• Aucun résultat trouvé

L’INSTITUTFOURIER DE ANNALES

N/A
N/A
Protected

Academic year: 2022

Partager "L’INSTITUTFOURIER DE ANNALES"

Copied!
55
0
0

Texte intégral

(1)

A L

E S

E D L ’IN IT ST T U

F O U R

ANNALES

DE

L’INSTITUT FOURIER

Vincent PIT & Barbara SCHAPIRA

Finiteness of Gibbs measures on noncompact manifolds with pinched negative curvature

Tome 68, no2 (2018), p. 457-510.

<http://aif.cedram.org/item?id=AIF_2018__68_2_457_0>

© Association des Annales de l’institut Fourier, 2018, Certains droits réservés.

Cet article est mis à disposition selon les termes de la licence CREATIVECOMMONS ATTRIBUTIONPAS DE MODIFICATION3.0 FRANCE. http://creativecommons.org/licenses/by-nd/3.0/fr/

L’accès aux articles de la revue « Annales de l’institut Fourier »

(http://aif.cedram.org/), implique l’accord avec les conditions générales d’utilisation (http://aif.cedram.org/legal/).

cedram

Article mis en ligne dans le cadre du

(2)

FINITENESS OF GIBBS MEASURES ON NONCOMPACT MANIFOLDS WITH PINCHED

NEGATIVE CURVATURE

by Vincent PIT & Barbara SCHAPIRA

Abstract. — We characterize the finiteness of Gibbs measures for geodesic flows on negatively curved manifolds by several criteria, analogous to those pro- posed by Sarig for symbolic dynamical systems over an infinite alphabet. These criteria should be useful in the future to find more examples with finite Gibbs measures. As an application, we recover Dal’bo–Otal–Peigné criterion of finiteness for the Bowen–Margulis measure on geometrically finite hyperbolic manifolds, as well as Peigné’s examples of geometrically infinite manifolds having a finite Bowen–

Margulis measure.

Résumé. — Nous donnons plusieurs critères caractérisant la finitude des me- sures de Gibbs pour le flot géodésique sur les variétés à courbure négative, analogues à ceux proposés par Sarig pour les sous-décalages sur des alphabets infinis. Ces cri- tères effectifs devraient permettre de trouver davantage d’exemples de mesures de Gibbs finies. En application, nous retrouvons le critère de Dal’bo–Otal–Peigné sur la finitude de la mesure de Bowen–Margulis pour des variétés hyperboliques géomé- triquement finies, ainsi que les exemples de Peigné de variétés à courbure négative géométriquement infinies possédant une mesure de Bowen–Margulis finie.

1. Introduction

Hyperbolic dynamical systems are, heuristically, so chaotic that all be- haviours that one can imagine indeed happen for some orbits. From the point of view of ergodic theory, this can be expressed by the existence of Gibbs measures. When the space is compact, choosing any regular enough

“weight function”, called apotential, i.e. usually a Hölder continuous map on the space, we can find an invariant ergodic probability measure which gives, roughly speaking, large measure to the sets where the potential is

Keywords:Gibbs measures, thermodynamic formalism, geodesic flow, geometrically in- finite manifolds, Kac lemma.

2010Mathematics Subject Classification: 37D40, 37D35, 28D20, 37A35, 37A40,

(3)

large, and small measure to those where the potential is small. This is a quantified way of saying that all behaviours that one can imagine, repre- sented by the choice of a potential, indeed happen for a hyperbolic dynam- ical system.

Such measures are calledGibbs measures, and their existence for Hölder potentials and uniformly hyperbolic flows was proved by Bowen–Ruelle [6].

Thermodynamical formalism, i.e. the study of the existence and proper- ties of these measures, has been extended to noncompact situations in two main cases. In symbolic dynamics, Sarig ([17, 18]) studied thermodynami- cal formalism of shifts over a countable alphabet. In the context of geodesic flows of noncompact negatively curved manifolds, the measure of maximal entropy, which is associated with the constant potential, now commonly called theBowen–Margulis measure, has been extensively studied, first by Sullivan on geometrically finite hyperbolic manifolds [19], and later with his ideas by many others. Among them, let us cite Otal–Peigné who ob- tained an optimal variational principle in [12], and Roblin [16] who estab- lished an equidistribution result of measures supported by periodic orbits towards the Bowen–Margulis measure. Generalizations of their results to Gibbs measures with general Hölder potentials have been proved in [14].

All these results hold if and only if the Gibbs measure associated with the potential is finite. It is therefore very important to be able to charac- terize, or at least to give sufficient conditions for the finiteness of Gibbs measures. Some partial results and examples have already been proved in the past. In [17, 18], in a symbolic context, Sarig established two finite- ness criteria for the measure associated with a Hölder potentialF. Iommi and its collaborators extended this study to suspension flows over such shifts, in [4, 11]. In the case of geodesic flows on the unit tangent bundle of noncompact manifolds, the first criterion appeared in [8]. In the par- ticular case of geometrically finite manifolds, Dal’bo–Otal–Peigné showed that the Bowen–Margulis measure is finite if and only if a series involving the parabolic elements of the fundamental group converges. This criterion has been extended later by Coudène [7] to Gibbs measures on geometrically finite manifolds. Finally, Peigné constructed in [15] the first examples of ge- ometrically infinite hyperbolic manifolds whose Bowen–Margulis measure is finite. His proof, once again, involved the convergence of a certain series.

Ancona [2] also obtained such examples, but through harmonic analysis.

Our main motivation is to give a unified way to check whether a Gibbs measure is finite, not specific to a certain class of manifolds, and allowing the recovery of all results mentioned above, the geometric ones as well as

(4)

Sarig’ symbolic criteria. We will provide three equivalent criteria for the finiteness of Gibbs measures, all involving the convergence of some series, two of them being the geometric analogues of Sarig’s criteria in terms of lengths of periodic orbits, the other one being a reformulation in terms of the action of the fundamental group Γ ofM on its universal cover, which is more convenient in our geometric context.

To this end, we will not rely on having a symbolic coding for the geodesic flow, which cannot be ensured in the general case, but on geometrical esti- mates andKac’s Lemma. Recall that this lemma asserts that if a measureµ is conservative andAis a measurable set with positive finite measure, then the measure of the whole space equals P

n>1nµ(An), where An denotes the subset of points ofA that return inAafter exactlyniterations of the dynamics. Therefore, the finiteness of µ is equivalent to the convergence of a certain series. However, Kac’s Lemma is in general more an abstract result than an useful criterion, due to the difficulty of estimating the mea- sure of the setsAn, but the geometry of negatively curved manifolds will allow us to convert it into an explicit efficient criterion.

Let us first give some notations in order to state our results. We are interested in the geodesic flow (gt) on the unit tangent bundleT1M of a complete manifold with pinched negative curvature. Our results also hold whenM is a negatively curved orbifold, that is the quotient of a complete simply connected negatively curved manifold by a discrete nonelementary group Γ which can contain torsion elements. We study this flow in restric- tion to itsnonwandering setΩ. We denote byP the set of periodic orbits, byPW the set of periodic orbits which intersect some setW ⊂T1M, and byPW0 the subset of primitive periodic orbits. For a given periodic orbit p∈ PW, we denote byl(p) its length, and bynW(p) the “number of times that the geodesicpcrossesW” (see Section 3 for a more precise definition).

We consider a potentialF :T1M →R, i.e. a Hölder continuous map, and denote byFe:T1Mf→Rits lift and P(F) its pressure. It is shown in [14]

how one can build a Gibbs measuremF associated with the potentialF.

The Hopf–Tsuji–Sullivan Theorem (see [14, Theorem 5.4]) asserts that a Gibbs measuremF is either ergodic and conservative (possibly finite or infinite) or totally dissipative, depending on whether thePoincaré seriesof (Γ, F) is respectively divergent or convergent. Therefore, before investigat- ing the finiteness of a Gibbs measure, we investigate when it is ergodic and conservative.

Definition 1.1 (Recurrence). — A potentialF :T1M →Ris said to be recurrent when there exists an open relatively compact subset W of

(5)

T1M, which intersects the nonwandering setΩ, such that

X

p∈P

nW(p)e R

p(F−P(F))

= +∞.

By analogy with the recurrence property for potentials on infinite sub- shifts developed in [18], where a periodic orbit may be made of several periodic points beginning with the same letter, this integernW(p) can be interpreted as how many changes of origins are possible along the geodesic orbitpso that the parametrization starts inW.

Our first result is a reformulation of the divergence of the Poincaré series in terms of periodic orbits.

Theorem 1.2 (Ergodicity criterion). — LetM be a negatively curved orbifold with pinched negative curvature, and F : T1M → R a Hölder continuous potential with P(F) < +∞. Then the Gibbs measure mF is ergodic and conservative if and only ifF is recurrent.

Unfortunately, this equivalence is unlikely to be very useful in practice.

The main interest of Theorem 1.2 is to enlighten the very strong analogy between our results on geodesic flows on noncompact manifolds and Sarig’s work in symbolic dynamics over a countable alphabet, despite the fact that no general coding result of the geodesic flow by a symbolic dynamical system is known in this context.

Definition 1.3 (Positive recurrence for the geodesic flow). — Let M be a negatively curved orbifold with pinched negative curvature. A Hölder continuous potential F : T1M → R with P(F) < +∞ is said positive recurrent relatively to a set W ⊂ T1M intersecting Ω and the integer N>1 if it is recurrent and

X

p∈P0W nW(p)6N

l(p)e R

p(F−P(F))

<+∞.

Our main finiteness criterion is the following result, which is the geomet- ric analogue of the symbolic criterion of Sarig [18].

(6)

Theorem 1.4 (First finiteness criterion). — Let M be a negatively curved orbifold with pinched negative curvature, and F : T1M → R a Hölder continuous potential withP(F)<+∞. Denote bymF its associ- ated Gibbs measure.

(1) IfF is recurrent, and there exists an open relatively compact set WMmeetingπ(Ω)such thatF is positive recurrent with respect toW=T1W and someN >KW (whereKW depends only on the geometry ofW andM), thenmF is finite.

(2) If mF is finite, then F is recurrent, and positive recurrent with respect to W = T1W and any N > 1 for every open relatively compact setWM meetingπ(Ω).

We shall see in the proof that when M is a manifold, andW is a small open ball, the constantKW equals one.

In particular, when F is recurrent, then F is positive recurrent with respect to W =T1W for some open relatively compact setW which in- tersectsπ(Ω) and for someN >KW if and only if it is positive recurrent relatively to any such set.

The proof of Theorem 1.4 follows from Theorem 1.6 below, which ex- presses the finiteness ofmF in terms of the action of Γ on Mfinstead of periodic orbits on T1M. This criterion does not appear in Sarig’s work because it has no meaning in a purely symbolic setting. However, it is very useful in our geometrical context.

We start by introducing the following notation. Given a subsetWf⊂Mf, denote by

Γ

We = (

γ∈Γ

y, y0 ∈fW ,[y;γy0]∩gfW 6=∅ ⇒Wf∩gfW 6=∅ orγfWgfW 6=∅

)

the set of elementsγsuch that there exists a geodesic starting fromWfand finishing inγfW that meets the orbit ΓfW only at the beginning or at the end.

Definition 1.5(Positive recurrence in the universal cover). — The pair (Γ,Fe)is said to be positive recurrent with respect to a setWMfsuch thatW=T1W intersectsΩifF is recurrent and

xM ,f X

γ∈Γ

We

d(x, γx)e Rγx

x (F−Pe (F))<+∞.

(7)

Theorem 1.6 (Second finiteness criterion). — Let M be a negatively curved orbifold with pinched negative sectional curvature. LetF :T1M → Rbe a Hölder continuous potential withP(F)<+∞, and denote bymF the associated Gibbs measure onT1M.

(1) IfF is recurrent, and if(Γ,Fe)is positive recurrent with respect to some open relatively compact setWf⊂Mfmeeting π(eΩ), then mF

is finite.

(2) IfmF is finite, thenF is recurrent, and(Γ,Fe)is positive recurrent with respect to any open relatively compact setWf ⊂Mfmeeting π(eΩ).

In particular, whenF is recurrent, then (Γ,Fe) is positive recurrent rela- tively to some open relatively compact setWfwhich intersectsπ(eΩ) if and only if it is positive recurrent relatively to any such set.

In [17], Sarig proved a finiteness criterion, established earlier than the symbolic analogue of Theorem 1.4. This criterion seems less practical than his later work. However, we wanted to show a complete analogy between the symbolic and our geometric settings, so we established the same crite- rion in our situation. The proof is different from the previous criteria and relies on equidistribution of weighted periodic orbits. This criterion requires the assumption that the geodesic flow is topologically mixing on Ω. This extremely classical assumption is satisfied in most interesting situations, even if its validity is open in general.

Definition 1.7. — The potentialF is said to be positive recurrent in the first sense of Sarig [17]if there exists an open relatively compact subset W ofT1M meeting Ω, and constants c >0,t0>0andC >0such that

t>t0, 1

C 6 X

p∈P t−c<l(p)6t

nW(p)e R

p(F−P(F))

6C .

Theorem 1.8 (Third finiteness criterion). — Let M be a negatively curved complete orbifold, with pinched negative curvature. Assume that its geodesic flow is topologically mixing. Let F : T1M →R be a Hölder continuous potential with P(F)<+∞, and denote by mF its associated Gibbs measure on T1M. Then mF is finite if and only if F is positive recurrent in the first sense of Sarig with respect to some open relatively compact setW meetingΩ.

When this theorem holds, F is actually positive recurrent in the first sense of Sarig with respect to any open relatively compact setWmeeting Ω.

(8)

The structure of this paper goes at follows. Section 2 introduces the background on geometry and thermodynamic formalism, including some elementary lemmas of hyperbolic geometry stated in a convenient way for our purposes. In particular, Lemma 2.4 plays a crucial role in the proof.

Section 3 introduces the notion of number of returns of a periodic orbit, which is used to state Theorem 1.2, 1.4 and 1.8. The proof of Theorem 1.2 is given in Section 4. Theorem 1.6 is proved in Section 5, from which an intermediate Theorem 6.1 is derived in Section 6, and Theorem 1.4 is itself derived in Section 7. In Section 8, we state and prove a couple of equidistri- bution results for nonprimitive periodic orbits, and derive from it the proof of Theorem 1.8. Finally, we show in Section 9 how to retrieve previous finiteness results from ours.

We believe that these criteria will lead to new examples of interesting manifolds with finite Gibbs measure. This will be done in the future.

2. Preliminaries

2.1. Geodesic flow in negative curvature

In the following,Mfis a Hadamard manifold with pinched negative sec- tional curvature −b2 6 k 6−a2 <0, Γ is a discrete group of isometries preserving orientation of Mf, M = Γ\Mf is the quotient orbifold (mani- fold whenever Γ has no torsion elements), andT1M = Γ\T1Mfis its unit tangent bundle. Observe once and for all that Riemannian/differential con- cepts are still well defined onM or T1M by defining objects first on the universal cover, and then going down toM. With a slight abuse of notation, we denote by π:T1MM or π :T1Mf→Mfthe canonical projection, and byPΓ:T1Mf→T1M or PΓ :Mf→M the quotient maps.

The geodesic flow ofT1Mfand ofT1M is denoted by (gt)t∈R. The bound- ary at infinityMfis the set of equivalence classes of geodesic rays staying at bounded distance from each other. IfvMf, we denote byv± the posi- tive and negative endpoints of the geodesic it defines.

The limit set Λ(Γ) is the set of accumulation points in Mfof the Γ- orbit of any point inMf. We will only consider the nontrivial case where Γ is nonelementary, that is Λ(Γ) is infinite. Eberlein [9] proved that the nonwandering set Ω of the geodesic flow coincides with the set of vectors vT1Mfsuch thatv±∈Λ(Γ).

TheHopf coordinates relatively to any base pointx0Mfare given by vT1Mf7→(v, v+, βv+(x0, π(v)).

(9)

where the so-called Busemann cocycle βv+(x0, v) is the limit of d(x0, π(gtv))d(π(v), π(gtv)) when t →+∞. They induce an extremely useful homeomorphism betweenT1Mfand (∂MMf)\diagonal×R. The geodesic flow acts by translation on the real coordinate, so that all dynamically relevant sets can be expressed nicely in terms of these coordi- nates.

The set of periodic orbits (respectively primitive periodic orbits) of the geodesic flow is denoted byP (respectively P0). Recall that each periodic orbit of the geodesic flow corresponds to exactly one conjugacy class of hyperbolic elements of Γ. More precisely, let Γh be the set of hyperbolic (or loxodromic) isometries of Γ, and Γ0h those which are primitive. By definition, such a γ has two fixed points in Mf, one repulsive and the other attractive. It acts by translation on the geodesic line of Mfjoining them so that the geodesic orbit ofT1Mffrom the repulsive to the attractive endpoint induces on T1M a periodic orbit of the geodesic flow. We will denote byl(γ) forγ∈Γh, or equivalentlyl(p) forp∈ P, the period of this orbit.

2.2. Thermodynamical formalism

In this section, we recall briefly some facts about thermodynamical for- malism on negatively curved manifolds, which are either classical or can be found in [14].

Let F : T1M →R be a Hölder continuous map (or potential), and Fe be its Γ-invariant lift to T1Mf. Its topological pressure is defined as the supremum

P(F) = sup

µ∈M

h(µ) +

Z F

,

where the supremum is taken over the setM of all invariant probability measures, andh(µ) is the Kolmogorov–Sinai entropy of µ.

A dynamical ballB(v, T, ε) is the set B(v, T, ε) =PΓ

n

wT1Mf

d(π(gtv)), π(gt(w)))6εfor all 06t6To . Here, there is a slight abuse of notation : we use the distancedonMfinstead of a distance on T1Mf, for example the Sasaki metric. However, standard results about geodesic flows in negative curvature show that these two points of views are equivalent. We refer to [14] for details. An invariant Radon measureµsatisfies theGibbs property (see [14, Section 3.8]) if for

(10)

all compact subsetsKT1M, there exists a constant CK >0, such that for allvK andT >0 such thatgT(v)∈K, one has

(2.1) 1 CK

e RT

0 F(gt(v)) dt−T P(F)

6µ(B(v, T, ε))6CKe RT

0 F(gt(v)) dt−T P(F)

. The careful reader will observe that this definition is slightly simplified compared to [14], but describes the same measures.

When P(F) is finite, the Patterson–Sullivan–Gibbs construction, de- tailed in [14], allows to build a measure mF, which satisfies the above Gibbs property, whose lift ˜mF onT1Mfhas the following nice expression in the Hopf coordinates

(2.2) d ˜mF(v) = 1

DF−P(F),x0(v, v+)2F◦ιx

0 (v) dµFx

0(v+) dt , whereιis the flip mapι:v→ −v,µFx0 is the so-calledPatterson–Sullivan–

Gibbs conformal densityon the boundary, andDF is theF-gap map from xdefined as

DF,x(ξ, η) = exp1 2

t→+∞lim Z ηt

x

Fe− Z ηt

ξt

Fe+ Z x

ξt

Fe

.

This map is continuous and positive onMMf\ {diagonal}, and therefore bounded away from 0 and +∞ on all compact sets of M

Mf\ {diagonal}. Moreover, the pointx0 being arbitrary, the Patterson–

Sullivan–Gibbs conformal densities (µFx)

x∈Me

form a family of measures that have full support in Λ(Γ) and are Γ-quasi-invariant.

The Gibbs measure mF satisfies the following alternative, known as the Hopf–Tsuji–Sullivan Theorem, proved by Roblin [16] in full generality when F≡0, and whose proof has been adapted to Gibbs measures in [14]. First, recall that the pressureP(F) is also the critical exponent of the following Poincaré series

PΓ,x,F(s) =X

γ∈Γ

e Rγx

x (F−s)e .

This Hopf–Tsuji–Sullivan–Roblin theorem for Gibbs measures illuminates how the convergence or divergence of the above series for s =P(F) is a crucial point for the ergodicity of the Gibbs measuremF.

Theorem. — LetM be a negatively curved orbifold with pinched neg- ative curvature, andF : T1M → R a Hölder continuous map with finite pressure. Then the measuremF is ergodic and conservative if and only if

(11)

the Poincaré seriesPΓ,x,F(s)diverges ats=P(F), i.e.

X

γ∈Γ

e Rγx0

x0

(F−P(F))e = +∞,

and the measuremF is totally dissipative otherwise.

In fact, one can show that “the” measuremF built in [14] is well defined if and only if the above series diverges. As this is the only interesting case for us, we do not care about this problem of terminology.

When a Gibbs measure mF is finite, it is automatically ergodic and conservative. However, of course, the converse is not true, and it is precisely the purpose of this paper to propose criteria of finiteness.

2.3. Some exercises in hyperbolic geometry

This section gathers some well-known lemmas about the geometry of manifolds with pinched negative sectional curvature.

We start by recalling a very classical comparison lemma, as stated for example in [13, Lemma 2.1], from which we will derive the next lemmas.

Lemma 2.1. — Let(X, d)be a CAT(−1)-space. For all pointsx, yinX andz in XX, and every t ∈[0;d(x, z)](finite if zX), if xt is the point on[x;z]at distance tfromx, then

d(xt,[y;z])6e−tsinh(d(x, y)).

In particular, applying this lemma twice leads to the following lemma.

Lemma 2.2. — Letr, r0 >0. Let x, x0, y, y0Mfsuch thatd(x, y)6r and d(x0, y0) 6 r0. For every t ∈ [0;d(x, x0)], denote by xt the point on [x;x0]at distancet fromx. Then

d(xt,[y;y0])6sinh(r)e−t+ sinh(r0)et−d(x,x0)+sinh(r). Recall also the following.

Lemma 2.3([14, Remark 2 following Lemma 3.2]). — For everyR>0, there is a constant C that only depends on Fe, R and the bounds on the sectional curvature of Mf such that for every x, x0, y, y0 in Mf satisfying d(x, x0), d(y, y0)6R

Z y x

Fe− Z y0

x0

Fe

6C .

(12)

2.4. Parallel geodesic segments avoiding images of a compact set

The following geometrical lemma is the key ingredient of the proof of Theorem 1.6, from which Theorem 1.4 is derived. It asserts that, if a geo- desic segment [y;y0] is known to avoid the Γ-orbit of balls B(x, R) except maybe at its beginning or at its end, then every other geodesic segment whose endpoints are close fromy and y0 will also essentially avoid the Γ- orbit of ε-shrinked balls B(x, Rε), provided that the guiding segment [y;y0] is long enough.

Lemma 2.4 (Long range subset avoidance). — Let fW be an open rel- atively compact subset of Mf. For every R > 0, there exists ρ = ρ(R) such that for allε >0, all Wf0 ⊂fW open relatively compact subsets with d(fW0,Mf\fW) >ε, all y, y0, z, z0Mf with d(y, z), d(y0, z0) 6R, and all γ∈Γ, if

[y;y0]∩γfW =∅ and [z;z0]∩γWf06=∅. then we have

min

d(y, γfW), d(y0, γWf)

6ρ−logε .

Proof. — Denote by l=d(z, z0). Letzt be a point of [z;z0] inside γfW0, witht=d(z, zt). Without loss of generality, we can assume thatt6 2l. By Lemma 2.2, we have

d(zt,[y;y0])6sinh(R)

e−t+et−l+sinh(R)

6C e−t+et−l

62Ce−t with C = sinh(R)esinh(R). The assumption d(zt,[y;y0]) > ε ensures that t 6log(2Cε ) and henceforth thatd(y, zt)6ρ−logε with ρ = log(2C) +

diamfW.

2.5. About finding hyperbolic isometries

In the proofs of Theorems 1.2 and 1.4, we will need to compare sums indexed on periodic orbits of the geodesic flows, i.e. on conjugacy classes of hyperbolic elements of Γ, with sums indexed on the whole group Γ. To this end, we need some technical tools to go from the former to the latter.

We start by recalling a variant of Anosov closing lemma, which is easily obtained by combining [10, Corollaire 8.22] with Lemma 2.2.

(13)

Lemma 2.5. — For everyl, ε >0, there existsε0∈]0; 1]withlimε→0ε0= 0such that for every isometryγof any proper geodesic CAT(−1)-spaceX, for everyx0in X, ifd(x0, γx0)>l andd(x0,

γ−1x0;γx0

)6ε0, thenγ is hyperbolic andd(x0, Aγ)6ε, whereAγ is the translation axis ofγin X.

Let Γh be the set of hyperbolic elements of Γ. Ifγ ∈Γh, we denote by AγT1Mfitsaxis, i.e. the set of vectorsvT1Mfsuch that gl(γ)v=γv, wherel(γ) is the minimal displacement of a point byγ. In other words,Aγ

is the set of unit vectors on the geodesic joining the repulsive fixed point to the attractive fixed point, oriented towards the latter.

If xMf, l > 0 and UMf is open, then the angular sector at distancelbased atxand supported byU is the open set

Cx,l(U) =n zMf

d(z, x)> l and ∃ξU, z∈]x;ξ[o .

Lemma 2.6. — LetW ⊂f T1Mfbe an open relatively compact set inter- sectingΩ,e ε >0 and xπ(fW)∩Conv(Λ(Γ)). There existg1, . . . , gk ∈Γ and a finite setS⊂Γsuch that for everyγ∈Γ\S, there existi, jsuch that γ0=gj−1γgi is hyperbolic and its axis satisfiesAγ0∩W ∩f T1B(x, ε)6=∅.

Proof. — Take ε0 =ε0(ε,1) given by Lemma 2.5. Let U and V be two non empty open sets of Mfwith disjoint closures, both meeting Λ(Γ), such that any geodesic orbit ofT1MffromUtoV meetsW ∩Tf 1B(x, ε0). Fix two non empty open setsU0, V0Mfmeeting Λ(Γ) such that U0U and V0V. There exists l0 >1 such that for every yU0 =Cx,l0(U0) and zV0 = Cx,l0(V0), the geodesic orbit of T1Mf from y to z meets W ∩f T1B(x, ε0).

As Γ acts minimally on Λ(Γ), there existg1, . . . , gp, . . . , gk∈Γ such that Λ(Γ)⊂

p

[

i=1

giU0 and Λ(Γ)⊂

k

[

i=p+1

giV0.

LetR0= sup

d(x, g−1i x)

i= 1, . . . , k ,R1=R0+ 2ε, and define U00=n

yMf

B(y, R1)⊂U0o

and V00=n zMf

B(z, R1)⊂V0o . Denote byU00andV00their closures insideMf∪∂Mf. Observe thatU00= U0andV00=V0, so that we still have inMf∪Mf

Λ(Γ)⊂

p

[

i=1

giU00 and Λ(Γ)⊂

k

[

i=p+1

giV00.

(14)

Therefore, both sets K= Conv(Λ(Γ))\

p

[

i=1

giU00 and L= Conv(Λ(Γ))\

k

[

i=p+1

giV00

do not meet Mf and are relatively compact in Mf. Therefore, S = γ∈Γ

γ−1xK or γxL is finite.

Now if γ∈Γ\S, by construction, there existi, jsuch thatγ−1xgiU00 andγxgjV00. Ifγ0 =gj−1γgi, thenu=γ0−1g−1j xU00andv=γ0gi−1xV00, which means thatγ0xsatisfies

d(γ0x, v) =d(x, g−1i x)6R0< R1,

i.e. γ0xV0 and similarly γ0−1xU0. By our choice of U0 and V0, this ensures that d(x, γ0x) > l and h

γ0−1x;γ0xi

meets B(x, ε0), so that by Lemma 2.5,γ0 is hyperbolic and its axisAγ0 meetsT1B(x, ε).

Finally, let z∈eπ(Aγ0)∩B(x, ε). Since

d(γ0z, v)6d(γ0z, γ0x) +d(γ0x, v)6ε+R0< R1,

we deduce that γ0zV0 and likewise γ0−1zU0. This implies that the geodesic orbit ofT1Mffrom γ0−1z toγ0z, i.e.Aγ0, also meetsW.f

2.6. Shadows

Ify, y0are two distinct points ofMf∪∂Mf, letv(y, y0)∈Mf(respec- tivelyv+(y, y0)∈Mf) be the endpoint of the one-sided infinite geodesic ray going fromy0toy(respectively fromytoy0). The mapsv+ andv are continuous for the usual topology on (fMMf)2\ {diagonal}.

With the above notations, ifxMf∪MfandW is an open, relatively compact, geodesically convex subset ofMf, the shadow ofW viewed from xis the set

OxW ={v+(x, y)|yW}=n

ξMf

]x;ξ[W 6=∅o .

We start by stating a classical lemma that asserts that, if the base point is far enough from the set that casts the shadow, then it can be moved around by a bounded amount almost without changing the shadow.

Lemma 2.7. — For everyr >0, every 0< ε < r and everyδ >0, there existsl0 >0 such that for allx, yMfsatisfyingd(x, y)>l0 and for all zB(y, δ)we have

OyB(x, rε)⊂ OzB(x, r)⊂ OyB(x, r+ε).

(15)

We will also need the two following lemmas about products of shadows.

Their proofs are very similar to [14, Lemma 3.17] and therefore ommitted.

Lemma 2.8. — For all r > 0, r0 > 0 and ε > 0, there exists l0 > 0 such that for allx, x0Mfsatisfyingd(x, x0)>l0, for allyB(x, r) and y0B(x0, r0), we have

(v(y, y0), v+(y, y0))∈ Ox0B(x, r+ε)× OxB(x0, r0+ε).

Lemma 2.9. — For everyr >0, r0 >0, andε >0, there exists l0 >0 such that the following holds : for everyx, x0Mfsatisfyingd(x, x0)>l0, for every v ∈ Ox0B(x, r) and every v+ ∈ OxB(x0, r0), there exist yB(x, r+ε)andy0B(x0, r0+ε)such that

(v, v+) = (v(y, y0), v+(y, y0)).

The next lemma states that when two balls are far enough one from each other, their shadows relative to each other’s center cannot intersect.

Lemma 2.10. — For every R > 0, there exists l0 > 0 such that, for everyx, yMfsatisfyingd(x, y)>l0, one has

OxB(y, R)∩ OyB(x, R) =.

Proof. — Suppose not, and take ξ in the intersection. In the triangle (x, y, ξ), we would have

d(x,[y;ξ[)6R and d(y,[x;ξ[)6R .

Triangles ofMfareδ-hyperbolic for some positive constantδdepending only on the upper bound of the sectional curvature. Therefore, this situation is possible only if d(x, y) 6 2R + 2δ. Thus, the lemma is proved with

l0= 2R+ 2δ.

Measure of shadows

The Shadow Lemma, initially due to Sullivan, estimates the measure given by Patterson–Sullivan–Gibbs densities to shadows of balls in terms of integrals of the normalized potential. It has been proven by Mohsen in our setting, and asserts the following.

Lemma 2.11 (Mohsen’s Shadow Lemma, [14, Lemma 3.10]). — Let (µFx)x∈

Me be the Patterson–Sullivan–Gibbs conformal density associated withF, and Kbe a compact subset ofMf. There existsR0>0such that, for allR>R0, there existsC >0such that for allγ∈Γandx, yK

1 Ce

Rγy

x (F−P(F))e 6µFx (OxB(γy, R))6Ce Rγy

x (F−P(F))e .

(16)

A careful examination of the proof of this lemma shows that the condition R>R0is actually only necessary for the lower bound. In the following, we will only use this lemma for its upper bound, so we can forget about this restriction.

However, we will also need some lower bound estimates for the mF- measure of dynamical balls. To this end, we will use the following variant of the Shadow Lemma for product of shadows of balls, which replaces the restriction on the size of balls by the assumption that the ball intersects the nonwandering set of the geodesic flow.

Lemma 2.12(Shadow product lemma). — Let(µFx)

x∈Me

and(µFx◦ι)

x∈Me be the Patterson–Sullivan–Gibbs conformal densities respectively associ- ated with F and Fι. Assume thatB(x, R)Mf intersects the base of the nonwandering setΩ. Then there existe C >0 andS, G⊂Γ finite such that for everyγ∈Γ\S there existg, hGsuch that

1 Ce

Rγx

x (F−P(F))e 6µF◦ιx (OγxB(gx, R))µFx(OxB(γhx, R)) 6Ce

Rγx

x (F−Pe (F)).

Proof. — If T1B(x, R)∩Ωe 6= ∅, then we can find η, ξ ∈ Λ(Γ) distinct such that the geodesic (ξη) intersects B(x, R). In particular there exists ε >0 such that η ∈ OξB(x, Rε) andξ∈ OηB(x, Rε). By continuity of the shadows, there exist two neighbourhoodsU, V of respectivelyξ and ηin Mf∪Mfsuch that

yU , OξB(x, Rε)⊂ OyB(x, R)

and ∀zV , OηB(x, Rε)⊂ OzB(x, R). By using the same technique as in the proof of Lemma 2.6, we can find S, G⊂Γ finite such that, for everyγ∈Γ\S, there existg, hGsuch that g−1γxU andh−1γ−1xV.

Since the Patterson–Sullivan–Gibbs densities charge any open set that intersects the limit set,

α= minn

µF◦ιg−1x(OξB(x, Rε)) gGo

>0

and β=µFx(OηB(x, Rε))>0. Letγ∈Γ\S and take g, hGsuch thatg−1γxU andh−1γ−1xV. By the invariance property of the densities, we have on the one hand µF◦ιx (OγxB(gx, R)) =µF◦ιg−1x(Og−1γxB(x, R))>µF◦ιg−1x(OξB(x, R−ε))>α ,

(17)

and on the other hand

µFγhx(OxB(γhx, R)) =µFx(Oh−1γ−1xB(x, R))>β . But the conformal density property of (µFx) ensures that

µFx(OxB(γhx, R)) = Z

ζ∈OxB(γhx,R)

e−CF−P(F),ζ(x,γhx)Fγhx(ζ), where CF,ζ(x, y) = limt→∞Rζt

y Fe−Rζt

x Fe is the Gibbs cocycle associated withF. By applying [14, Lemma 3.4] (2), we get the existence of a constant C1>1 independent ofγ andhsuch that

1 C1

e Rγhx

x (F−P(F))e 6 µFx(OxB(γhx, R))

µFγhx(OxB(γhx, R)) 6C1e Rγhx

x (F−P(F))e .

This implies that αβ C1e

Rγhx

x (F−P(F))e 6µF◦ιx (OγhxB(x, R))µFx(OxB(γhx, R)) 6C1e

Rγhx

x (F−P(F))e .

Finally, after noting that d(γx, γhx) = d(x, hx) is bounded from above independently fromγ, we apply Lemma 2.3 to obtain a constantC2 >0 that only depends onMf, F,e xandRsuch that

Z γhx x

(Fe−P(F))− Z γx

x

(Fe−P(F))

6C2.

This concludes the proof withC=αβC1eC2.

3. Number of returns of a periodic orbit

The aim of this section is to introduce a useful mathematical definition of the “number of times that a periodic geodesic enters in a given setW”.

Observe that as soon asW is non convex, or has holes, it may be highly non trivial and cannot be done in a naive way.

Let Wfbe a relatively compact subset ofT1Mf. Ifγ∈Γh, we define the number of copies of the axis ofγ intersectingWfas the quantity

n

We(γ) = #n γ0∈Γ

g∈Γ, γ0 =g−1γg and Aγ0∩W 6=f ∅o . By definition, this number depends only on the conjugacy class ofγ∈Γh. Of course, it is also Γ-invariant, in the sense that

n

We(γ) =n

gWe (γ).

(18)

We shall now extend this definition to relatively compact subsets ofT1M in the following way. First note that ifW ⊂ T1M is open and relatively compact, then it admits an open relatively compactlift(actually many of them), i.e. an open relatively compact setW ⊂f T1Mfsuch thatPΓ(W) =f W where PΓ : T1Mf → T1M is the covering map. Indeed, it is enough to cover W by trivializing open sets for the covering map PΓ, to take for each of these sets the image of its intersection with W by one of the inverse branches ofPΓ, and then letWfto be the union of these preimages.

However, there might not exist an open liftWfofWsuch thatPΓ:W → Wf is 1−1 if, for example, the baseπ(W) contains a ballB(x, R) whose radius Ris larger than the injectivity radius atx.

GivenW ⊂T1M open relatively compact, and any periodic orbitp∈ P, this leads us to define thenumber of returns of pintoW by

nW(p) = infn

We(γp),

whereγp is any hyperbolic isometry in the conjugcy class associated with p, and the infimum is taken over all open relatively compact lifts WfofW toT1Mf. By definition, nW(p)>1 if and only ifpintersects W.

Note that the quantities n

We(γ) andnW(p) do not depend on the mul- tiplicity ofγ or p. Indeed, two isometries γ and γ0 are conjugated by an element g if and only ifγk and γ0k are conjugated by this element g, for k>1. Moreover, the axiiAγ and Aγk are equal for allk >1. Therefore, we get

k>1, n

We(γk) =n

We(γ).

Equivalently, ifp∈ Pis a periodic orbit with multiplicity whose associated primitive orbit isp0∈ P0, we have

nW(p) =nW(p0).

Although there might not exist a lift of W that realizes the number of returnsnW(p) ofp∈ PintoWas a number of copies, the numbers of copies of the axis ofγ∈Γhintersecting two distinct open relatively compact lifts ofW are uniformly commensurable with each other.

Lemma 3.1. — LetWf1,Wf2be two open relatively compact lifts toT1Mf of an open relatively compact setW.

IfWf1⊂Wf2, then for allγ∈Γh, we haven

We1(γ)6n

We2(γ).

If Wf1 and Wf2 are isometric, then for all γ ∈ Γh, we have n

We1(γ) = n

We2(γ).

(19)

More generally, there is C=C

We1,We2

such that for allγ∈Γh, we have 1

Cn

We2(γ)6n

We1(γ)6Cn

We2(γ).

Proof. — The first assertions are clear. For the last one, it is enough to show that n

We10) 6 Cn

We20) for any γ0 ∈ Γ0h. Take γ0 = g−1γ0g for someg∈Γ such that there existsvAγ0∩Wf1. SinceWf1andWf2are both lifts of the same setW, the set

Hv=n h∈Γ

hv∈Wf2

o

is non-empty. Note that ifhHv, thenh(v)∈Wf2∩Aγ00whereγ00=0h−1 is a conjugate ofγ0. This ensures that

n γ0∈Γ

g∈Γ, γ0 =g−1γ0g and Aγ0∩Wf16=∅o

⊂ [

h∈H

h−1n

γ00∈Γ

g∈Γ, γ00=g−1γ0g and Aγ00∩Wf26=∅o h ,

whereH =∪

v∈We1

Hvdepends only onWf1andWf2but not onγ0. In order to conclude, it is enough to show thatH is finite. Indeed, letWf3=Wf1∪Wf2. It is a compact subset ofT1Mfand we have

H=n h∈Γ

v∈Wf1, hv∈Wf2

o⊂n h∈Γ

Wf3hfW36=∅o , which is finite since the action of Γ onT1Mfis proper.

In particular, if one takes a relatively compact lift Wf1 ofW, then there exists a constantCwhich depends only onWf1such that for everyp0∈ P0 and everyγ0 in the conjugacy class associated withp0 we have

1 Cn

We10)6nW(p0)6Cn

We10).

Lemma 3.2. — IfW ⊂T1Mis covered by a finite collection(Wi)i=1,...,n

of open relatively compact subsets of T1M, then there exists C = CW,W1,...,Wn>0such that for allp∈ P, we have

nW(p)6C

n

X

i=1

nWi(p).

Proof. — We may assume that p∈ P0. Fix γp ∈ Γ0h in the conjugacy class associated withp. For eachi, take an open relatively compact liftWfi

ofWi to T1Mf, as well as a constant Ci independent from pand γp such that

n

Weip)6CinWi(p).

(20)

Observe that

Wf=P−1Γ (W)∩

n

[

i=1

Wfi

is an open relatively compact lift ofW, and that if an axis Aγ meets Wf then it meets at least one of theWfi. Therefore

nW(p)6n

We(γp)6

n

X

i=1

n

Weip)6max(Ci)

n

X

i=1

nWi(p).

4. Ergodicity of Gibbs measures for recurrent potentials

LetF :T1M →Rbe a Hölder continuous potential onT1M. According to the Hopf–Tsuji–Sullivan Theorem, we know that mF is ergodic and conservative if and only if the Poincaré series associated withF diverges at the critical exponents=P(F), in which case (Γ, F) is said to bedivergent following the terminology in [14]. In this section, we will prove Theorem 1.2 which asserts that it is also equivalent to the divergence of the series

X

p∈P

nW(p)e R

p(F−P(F))

,

forW an open relatively compact set intersecting Ω.F is said to berecur- rentrelatively to W when this series diverges.

Note that periodic orbits meeting W are the only periodic orbits to consider in the above sum, because otherwisenW(p) = 0.

Theorem 1.2. — LetM be a negatively curved orbifold with pinched negative curvature, andF :T1M →Ra Hölder continuous potential with P(F)<+∞. Then the Gibbs measuremF is ergodic and conservative if and only ifF is recurrent with respect to some open relatively compact set intersectingΩ.

In particular, the recurrence property does not depend on the choice of the open relatively compact subsetW.

Observe first that, for any real numberk, this equivalence is satisfied for a potentialF if and only if it is satisfied for the potential F+k, as the Gibbs measuresmF and mF+k are equal. We may therefore assume from now on thatP(F) = 0. We will also denote byFe the Γ-invariant lift ofF toT1Mf.

Références

Documents relatifs

The next proposition relies upon the criterion for the discreteness of the spectrum of ∆ M in terms of the isocapacitary function of M given in Theorem 2.1, and provides us with

As an application, we prove Ruelle’s inequality and Pesin’s entropy formula for the geodesic flow in manifolds with pinched negative sectional curvature.. Motivation and statements

Actually, the characterization given by Theorem 1.2 is very particular to constant curvature spaces and it does not generalize, as it is, to locally symmetric spaces: indeed, it is

In Section §4 we discuss several notions of equivalence of rays related to the Busemann equivalence; then we give a characterization (Theorem 28) of the Busemann equivalence for rays

As far as we know, even in the classical case of Riemannian negatively curved symmetric spaces of rank one (which are 1/4-pinched, cp. [15]), there does not exist an elementary proof

For any sequence of initial data microlocalized on the unit cotangent bundle, we look at the quantum evolution (below the Ehrenfest time) under small pertur- bations of the

Our approach for constructing the maximal entropy measure and the harmonic measure for the geodesic flow on Riemannian negatively curved manifolds can be

— Let M be a connected, complete, nonpositively curved manifold, whose geodesic flow admits at least three different periodic orbits, that do not bound a flat strip... Recall here