Pressure gaps, Geometric potentials, and nonpositively curved manifolds
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.
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 be a closed, connected, smooth -dimensional manifold, and be a () rank 1 Riemannian metric on . Let be the geodesic flow on the unit tangent bundle of the Riemannian manifold . The topological pressure of a potential is the supremum of the free energy over -invariant Borel probability measures, where is the measure-theoretic entropy. A measure that achieves this supremum is called an equilibrium state of .
For the geodesic flow , "nonuniform hyperbolicity" arises from a geometric object on , namely, the singular set (see Section 2 for the precise definition). A potential is said to have a pressure gap if , where is the pressure restricted to the -invariant set . In this paper, we prove that if decays rapidly enough near the singular set, then 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 :
-
(C1)
is a closed rank 1 manifold with nonpositive sectional curvature.
-
(C2)
is either or the flat strip case , where is a codimension 1 totally geodesic flat torus.
-
(C3)
is negatively curved outside or the flat strip .
For these types of manifolds, it is convenient to study the behavior of geodesics near or using Fermi coordinates (see Section 2.3 for more details). For , let be the signed distance from to or , and the closest point on or from . The map is a projection onto , which induces a projection .
For , we consider two components from its Fermi coordinates: , and the signed angle between and the hypersurface . We define the radial curvature at as the sectional curvature of the tangent plane spanned by where .
We are particularly interested in the following two types of manifolds:
-
•
is type 1 if has an order uniform curvature bound, i.e., vanishes uniformly to order over or the flat strip (see (4.2) for more details).
- •
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.




In this paper, in addition to conditions (C1)-(C3), we assume the following condition on the continuous potential :
-
(C4)
If is type 1, then is constant on . If is type 2, then is transversally constant on , meaning depends only on the image of the projection .
Note that when is type 2, we do not assume that is constant on .
Any potential function can be extended to via
(1.1) |
With this extension, the integral of 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 the radius neighborhood of in .
Theorem A (Pressure gap criterion).
Suppose and satisfy conditions (C1)-(C4), and there exist such that
(1.2) |
for any . Then .
In particular, any potential that is Hölder continuous with sufficiently large exponents at satisfies (1.2). We also note that it is sufficient to have a lower bound on near , since when is larger than , accumulates more pressure outside the singular set. When is locally constant, we achieve [BCFT18, Theorem B] in our setup:
Corollary (Locally constant potentials).
Suppose is type 1 or type 2, and is locally constant in a neighborhood of , then .
Remark 1.1.
We briefly discuss the improvements and differences between Theorem A and [BCFT18, Theorem B]:
-
(1)
We do not assume that is constant on when is type 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)
Another difference is that the topological entropy of our singular set is zero, allowing us to better estimate . This is the primary reason we can dispense with the locally constant assumption in [BCFT18, Theorem B]. However, when has a flat strip, the transversally constant condition in Theorem A is necessary because increasing 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 contains only one flat torus or one flat strip. However, Theorem A holds when is induced by finitely many codimension 1 totally geodesic flat tori or flat strips. Specifically, Theorem A is valid under the following assumptions:
-
(1)
, where or for a codimension 1 totally geodesic flat subtorus .
-
(2)
Without loss of generality, we assume . We only need to satisfy the assumptions of Theorem A near .
-
(3)
satisfies the curvature bounds near or as a type 1 or type 2 manifold.
We say that a potential has a phase transition at if the pressure map fails to be differentiable at . 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 decays rapidly near :
Theorem B (No phase transition).
Let and satisfy conditions (C1)-(C4), and let be a Hölder continuous potential such that
(1.3) |
Then has a unique equilibrium state for each ; thus, does not have phase transitions.
Remark 1.3.
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 satisfies the following conditions: (1) a "strong" specification property, (2) the singular set can be shadowed by nearby vectors, and (3) the potential decays rapidly enough, then exhibits a pressure gap.
Results for Geometric Potentials
The second theme of this paper is dedicated to studying the behavior of the geometric potential near . Recall that the geometric potential is defined as
where 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 at . 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 near , if there exists a neighborhood of and such that for . We say has order Ricci curvature bounds if the Ricci curvature vanishes uniformly to order over (see (6.3)).
Theorem C (Geometric Potentials).
Let be a type 1 manifold without flat strips. Suppose has an order Ricci curvature bound, then near we have
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 is Hölder continuous at .
We note that our method, inspired by [GW99], currently only establishes the Hölder continuity of the geometric potential at . Achieving global Hölder continuity of may require the Hölder continuity of the unstable Jacobian tensor , which is still unclear in higher-dimensional cases.
On the other hand, as a consequence of Theorem C, we know that is a borderline case of the pressure gap criterion given in Theorem A (i.e., ). 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 exhibits a phase transition at (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).

In Figure 1.3, each point corresponds to potentials satisfying near . The shaded region represents potentials that have a pressure gap and no phase transitions by Theorem A. The geometric potential lies at the vertex of the shaded region.
We conclude this subsection by posing an open question:
Question.
Suppose lies on the boundary of the shaded region in Figure 1.3. Is there a potential satisfying near such that has a phase transition?
Examples:
The prototype of type 1 manifolds is the surface of revolution with profile , 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 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 .
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 is locally constant near , then has a pressure gap. This work was inspired by Knieper [Kni98], where the entropy gap was established as a consequence of the uniqueness of the measure of maximal entropy. Recall that the topological entropy 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 for . They showed that the Liouville measure is the only equilibrium state not supported on . However, the Hölder continuity of is even less known. It is only assured by Gerber and Wilkinson [GW99, Theorem I] for type 2 surfaces. Much less is known about 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 be a closed nonpositively curved manifold, and the geodesic flow on the unit tangent bundle . The tangent bundle contains three -invariant continuous bundles , and . The bundle is one-dimensional along the flow direction, and the other two bundles , which are orthogonal to with respect to the Sasaki metric, can be described using Jacobi fields. If is negatively curved, these three bundles form a splitting of .
A Jacobi field along a geodesic is a vector field along satisfying the Jacobi equation
where is a Riemannian curvature tensor. A Jacobi field is orthogonal if there exists such that and are perpendicular to . It is well known that when this orthogonal property holds at some , then it holds for all . A Jacobi field is parallel if for all .
Denoting the space of orthogonal Jacobi fields along by , we can identify with as follows. Consider a vector . Using the Levi-Civita connection the tangent space at may be identified with the direct sum of horizontal and vertical subspace, respectively, each isomorphic to equipped with the norm induced from the Riemannian metric on . The Sasaki metric on is defined by declaring and to be orthogonal. Restricted to , the tangent space corresponds to under this identification. Then any vector for may be written as and can be identified with an orthogonal Jacobi field along with the initial conditions . Moreover, the Sasaki norm of satisfies
The stable subspace at is then defined as
Similarly, the unstable subspace consists of vectors where is bounded for . Notice that the subbundles and are integrable to the respective foliations and . The footprints of and on are called the unstable and stable horospheres, which are denoted by and , respectively.
The rank of a vector is the dimension of the space of parallel Jacobi fields along , which coincides with the number . 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
The singular set is closed and -invariant, and in the case where is a surface the singular set can be characterized as the set of vectors where the Gaussian curvature vanishes for all . The complement of the singular set is the regular set
The geodesic flow restricted to the regular set is hyperbolic, but the degree of hyperbolicity, which can be measured by the function
where is the minimum eigenvalue of the shape operator on the unstable horosphere at . Using we can define nested compact subsets of where
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 can be found in [BCFT18].
The geometric potential is an important potential that measures the infinitesimal volume growth in the unstable direction:
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 be a hypersurface orthogonal to at where is the canonical projection. An orthogonal Jacobi field is called a -Jacobi field along , if comes from varying through unit speed geodesics orthogonal to . We denote the set of - Jacobi fields by . The shape operator on is the symmetric linear operator defined by , where is the unit normal vector field toward the same side as .
We are particularly interested in the unstable horosphere at . In this case, coincides with the space of unstable Jacobi field . For , let be the shape operator of the unstable horosphere . We know is a a positive semidefinite symmetric linear operator on , and for any unstable Jacobi field it satisfies ; see [BCFT18, Lemma 2.9].
For any vector , let is the symmetric linear map defined via for . Using the Jacobi equation, for an unstable Jacobi field we know and . Thus, we get the operator-valued Riccati equation:
(2.1) |
see [BCFT18, (7.6)]. Using the above notation, the Ricci curvature at is defined as the trace of the map .
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 in a compact metric space .
For any , we define a metric on via
and the corresponding -ball around in -metric will be denoted by . We say a subset of is -separated if for distinct . Moreover, we will identify with the orbit segment of length starting at .
Let be a continuous function on , which we often call a potential. We define to be the integral of along an orbit segment . For any subset , we let be the subset of consisting of orbit segments of length . We define
The topological pressure of on is then defined by
When is the entire orbit space , then we denote it by and call it the topological pressure of . In the case where , the resulting pressure is called the topological entropy of the flow denoted by .
Denoting by the set of all -invariant measures on , the pressure satisfies the variational principle
where is the measure-theoretic entropy of . Any invariant measure attaining the supremum is called an equilibrium state for . Likewise, any invariant measure attaining the supremum when is called a measure of maximal entropy.
2.3. Codimension 1 totally geodesic flat torus and Fermi coordinates
Let be an -dimensional closed rank 1 nonpositively curved manifold and a totally geodesic -torus in with on any . We further suppose that the complement of is negatively curved and that curvature away from a small neighborhood of 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 the universal covering of and (abusing the notation) continue denoting the lifts of and by and , respectively. Recall that the Fermi coordinate of is given by where is an -dimensional coordinate on and measures the signed distance on to .
For near , by we mean the -coordinate of . For any with near , we define and denote by the signed angle between and the hypersurface ; we adopt the convention that when . We also define
(2.2) |
When there is no confusion, we may write and for and , respectively.
Remark 2.1.
With respect to Fermi coordinates , the curve is always a geodesic perpendicular to , while with for some is not a geodesic unless and is linear.
For small, the Riemannian metric near can be written as
(2.3) |
where is the Riemannian metric on . In particular, is the Euclidean metric on .
Denoting by the vertical vector field, the second fundamental form on is defined via
for any . The shape operator is defined via
As II is bilinear and symmetric, the shape operator is diagonalizable. Its eigenvalues
are called principal curvatures at .
For any geodesic near , by the first variation formula, we have
(2.4) |
where , which then gives .
We denote by
the component of that is orthogonal to . Then .
Lemma 2.2.
If is a geodesic on near , we have
Proof.
3. Preparation and outline for pressure gaps results
3.1. Shadowing map
To distinguish vectors in and generic vectors in , we will use different fonts to denote them; more precisely, we will write and . Given an orbit segment with , we now describe a method for constructing a new orbit segment that shadows . Though simple, this construction will be crucial in proving Theorem A and Theorem D.
For any , suppose for some . For any and any such that the Fermi coordinates are well-defined for , there exists such that the distance on between and is equal to ; see Figure 3.1. From the triangle inequality we know that .
Denoting by the geodesic connecting these two points, we define
(3.1) |
Throughout the paper, we will often write , or simply , to denote whenever the context is clear.

The next few lemmas establish a few properties on the map .
Lemma 3.1.
For any and , the following statements hold:
-
(1)
For any , the function is convex for any .
-
(2)
There exists such that for any ,
for all .
-
(3)
For any , there exists such that .
Proof.
For (1), it is not hard to see from the construction of that and for all . The statement then is a consequence of from (2.4).
For (2), it is clear that
for all . We then observe that where is the geodesic with the same initial point as that is forward asymptotic to . The statement then follows as the function
vanishes when and varies continuously in . By repeating the same argument with the unstable manifolds for , we can find with the desired property.
For (3), any unit vector satisfies belongs to . Since is exhausted by compact subsets and the flat torus in is compact, the statement follows. ∎
Remark 3.2.
From here on, we will assume that belongs to with defined as in above lemmas. In particular, we will often evoke Lemma 3.1 to use the inequality
for any and .
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 is a convex function on for any , and hence there exists a well-defined number such that
(3.2) |
and that is the smallest among all such numbers. In the case where is strictly convex, there is a unique which attains the minimum of . Note that and for .
The goal of the next few sections is to establish bounds on the distance and the angle under the assumption that the radial curvature vanishes to the order of at and that the curvature near is controlled; see Section 4 and 5 for the precise description of the setting. In particular, we will show that for suitable , there exists independent of such that the shadowing vector for any and any satisfies
(3.3) |
for any . From its derivation in Proposition 4.2 and 5.3, it will be clear that the analogous inequality holds for by simply applying the symmetric argument starting from instead of :
(3.4) |
In the setting considered in Section 4 where a uniform control on the curvature is assumed on the entire , the angle 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 is a surface (see Section 5 for details) fits into the assumption of described in the above paragraph, and 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 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 to have the pressure gap. From direct integration using the estimates (3.3) and (3.4) we immediately get
Proposition 3.3 (Key inequality).
Proof.
We firstly prove the inequality for . For type 1 manifold is easier as on Sing. By (1.2), (3.3) and (3.4), we have
as
For Type 2 manifolds, use (1.2), (3.3)
and (3.4) again we have
For general , we begin by intersecting the geodesic with the hyperplane so that there are exactly two intersection points, each near and ; see Figure 3.2. We denote by the orbit segment connecting such intersection points. Since two orbit segments and differ only at either ends by length at most , there exists a constant depending only on and such that

Notice from its construction that is equal to for some near , and hence the integral admits a uniform lower bound independent of of the above proposition. Therefore, the same is true for for . ∎
4. Estimates for type 1 manifolds
Recall that is the vertical vector field. For any that is not collinear with we define the radial curvature of by
(4.1) |
that is, the sectional curvature of the plane
In this section, we will consider the first of the two settings in which vanishes uniformly to the order . Namely, if are the Fermi coordinates along , there exists such that
(4.2) |
for any with .
As outlined in Subsection 3.2, the main goal of this section is to prove that under the above assumption on , the shadowing vector for any satisfies the estimates on and claimed in (3.3).
Indeed, in this subsection, we will derive estimates on for generic vectors near ; namely, bouncing, asymptotic, and crossing vectors (see Definition 4.1). In Section 6, we will discuss behaviors of geometric potentials with respect to bouncing, asymptotic, and crossing vectors. Notice that by Definition 4.1, shadowing vectors are bouncing vectors.
Let be a neighborhood of , and in the Fermi coordinates.
Definition 4.1.
Suppose such that . Let and . We say that (relative to ) is
-
(1)
bouncing if and for ,
-
(2)
asymptotic if or ,
-
(3)
crossing if and for .
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 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 with the minimal of (when ) occurring at where, recalling from (3.2), is the the time (unique in this case) in which attains its minimum.

For type 1 manifolds, the condition (3.3) is established in the proposition below (as for bouncing vectors):
Proposition 4.2.
For any sufficiently small (see Lemma 4.5 for the domain can take), there exists independent of such that for any with and ,
-
(1)
if is a bouncing vector relative to , then for we have
-
(2)
if is an asymptotic vector relative to , then for we have
-
(3)
if is a crossing vector relative to , then for , we have
Remark 4.3.
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 such that for any , and any solution of the following Riccati equation
(4.3) |
we have
Proof.
Let be the solution of
Then satisfies (4.3). Since the solution of any first-order ODE is unique; we have .
Now we estimate . Since , we have
establishing the required upper bound on .
For the lower bound, let . Then for any the upper bound for gives
which then gives as required. ∎
Recalling the notations and from (2.2), the following lemma uses the curvature bound (4.2) to compare with for any shadowing vector .
Lemma 4.5.
There exists such that for any and any with , we have
as long as . In particular, is strictly convex and positive.
Proof.
Since we have
where is the symmetric linear map such that for . Using from the previous lemma, we claim that we can take
For and , let be the solution of
By the main theorem in [EH90], the solutions satisfy
on . By Lemmas 2.2 and 4.4 and Remark 3.2, we have
and
If vanishes somewhere in , and assume is the smallest zero of in . It is clear that because if , then it would imply which is impossible. Thus , and denote by the next zero of . Then is a geodesic connecting two distinct points on a totally geodesic submanifold , which would imply , again resulting in a contradiction. Therefore, is positive for all . ∎
We also need the following auxiliary lemma.
Lemma 4.6.
Let be a piecewise smooth, strictly decreasing function with finitely many discontinuities. Assume that and that there exists with such that
when is smooth at . Then exists a constant independent of such that
for all .
Proof.
We first consider the lower bound of . Firstly we have
(4.4) |
whenever is smooth. Thus,
Hence,
Now, we compute the upper bound. Similar to the lower bound, we get
whenever is smooth.
We then define an auxiliary piecewise smooth function
which is strictly increasing on with and . Moreover,
(4.5) |
Let be the solution of the ODE
(4.6) |
Since , from (4.5) and (4.6) we know that , thus when is slightly larger than . In fact, we have for all . This is because if for some , then we can define to be the smallest with . Since on , the condition implies . On the other hand, the condition considered with (4.5) and (4.6) implies , deriving a contradiction. Thus on .
By (4.6), we have
Thus, for any ,
(4.7) |
where
It is clear that is convex on , thus for . Hence, for any ,
(4.8) |
On the other hand, since is convex and increasing, is also increasing on . Thus, for any
(4.9) |
where is the beta function. By combining (4.7), (4.8), and (4.9), we get
Hence
Setting , we have
This completes the proof. ∎
We are ready to prove Proposition 4.2.
Proof of Proposition 4.2.
We will use to denote for simplicity. We will also use 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: is a bouncing vector. We will first prove the lower bound for when . Noting that , by Lemma 4.5, we have
and
By taking in Lemma 4.6, we know that there exists independent of such that for any ,
For , recall that . Thus . By Lemma 4.5 we have
Taking the integral on , we get
Since , we have the same bounds for .
Case 2: is an asymptotic vector. It is not hard to see that the above argument is valid when
Case 3: is a crossing vector. Denote by . By Lemma 4.5, we know that whenever ,
Thus
Taking integral, we get
(4.10) |
Hence
Since , we have . Thus there exists independent of such that
Setting in Lemma 4.6, we have
(4.11) |
Since and , by compactness, has a uniform lower bound depending on . Thus, we have
Similar to Case 1, the bound of comes from and (4.11). ∎
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 is a complete nonpositively curved surface, and is a closed geodesic of some length on which the Gaussian curvature vanishes to order . Namely, if are the Fermi coordinates along , there exists and an interval for some such that
(5.1) |
for all and for all and
(5.2) |
for all and . To simplify the argument, whenever applicable, we will adopt the notation for the Riemannian metric 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 is weakened: the neighborhood of only a small subset of is assumed to satisfy the uniform curvature bound as in (4.2).
On the complement of in 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 and for . Furthermore, unlike in the previous section where from the definition (3.1) of the shadowing map had to be carefully chosen, this setting is less sensitive to the choice of .
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 .
The goal of this section is to show that under this different set of assumptions, the shadowing vector satisfies the estimates in and as claimed in (3.3). Recalling that is the length of , we state it as a proposition below, which is the analog of Proposition 4.2.
Proposition 5.3.
There exists independent of such that for any shadowing vector and we have
and
and for any ,
Remark 5.4.
We do not expect the lower bound of to hold for all since we have little control of the metric near . For instance, when the metric near is isometric to the surface of revolution of , by Proposition 3.1 we know that is of the same scale as when is near .
To prove the proposition, we need to exploit the assumptions on and establish a few auxiliary lemmas. Consider any for some and . Recall that is the smallest number in which attains the minimum. We can decompose into subintervals and for so that
and
see Figure 5.1. Notice that , thus is not empty (though it may be arbitrarily short), but may be empty.

Since the angle satisfies for any from Lemma 3.1, there exists so that
where the maximum is taken over all possible and the minimum is taken over in if (i.e. is nonempty) or else (i.e. and is empty) in in order to exclude which could be arbitrarily small.
The following lemma shows admits a similar bound as in Lemma 4.5 when belongs to for some .
Lemma 5.5.
There exists independent of such that
for all . Moreover, there exists such that whenever for some , we also have the lower bound
Proof.
The following lemma we let . For simplicity, in the remaining part of this section, we abbreviate and where for any and .
Lemma 5.6.
For any with (namely, those not containing 0 or ), we have
-
(1)
-
(2)
for any .
Proof.
For (1) consider any and . Since is convex and decreasing, we have and . In particular,
Since , we can divide into subintervals of the same length. The integral on each subinterval, whose length is at most , is no more than that on . Hence
For (2), by (1), we have
∎
Lemma 5.7.
There exists such that for any and ,
Proof.
Similarly, if , we have
We finish the proof by taking . ∎
Proof of Proposition 5.3.
As did in Proposition 4.2, we will use to denote a generic constant. The desired lower bound for can be established just as done in Proposition 4.2. This is because the upper bound from Lemma 5.5 holds for all , and this is the only ingredient needed for the lower bound on in Proposition 4.2. In particular, we have for all .
On the other hand, the desired upper bound for is more difficult to obtain. The reason for introducing and establishing Lemma 5.7 was to obtain the upper bound. We will prove the case where and . Other cases are similar, and we will comment on them at the end of the proof. We let be the lower bound on from the above paragraph.
First, define a sequence
Then we have . Define a function via
Then is a piecewise smooth function with discontinuities at each . Moreover, Lemma 5.7 shows that is strictly decreasing function satisfying the assumption of Lemma 4.6 with . Thus by Lemma 4.6, there exists such that
for all . In particular, this inequality provides an upper bound for for for ; here should be replaced by when .
For , we have from the choice of . Thus for ,
For the last remaining subset of the domain when , which is due to the assumption that , we have . Therefore,
In sum, we can find such that
for all .
For , using (2.4) and Lemmas 3.1 (1) and 5.5, there exists such that
Thus we get the required upper bound for :
This completes the proof when and .
Other remaining cases can be dealt with similarly. When and , then exactly the same proof works; in fact, there is no need to separately consider like we did above. In the case where , we can use proceed just as we did above by bounding above by for .
Now, we consider the lower bound of . For any , the interval contains at least one . Let (resp. ) be the minimal (resp. maximal) with . We firstly compare the integrals of on and . Since ,
(5.3) |
Moreover, we have
Hence
Together with (5.3), since is non-increasing, we get
(5.4) |
Notice that , and for any . Thus, by (5.4),
∎
6. Geometric potentials
This section aims to prove Theorem C. Let be a closed rank 1 nonpositively curved manifold, and denote the geodesic flow on Recall that the geodesic potential is defined via
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:
where and is a unstable Jacobi field along such that . We also have where is the shape operator of the unstable horoshpere .
Let be the canonical projection. Its derivative sends onto We have , and thus
Thus
(6.1) |
For any , since is symmetric, we can take an orthonormal basis of so that with . Since for , for any , we have an orthonormal basis of , where is determined by
Thus and the matrix of with respect to these two orthonormal basis is . Hence
(6.2) |
Now we use the following Jacobi formula for :
For simplicity, denote by and . By (6.1) and (6.2), we have
Since , is positive semidefinite, and if are positive semidefinite, we have
When is sufficiently close to Sing, and are small nonnegative numbers, thus we have near . We summarize the above discussion below:
Proposition 6.1.
Suppose is a closed rank 1 nonpositively curved manifold. Then we have
In particular, we have near .
6.1. The proof of Theorem C
The strategy of the proof is to study the auxiliary function through the associated Riccati equation. We establish a version of Theorem C for . 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 and Ricci curvature are comparable. Since this observation is not in the mainstream of the current paper, we leave the proof in Appendix A.
Let be a type 1 manifold with order Ricci curvature bounds, that is, there exists such that
(6.3) |
for all with .
Proposition 6.2.
In particular, we have the same scale estimation for , and Theorem C follows. To prove Proposition 6.2, we need the the following lemma.
Lemma 6.3.
Assume there exist so that
-
(1)
for all , then there exists depending on so that
-
(2)
for all , and , then there exist depending on so that
for all
Proof.
Denote by . Since is diagonalizable and all eigenvalues are nonnegative, by Cauchy-Schwartz we have
Thus, by the Riccati equation,
On the other hand, denote by We have
-
(1)
Compare with the solution of
By the main theorem in [EH90], we have
Compare with the solution of
We have
-
(2)
Compare with the solution of
We have
Since is decreasing, so does . For , we get
Compare with the solution of
For ,
∎


Proof of Proposition 6.2.
We follow the main steps in the proof of [GW99, Lemma 3.3].
We use instead of for simplicity.
Case 1: is bouncing or asymptotic. See Figure 6.1a.
Since the asymptotic case is the bouncing case with ,
we only have to consider the bouncing . We may assume .
Denote by
For any , we have . Moreover we have
Lemma 6.4.
for some independent of .
Proof.
Case 2: , similar to Case 1.
Case 3: , and first decreases, then increases on . Assume satisfies . By Case 1 we know that . ∎
Take in Lemma 6.3(1), we get
By (3.6), we know that
Thus, we finish the proof of Proposition 6.2
in this case.
Case 2: is crossing. See Figure 6.1b. Recall
that , and . Denote by .
Since is close to , we may assume that
and . Since , by (4.10)
we have
Thus, it suffices to prove
(6.4) |
We prove (6.4) in the following two cases:
-
•
Case 2a: . In this case, and for . Let be the minimal solutions of
-
•
Case 2b: . In this case, we have . Since crosses and we do not have flat strips, by the result of Case 2a, we know that
∎
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
The flat strip is the part with , and the metric satisfies both curvature conditions by Lemma A.1. Let be a unit vector with and angle , meaning that is a vector exiting the flat strip. At time , the geodesic enters the strip with vector . By symmetry, the angle of is . Assume that the Hölder continuity in Proposition 6.2 holds for both and , namely, . As is in the flat strip for , satisfies the Riccati equation , and the solution is
Plug in and , we have
Contradictory to .
The Hölder continuity of 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 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, and are Hölder continuous in a small neighborhood of .
7. Sufficient criteria for the pressure gap
Let be a closed Riemannian manifold and the geodesic flow on . In this section, we will describe an abstract result to establish the pressure gap for a given potential . 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 satisfies the specification at scale if there exists such that given finite orbit segments and with for all , there is such that for all .
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 ’s such that ; that is, the transition time is exactly equal to .
7.2. Abstract result on the pressure gap
We now list the conditions together which establish the pressure gap. Let be a closed rank 1 manifold with a codimension 1 flat subtorus, and 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
to be the set of orbit segments with endpoints in , we require that the geodesic flow and the potential satisfy the following property:
-
(1)
Singular set zero entropy property: .
-
(2)
Specification property: For any and , the orbit segments satisfies the specification at scale .
-
(3)
Shadowing property: There exists such that for every there exists a map
with the following properties: denoting by the shadowing vector of an arbitrary ,
-
(a)
The conclusion of Lemma 3.1 hold for .
-
(b)
For any , there exists such that for any , the vector satisfies
for all .
-
(a)
-
(4)
Special Bowen property: For any , there exists independent of such that
for any .
Proposition 7.2.
Suppose the geodesic flow and the potential satisfy the above listed conditions (1), (2) and (3). Then, has a pressure gap.
Proof of Theorem A.
Each condition listed above can be verified as follows. By design, we know . The specification property (2) is already established in [BCFT18] for the geodesic flow over rank 1 nonpositively curved manifold. With 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 to be where is the constant from Proposition 4.2 and 5.3. Lastly, (4) follows from Proposition 3.3. Hence, 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 be a continuous flow on a compact space , and be an invariant probability measure. Then for every potential there exists some
for all .
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 by the integration of some special in along the flow. The second step is using the specification property (2) and shadowing property (3) to create a separated set that bounds . The last step is using the special Bowen property (4) to estimate the pressure.
Notice that without lost of generality, we may assume ; otherwise, we consider . Hence, it is sufficient to show that there exists a separated set on such that
(7.1) |
The first step is to find a special singular vector for shadowing. Since on is entropy-expansive, we know there exists an equilibrium state of . That is
By the singular set zero entropy property (1), we know ; and thus, . By Peres’ Lemma (Theorem 7.3), there exists such that for all
(7.2) |
The second step is to create such a separated set on . For given in the shadowing property (3), from Lemma 3.1 we get where for any . We also obtain the constant from (3b) corresponding to .
We now set be a constant such that , and for each consider
For any small such that , consider any size subset
with . Setting and , such a subset can be viewed as a partition of the interval into subintervals, each of length for . We denote , and we know .
Given , we consider singular vectors and the corresponding regular orbits where and for . Using the specification property (2), one can find a vector which shadows orbits (see Fig. 7.1), that is, for

We claim that is a -separated set. To see this, let for , and Without loss of generality, suppose , then the claim follows the inequalities below (see Fig. 7.2 ):

The last step is using the special Bowen property (4) to bound the pressure from below. For , the integral of during each transition period is bounded below by , by the special Bowen property (4) and (7.2) we get
(7.3) | ||||
Recall that , summing the above inequality over all possible subsets gives
As is a -separated set (where with and are arbitrary), this implies that
Combining the last two inequalities with , one establishes the pressure gap for by taking sufficiently small. ∎
Remark 7.4.
We list below several main differences between this proof and the proof of [BCFT18, Theorem B].
-
(1)
Our proof utilize 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 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)
Appendix A Ricci curvature bound and radial curvature bound comparison
Recall that is a dimensional manifold with nonpositive sectional curvature, and is a totally geodesic -subtorus. We assume that for any and any 2-plane . On the universal cover , we define the Fermi coordinate near in the following way: is the coordinate on , and measures the signed distance on to . is always a geodesic perpendicular to . The Riemannian metric near is
(A.1) |
where is the Riemannian metric on . In particular, is the Euclidean metric on
If is a warped product, namely, . We have . Since is totally geodesic, we have .
Lemma A.1.
If , then the following conditions are equivalent:
-
(1)
There exists such that
-
(2)
There exists so that
-
(3)
There exists so that
Proof.
Since , the radial curvature is
while the sectional curvature is given by
where be the angle between and .
Now we compute the Ricci curvature. If is normal, then and we are done. Otherwise, denote by the perpendicular complement in , the angle between and , and the horizontal subspace at . Since both and have codimension 1, has dimension . We can construct an orthonormal basis such that , , and the section spanned by is normal. Then we have
(A.2) |
When , we get . Thus, for any , the Ricci curvature can be calculated using (A.2).
Since and , we have . Since , . Therefore, by (A.2).
Assume for some . We have and . Thus, is the dominant term in (A.2). Therefore, .
Since and , we have and , thus . ∎
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 be a continuous flow on a compact space , and be an invariant probability measure. Then for every potential there exists some
for all .
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 be a probability space and be a measure-preserving flow on . If and , then
where
Proof of Theorem B.1.
For , we define
and
It is enough to show
To see this, we first notice that . We then apply the Maximal Ergodic Theorem on and , i.e., , and get
Observe that , and which guarantees for all . Since is continuous and is compact, we know ’s are nested compact sets, and thus
∎
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 , 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.