(\eta)\}

to be the set of orbit segments with endpoints in Reg(η)\mathrm{Reg}(\eta), we require that the geodesic flow {ft}\{f_{t}\} and the potential φ:T1M\varphi\colon T^{1}M\to\mathbb{R} satisfy the following property:

  1. (1)

    Singular set zero entropy property: htop(|Sing)=0h_{\mathrm{top}}(\mathcal{F}|_{\mathrm{Sing}})=0.

  2. (2)

    Specification property: For any η>0\eta>0 and ρ>0\rho>0, the orbit segments 𝒞(η)\mathcal{C}(\eta) satisfies the specification at scale ρ\rho.

  3. (3)

    Shadowing property: There exists R>0R>0 such that for every t>0t>0 there exists a map

    Πt,R:SingT1M\Pi_{t,R}\colon\mathrm{Sing}\to T^{1}M

    with the following properties: denoting by w𝗏:=Πt,R(𝗏)w_{\mathsf{v}}:=\Pi_{t,R}(\mathsf{v}) the shadowing vector of an arbitrary 𝗏Sing\mathsf{v}\in\mathrm{Sing},

    1. (a)

      The conclusion of Lemma 3.1 hold for Πt,R\Pi_{t,R}.

    2. (b)

      For any ε>0\varepsilon>0, there exists L:=L(ε,R)L:=L(\varepsilon,R) such that for any t>2Lt>2L, the vector w𝗏w_{\mathsf{v}} satisfies

      d(fτw𝗏,Sing)<εd(f_{\tau}w_{\mathsf{v}},\mathrm{Sing})<\varepsilon

      for all τ[L,tL]\tau\in[L,t-L].

  4. (4)

    Special Bowen property: For any δ>0\delta>0, there exists C=C(δ,R)>0C=C(\delta,R)>0 independent of tt such that

    0tφ(fτu)𝑑τ0tφ(fτ𝗏)𝑑τ>C\int_{0}^{t}\varphi(f_{\tau}u)d\tau-\int_{0}^{t}\varphi(f_{\tau}\mathsf{v})d\tau>-C

    for any uBt(w𝗏,δ)u\in B_{t}(w_{\mathsf{v}},\delta).

Proposition 7.2.

Suppose the geodesic flow {ft}t\{f_{t}\}_{t\in\mathbb{R}} and the potential φ\varphi satisfy the above listed conditions (1), (2) and (3). Then, φ\varphi has a pressure gap.

Assume we have Proposition 7.2, we finish the proof of Theorem A:

Proof of Theorem A.

Each condition listed above can be verified as follows. By design, we know htop(|Sing)=0h_{\mathrm{top}}(\mathcal{F}|_{\mathrm{Sing}})=0. The specification property (2) is already established in [BCFT18] for the geodesic flow {ft}\{f_{t}\} over rank 1 nonpositively curved manifold. With Πt,R\Pi_{t,R} defined as in (3.1) via the Fermi coordinates, (3a) is immediate as we have already proved Lemma 3.1. For (3b), we can take LL to be (Q/ε)m/2(Q/\varepsilon)^{m/2} where QQ is the constant from Proposition 4.2 and 5.3. Lastly, (4) follows from Proposition 3.3. Hence, φ\varphi has the pressure gap by the above proposition. ∎

We note that the proof of Proposition 7.2 draws inspiration from [BCFT18, Theorem B]. However, leveraging the singular set’s zero entropy property, Peres’ Lemma allows us to circumvent several technicalities and arrive at a more straightforward proof than that presented in [BCFT18]. See Remark 7.4 for more details.

Theorem 7.3.

[Per88, Lemma 2] Let ={ft}\mathcal{F}=\{f_{t}\} be a continuous flow on a compact space XX, and μ\mu be an \mathcal{F}-invariant probability measure. Then for every potential φ:X\varphi:X\to\mathbb{R} there exists some vXv\in X

1T0Tφ(fτv)𝑑τXφ𝑑μ\frac{1}{T}\int_{0}^{T}\varphi(f_{\tau}v)d\tau\geq\int_{X}\varphi d\mu

for all T>0T>0.

The authors believe this Peres’ result is known among the experts; however, we cannot find proof of the about flow version in the literature. For the completeness, we give a proof in the appendix; see Theorem B.1.

Proof of Proposition 7.2.

This proof has three steps. The first step is using the singular set zero entropy property (1) and Theorem 7.3 to bound P(φ,Sing)P(\varphi,\mathrm{Sing}) by the integration of some special 𝗏\mathsf{v} in Sing\mathrm{Sing} along the flow. The second step is using the specification property (2) and shadowing property (3) to create a (t,δ)(t,\delta)-separated set that bounds P(φ,Sing)P(\varphi,\mathrm{Sing}). The last step is using the special Bowen property (4) to estimate the pressure.

Notice that without lost of generality, we may assume P(φ,Sing)=0P(\varphi,\mathrm{Sing})=0; otherwise, we consider φP(φ,Sing)\varphi-P(\varphi,\mathrm{Sing}). Hence, it is sufficient to show that there exists a (t,δ)(t,\delta)-separated set FtF_{t} on T1MT^{1}M such that

(7.1) limt1twFte0tφ(fτw)𝑑τ>0.\lim_{t\to\infty}\frac{1}{t}\sum_{w\in F_{t}}e^{\int_{0}^{t}\varphi(f_{\tau}w)d\tau}>0.

The first step is to find a special singular vector 𝗏\mathsf{v} for shadowing. Since \mathcal{F} on Sing\mathrm{Sing} is entropy-expansive, we know there exists an equilibrium state μ(|Sing)\mu\in\mathcal{M}(\mathcal{F}|_{\mathrm{Sing}}) of φ\varphi. That is

0=P(φ,Sing)=hμ()+Singφ𝑑μ.0=P(\varphi,\mathrm{Sing})=h_{\mu}(\mathcal{F})+\int_{\mathrm{Sing}}\varphi d\mu.

By the singular set zero entropy property (1), we know hμ()=0h_{\mu}(\mathcal{F})=0; and thus, Singφ𝑑μ=0\int_{\mathrm{Sing}}\varphi\ d\mu=0. By Peres’ Lemma (Theorem 7.3), there exists 𝗏Sing\mathsf{v}\in\mathrm{Sing} such that for all T>0T>0

(7.2) 0Tφ(fτ𝗏)𝑑τTSingφ𝑑μ=0.\int_{0}^{T}\varphi(f_{\tau}\mathsf{v})d\tau\geq T\int_{\mathrm{Sing}}\varphi\ d\mu=0.

The second step is to create such a (t,δ)(t,{\color[rgb]{1,0,0}\delta})-separated set FtF_{t} on T1MT^{1}M. For R>0R>0 given in the shadowing property (3), from Lemma 3.1 we get η>0\eta>0 where w,ftwReg(η)w,f_{t}w\in\mathrm{Reg}(\eta) for any w=Πt,R(𝗏)w=\Pi_{t,R}(\mathsf{v}). We also obtain the constant LL from (3b) corresponding to ε=R/2\varepsilon=R/2.

We now set ξ\xi be a constant such that ξ>𝒯+2L\xi>\mathcal{T}+2L, and for each NN\in\mathbb{N} consider

𝒜:={ξ,2ξ,,(N1)ξ}[0,Nξ]\mathcal{A}:=\{\xi,2\xi,\ldots,(N-1)\xi\}\subseteq[0,N\xi]

For any small α>0\alpha>0 such that αN\alpha N\in\mathbb{N}, consider any size (αN1)(\alpha N-1) subset

J={N1ξ,,NαN1ξ}𝒜J=\{N_{1}\xi,\ldots,N_{\alpha N-1}\xi\}\subset\mathcal{A}

with Ni{1,,N1}N_{i}\in\{1,\ldots,N-1\}. Setting N0=0N_{0}=0 and NαN=NN_{\alpha N}=N, such a subset can be viewed as a partition of the interval [0,Nξ][0,N\xi] into αN\alpha N subintervals, each of length niξn_{i}\xi for ni=NiNi1n_{i}=N_{i}-N_{i-1}. We denote 𝕁Nα={J𝒜:#J=αN1}\mathbb{J}_{N}^{\alpha}=\{J\subset\mathcal{A}:\ \#J=\alpha N-1\}, and we know #𝕁Nα=(N1αN1)\#\mathbb{J}_{N}^{\alpha}={N-1\choose\alpha N-1}.

Given J𝕁NαJ\in\mathbb{J}_{N}^{\alpha}, we consider singular vectors 𝗏iJ:=fNiξ(𝗏)\mathsf{v}_{i}^{J}:=f_{N_{i}\xi}(\mathsf{v}) and the corresponding regular orbits {(wiJ,ti)}\{(w_{i}^{J},t_{i})\} where wiJ:=Πniξ𝒯,R(𝗏iJ)w_{i}^{J}:=\Pi_{n_{i}\xi-\mathcal{T},R}(\mathsf{v}_{i}^{J}) and ti=niξ𝒯t_{i}=n_{i}\xi-\mathcal{T} for i=1,,αNi=1,\cdots,\alpha N. Using the specification property (2), one can find a vector wJT1Mw_{J}\in T^{1}M which shadows orbits {(wiJ,ti)}i=1αN1\{(w_{i}^{J},t_{i})\}_{i=1}^{\alpha N-1} (see Fig. 7.1), that is, for i=1,,αNi=1,\cdots,\alpha N

fNi1ξwJBti(wiJ,ρ).f_{N_{i-1}\xi}w_{J}\in B_{t_{i}}(w_{i}^{J},\rho).
Refer to caption
Figure 7.1. Shadowing orbit

We claim that FNξ:={(wJ,Nξ):J𝕁Nα}F_{N\xi}:=\{(w_{J},N\xi):J\in\mathbb{J}_{N}^{\alpha}\} is a (Nξ,R/22ρ)(N\xi,R/2-2\rho)-separated set. To see this, let Jj={N1jξ,,NαN1jξ}J_{j}=\{N_{1}^{j}\xi,\cdots,N_{\alpha N-1}^{j}\xi\} for j=1,2j=1,2, and m=min{n:Nn1Nn2}.m=\min\{n:\ N_{n}^{1}\neq N_{n}^{2}\}. Without loss of generality, suppose Nm1<Nm2N_{m}^{1}<N_{m}^{2}, then the claim follows the inequalities below (see Fig. 7.2 ):

d(fNm1𝒯(wJ1),Sing)>Rρ and d(fNm1𝒯(wJ2),Sing)<R2+ρ.d(f_{N_{m}^{1}-\mathcal{T}}(w_{J_{1}}),\mathrm{Sing})>R-\rho\text{ and }d(f_{N_{m}^{1}-\mathcal{T}}(w_{J_{2}}),\mathrm{Sing})<\frac{R}{2}+\rho.
Refer to caption
Figure 7.2. Separated set

The last step is using the special Bowen property (4) to bound the pressure from below. For (wJ,Nξ)FNξ(w_{J},N\xi)\in F_{N\xi}, the integral of φ(fτwJ)\varphi(f_{\tau}w_{J}) during each transition period is bounded below by 𝒯φ-\mathcal{T}\|\varphi\|, by the special Bowen property (4) and (7.2) we get

(7.3) 0Nξφ(fτwJ)𝑑τ\displaystyle\int_{0}^{N\xi}\varphi(f_{\tau}w_{J})d\tau =i=1αNNi1ξNiξφ(fτwJ)𝑑τ\displaystyle=\sum_{i=1}^{\alpha N}\int_{N_{i-1}\xi}^{N_{i}\xi}\varphi(f_{\tau}w_{J})d\tau
i=1αN0tiφ(fτ𝗏iJ)𝑑ταNC2αNφ𝒯\displaystyle\geq\sum_{i=1}^{\alpha N}\int_{0}^{t_{i}}\varphi(f_{\tau}\mathsf{v}_{i}^{J})d\tau-\alpha NC-2\alpha N||\varphi||\mathcal{T}
i=1αNNi1ξNiξφ(fτ𝗏iJ)𝑑ταN(3𝒯φ+C)\displaystyle\geq\sum_{i=1}^{\alpha N}\int_{N_{i-1}\xi}^{N_{i}\xi}\varphi(f_{\tau}\mathsf{v}_{i}^{J})d\tau-\alpha N(3\mathcal{T}\|\varphi\|+C)
=0Nξφ(fτ𝗏)𝑑ταN(3𝒯φ+C)αN(3𝒯φ+C).\displaystyle=\int_{0}^{N\xi}\varphi(f_{\tau}\mathsf{v})d\tau-\alpha N(3\mathcal{T}\|\varphi\|+C)\geq-\alpha N(3\mathcal{T}\|\varphi\|+C).

Recall that #FNξ=#𝕁Nα=(N1αN1)\#F_{N\xi}=\#\mathbb{J}_{N}^{\alpha}={N-1\choose\alpha N-1}, summing the above inequality over all possible subsets JJ gives

wJFNξe0Nξφ(fτwJ)𝑑τ\displaystyle\sum_{w_{J}\in F_{N\xi}}e^{\int_{0}^{N\xi}\varphi(f_{\tau}w_{J})d\tau} (N1αN1)eαN(3𝒯φ+C).\displaystyle\geq{N-1\choose\alpha N-1}e^{-\alpha N(3\mathcal{T}\|\varphi\|+C)}.

As FNξF_{N\xi} is a (Nξ,δ)(N\xi,\delta)-separated set (where δ<R/2ρ\delta<R/2-\rho with RR and ρ\rho are arbitrary), this implies that

P(φ)limN1Nξlog(wJFNξe0Nξφ(fτwJ)𝑑τ).{\displaystyle P(\varphi)\geq\lim_{N\to\infty}\frac{1}{N\xi}\log\Big{(}\sum_{w_{J}\in F_{N\xi}}e^{\int_{0}^{N\xi}\varphi(f_{\tau}w_{J})d\tau}\Big{)}.}

Combining the last two inequalities with (N1αN1)αe(αlogα)N{N-1\choose\alpha N-1}\geq\alpha e^{\left(-\alpha\log\alpha\right)N}, one establishes the pressure gap for φ\varphi by taking α\alpha sufficiently small. ∎

Remark 7.4.

We list below several main differences between this proof and the proof of [BCFT18, Theorem B].

  1. (1)

    Our proof utilize htop(|Sing)=0h_{\mathrm{top}}(\mathcal{F}|_{\mathrm{Sing}})=0 fact and Peres’ Lemma (Theorem 7.3) to simplify two steps (that is [BCFT18, Sec. 8.2 & 8.3] in the [BCFT18, Theorem B]. Precisely, using htop(|Sing)=0h_{\mathrm{top}}(\mathcal{F}|_{\mathrm{Sing}})=0 and Peres’ Lemma, we can easily find a vector in the singular set to shadow regular orbit segments with good control of the integration of potentials.

  2. (2)

    The equation (7.3) is another major difference. The extra constant CC in (7.3) comes from the condition (4) given in Section 7.2. This constant CC allows us to have some wiggling room for the potential without losing the pressure gap.

Appendix A Ricci curvature bound and radial curvature bound comparison

Recall that MM is a nn-dimensional manifold with nonpositive sectional curvature, and T0MT_{0}\subset M is a totally geodesic (n1)(n-1)-subtorus. We assume that K(σ)=0K(\sigma)=0 for any xT0x\in T_{0} and any 2-plane σTxM\sigma\subset T_{x}M. On the universal cover M~\widetilde{M}, we define the Fermi coordinate (s,x)(s,x) near T~0\widetilde{T}_{0} in the following way: ss is the coordinate on T~0\widetilde{T}_{0}, and xx measures the signed distance on M~\widetilde{M} to T~0\widetilde{T}_{0}. s=const.s=\text{const.} is always a geodesic perpendicular to T~0\widetilde{T}_{0}. The Riemannian metric near T~0\widetilde{T}_{0} is

(A.1) g=dx2+gx,|x|εg=dx^{2}+g_{x},\quad|x|\leq\varepsilon

where gxg_{x} is the Riemannian metric on T~x:=T~0×{x}\widetilde{T}_{x}:=\widetilde{T}_{0}\times\{x\}. In particular, g0g_{0} is the Euclidean metric on T~0.\widetilde{T}_{0}.

If gxg_{x} is a warped product, namely, g=dx2+f(x)2g0g=dx^{2}+f(x)^{2}g_{0}. We have f(0)=1f(0)=1. Since T0T_{0} is totally geodesic, we have f(0)=0f^{\prime}(0)=0.

Lemma A.1.

If g=dx2+f(x)2g0g=dx^{2}+f(x)^{2}g_{0}, then the following conditions are equivalent:

  1. (1)

    There exists C1,C2,ε>0C_{1},C_{2},\varepsilon>0 such that

    C1|xv|mK(v)C2|xv|m,for any vX with |xv|<ε.-C_{1}|x_{v}|^{m}\leq K_{\perp}(v)\leq-C_{2}|x_{v}|^{m},\text{for any $v\perp X$ with }|x_{v}|<\varepsilon.
  2. (2)

    There exists C1,C2,ε>0C_{1},C_{2},\varepsilon>0 so that

    C1|xv|mRic(v)C2|xv|m,for any v with |xv|<ε.-C_{1}|x_{v}|^{m}\leq\text{Ric}(v)\leq-C_{2}|x_{v}|^{m},\text{for any $v$ with }|x_{v}|<\varepsilon.
  3. (3)

    There exists C1,C2,ε>0C_{1},C_{2},\varepsilon>0 so that

    C1|x|m+2f(x)1C2|x|m+2.C_{1}|x|^{m+2}\leq f(x)-1\leq C_{2}|x|^{m+2}.
Proof.

Since g=dx2+f(x)2g0g=dx^{2}+f(x)^{2}g_{0}, the radial curvature is

K(v)=f′′(xv)f(xv),K_{\perp}(v)=-\frac{f^{\prime\prime}(x_{v})}{f(x_{v})},

while the sectional curvature is given by

Kσ=f′′fcos2θ(ff)2sin2θ,K_{\sigma}=-\frac{f^{\prime\prime}}{f}\cos^{2}\theta-\left(\frac{f^{\prime}}{f}\right)^{2}\sin^{2}\theta,

where θ\theta be the angle between XX and σ\sigma.

Now we compute the Ricci curvature. If vv is normal, then Ric(v)=(n1)K(v)\text{Ric}(v)=(n-1)K_{\perp}(v) and we are done. Otherwise, denote by vv^{\perp} the perpendicular complement in TpMT_{p}M, θ\theta the angle between vv and XX, and H(v)H(v) the horizontal subspace at vv. Since both vv^{\perp} and H(v)H(v) have codimension 1, vH(v)v^{\perp}\cap H(v) has dimension n2n-2. We can construct an orthonormal basis {ei}i=1n\{e_{i}\}_{i=1}^{n} such that e1=ve_{1}=v, e3,,envH(v)e_{3},\cdots,e_{n}\in v^{\perp}\cap H(v), and the section spanned by e1,e2e_{1},e_{2} is normal. Then we have

(A.2) Ric(v)=f′′f+(n2)(f′′fcos2θ(ff)2sin2θ)\text{Ric}(v)=-\frac{f^{\prime\prime}}{f}+(n-2)\left(-\frac{f^{\prime\prime}}{f}\cos^{2}\theta-\left(\frac{f^{\prime}}{f}\right)^{2}\sin^{2}\theta\right)

When θ=0\theta=0, we get Ric(v)=(n1)K(v)\text{Ric}(v)=(n-1)K_{\perp}(v). Thus, for any vv, the Ricci curvature can be calculated using (A.2).

(1)(2):(1)\Rightarrow(2): Since K=f′′/fK_{\perp}=-f^{\prime\prime}/f and f(0)=1f(0)=1, we have f′′|x|mf^{\prime\prime}\approx|x|^{m}. Since f(0)=0f^{\prime}(0)=0, f|x|m+1f^{\prime}\approx|x|^{m+1}. Therefore, Ric(v)|xv|m\text{Ric}(v)\approx-|x_{v}|^{m} by (A.2).

(2)(3):(2)\Rightarrow(3): Assume f1|x|kf-1\approx|x|^{k} for some k>2k>2. We have (f/f)2|x|2k2(f^{\prime}/f)^{2}\approx|x|^{2k-2} and f′′/f|x|k2f^{\prime\prime}/f\approx|x|^{k-2}. Thus, f′′/ff^{\prime\prime}/f is the dominant term in (A.2). Therefore, k=m+2k=m+2.

(3)(1):(3)\Rightarrow(1): Since f1xm+2f-1\approx x^{m+2} and f(0)=1f^{\prime}(0)=1, we have fxm+1f^{\prime}\approx x^{m+1} and f′′xmf^{\prime\prime}\approx x^{m}, thus K=f′′/f|x|mK_{\perp}=-f^{\prime\prime}/f\approx-|x|^{m}. ∎

Appendix B A lemma of Peres

In this section, we prove a proof of Theorem 7.3 as Peres’ original theorem [Per88, Lemma 2] is for transformations.

Theorem B.1.

[Per88, Lemma 2] Let ={ft}\mathcal{F}=\{f_{t}\} be a continuous flow on a compact space XX, and μ\mu be an \mathcal{F}-invariant probability measure. Then for every potential φ:X\varphi:X\to\mathbb{R} there exists some vXv\in X

1T0Tφ(fτv)𝑑τXφ𝑑μ\frac{1}{T}\int_{0}^{T}\varphi(f_{\tau}v)d\tau\geq\int_{X}\varphi d\mu

for all T>0T>0.

Peres’ proof is based on the Maximal Ergodic Theorem, and it works almost line by line in the flow case. A version of the Maximal Ergodic Theorem for flows can be found in [Pet83, Theorem 1.2, P.76].

Theorem B.2 (Maximal Ergodic Theorem).

Let (X,,μ)(X,\mathcal{B},\mu) be a probability space and ={ft}\mathcal{F}=\{f_{t}\} be a measure-preserving flow on XX. If φL1(μ)\varphi\in L^{1}(\mu) and α\alpha\in\mathbb{R}, then

{φ>α}φ𝑑μαμ{φ>α}\int_{\{\varphi^{*}>\alpha\}}{\varphi d\mu}\geq\alpha\cdot\mu\{\varphi^{*}>\alpha\}

where φ(x)=supT>01T0Tφ(fτ(x))𝑑τ.{\displaystyle{\varphi^{*}(x)=\sup_{T>0}\frac{1}{T}\int_{0}^{T}{\varphi(f_{\tau}(x))d\tau}}.}

Proof of Theorem B.1.

For ϵ>0\epsilon>0, we define

Eϵ:={xX:t0,1t0tφ(fτx)𝑑τ>Xφ𝑑με}E_{\epsilon}:=\{x\in X:\ \forall t\geq 0,\ \frac{1}{t}\int_{0}^{t}\varphi(f_{\tau}x)d\tau>\int_{X}\varphi d\mu-\varepsilon\}

and

Ψ(x):=Xφ𝑑μφϵ.\Psi(x):=\int_{X}\varphi d\mu-\varphi-\epsilon.

It is enough to show

ϵ>0Eϵ.\bigcap_{\epsilon>0}E_{\epsilon}\neq\emptyset.

To see this, we first notice that Eϵ={x:Ψ(x)0}E_{\epsilon}=\{x:\ \Psi^{*}(x)\leq 0\}. We then apply the Maximal Ergodic Theorem on Ψ\Psi and X\Eϵ={x:Ψ(x)>0}X\backslash E_{\epsilon}=\{x:\ \Psi^{*}(x)>0\}, i.e., α=0\alpha=0, and get

X\EϵΨ𝑑μ0.\int_{X\backslash E_{\epsilon}}\Psi d\mu\geq 0.

Observe that XΨ𝑑μ=ϵ\int_{X}\Psi d\mu=-\epsilon, and which guarantees EϵE_{\epsilon}\neq\emptyset for all ϵ>0\epsilon>0. Since φ\varphi is continuous and XX is compact, we know EϵE_{\epsilon}’s are nested compact sets, and thus

ϵ>0Eϵ.\bigcap_{\epsilon>0}E_{\epsilon}\neq\emptyset.

References

  • [ALP] Ermerson Araujo, Yuri Lima, and Mauricio Poletti, Symbolic dynamics for nonuniformly hyperbolic maps with singularities in high dimension, preprint, arXiv:2010.11808.
  • [Bal95] Werner Ballmann, Lectures on spaces of nonpositive curvature, vol. 25, Springer Science & Business Media, 1995.
  • [BBE85] Werner Ballmann, Misha Brin, and Patrick Eberlein, Structure of manifolds of nonpositive curvature. I, Ann. of Math. (2) 122 (1985), no. 1, 171–203.
  • [BBFS21] Keith Burns, Jérôme Buzzi, Todd Fisher, and Noelle Sawyer, Phase transitions for the geodesic flow of a rank one surface with nonpositive curvature, Dynamical Systems 36 (2021), no. 3, 527–535.
  • [BCFT18] Keith Burns, Vaughn Climenhaga, Todd Fisher, and Daniel J Thompson, Unique equilibrium states for geodesic flows in nonpositive curvature, Geometric and Functional Analysis 28 (2018), no. 5, 1209–1259.
  • [BG89] Keith Burns and Marlies Gerber, Real analytic Bernoulli geodesic flows on S2S^{2}, Ergodic Theory Dynam. Systems 9 (1989), no. 1, 27–45.
  • [BG14] Keith Burns and Katrin Gelfert, Lyapunov spectrum for geodesic flows of rank 1 surfaces, Discrete & Continuous Dynamical Systems 34 (2014), no. 5, 1841–1872.
  • [Bow74] Rufus Bowen, Some systems with unique equilibrium states, Mathematical systems theory 8 (1974), no. 3, 193–202.
  • [CCE+] Benjamin Call, David Constantine, Alena Erchenko, Noelle Sawyer, and Grace Work, Unique equilibrium states for geodesic flows on flat surfaces with singularities, preprint, arxiv:2101.11806.
  • [CKP20] Dong Chen, Lien-Yung Kao, and Kiho Park, Unique equilibrium states for geodesic flows over surfaces without focal points, Nonlinearity 33 (2020), no. 3, 1118–1155.
  • [CKP21] by same author, Properties of equilibrium states for geodesic flows over manifolds without focal points, Advances in Mathematics 380 (2021), 107564.
  • [CKW21] Vaughn Climenhaga, Gerhard Knieper, and Khadim War, Uniqueness of the measure of maximal entropy for geodesic flows on certain manifolds without conjugate points, Adv. Math. 376 (2021), Paper No. 107452, 44.
  • [CS14] Yves Coudène and Barbara Schapira, Generic measures for geodesic flows on nonpositively curved manifolds, J. Éc. polytech. Math. 1 (2014), 387–408. MR 3322793
  • [CT16] Vaughn Climenhaga and Daniel J. Thompson, Unique equilibrium states for flows and homeomorphisms with non-uniform structure, Adv. Math. 303 (2016), 745–799.
  • [CT21] by same author, Beyond Bowen’s specification property, Thermodynamic formalism, Lecture Notes in Math., vol. 2290, Springer, Cham, 2021, pp. 3–82.
  • [CT22] Benjamin Call and Daniel J. Thompson, Equilibrium states for self-products of flows and the mixing properties of rank 1 geodesic flows, J. Lond. Math. Soc. (2) 105 (2022), no. 2, 794–824.
  • [CX08] Jianguo Cao and Frederico Xavier, A closing lemma for flat strips in compact surfaces of non-positive curvature, preprint.
  • [Ebe01] Patrick Eberlein, Geodesic flows in manifolds of nonpositive curvature, Proceedings of Symposia in Pure Mathematics, vol. 69, Providence, RI; American Mathematical Society; 1998, 2001, pp. 525–572.
  • [EH90] J-H Eschenburg and Ernst Heintze, Comparison theory for riccati equations, Manuscripta mathematica 68 (1990), no. 1, 209–214.
  • [Fra77] Ernesto Franco, Flows with unique equilibrium states, American Journal of Mathematics 99 (1977), no. 3, 486–514.
  • [GN99] Marlies Gerber and Viorel Niţică, Hölder exponents of horocycle foliations on surfaces, Ergodic Theory Dynam. Systems 19 (1999), no. 5, 1247–1254.
  • [GR19] Katrin Gelfert and Rafael O. Ruggiero, Geodesic flows modelled by expansive flows, Proc. Edinb. Math. Soc. (2) 62 (2019), no. 1, 61–95.
  • [GS14] Katrin Gelfert and Barbara Schapira, Pressures for geodesic flows of rank one manifolds, Nonlinearity 27 (2014), no. 7, 1575–1594. MR 3225873
  • [GW99] Marlies Gerber and Amie Wilkinson, Hölder regularity of horocycle foliations, Journal of Differential Geometry 52 (1999), no. 1, 41–72.
  • [Kni98] Gerhard Knieper, The uniqueness of the measure of maximal entropy for geodesic flows on rank 1 manifolds, Annals of mathematics (1998), 291–314.
  • [LLS16] François Ledrappier, Yuri Lima, and Omri Sarig, Ergodic properties of equilibrium measures for smooth three dimensional flows, Comment. Math. Helv. 91 (2016), no. 1, 65–106.
  • [LMM] Yuri Lima, Carlos Matheus, and Ian Melbourne, Polynomial decay of correlations for nonpositively curved surfaces, preprint, arxiv:2107.11805.
  • [OW98] Donald Ornstein and Benjamin Weiss, On the Bernoulli nature of systems with some hyperbolic structure, Ergodic Theory Dynam. Systems 18 (1998), no. 2, 441–456.
  • [Per88] Yuval Peres, A combinatorial application of the maximal ergodic theorem, Bull. London Math. Soc. 20 (1988), no. 3, 248–252.
  • [Pes77] Ja. B. Pesin, Characteristic Ljapunov exponents, and smooth ergodic theory, Uspehi Mat. Nauk 32 (1977), no. 4 (196), 55–112, 287.
  • [Pet83] Karl Petersen, Ergodic theory, Cambridge Studies in Advanced Mathematics, vol. 2, Cambridge University Press, Cambridge, 1983. MR 833286
  • [Rue78] David Ruelle, An inequality for the entropy of differentiable maps, Bol. Soc. Brasil. Mat. 9 (1978), no. 1, 83–87.
  • [TW21] Daniel J. Thompson and Tianyu Wang, Fluctuations of time averages around closed geodesics in non-positive curvature, Comm. Math. Phys. 385 (2021), no. 2, 1213–1243.