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

Correlation of the renormalized Hilbert length
for convex projective surfaces

Xian Dai Mathematical institute of Heidelberg University, Heidelberg, Germany, 69117 [email protected]  and  Giuseppe Martone Department of Mathematics, University of Michigan, Ann Arbor, MI 41809 [email protected]
Abstract.

In this paper we focus on dynamical properties of (real) convex projective surfaces. Our main theorem provides an asymptotic formula for the number of free homotopy classes with roughly the same renormalized Hilbert length for two distinct convex real projective structures. The correlation number in this asymptotic formula is characterized in terms of their Manhattan curve. We show that the correlation number is not uniformly bounded away from zero on the space of pairs of hyperbolic surfaces, answering a question of Schwartz and Sharp. In contrast, we provide examples of diverging sequences, defined via cubic rays, along which the correlation number stays larger than a uniform strictly positive constant. In the last section, we extend the correlation theorem to Hitchin representations.

1. Introduction

In this paper, we study the correlation of length spectra of pairs of convex projective surfaces. We describe how length spectra of two convex projective structures ρ1\rho_{1} and ρ2\rho_{2} correlate on a closed connected orientable surface SS of genus g2g\geq 2. More precisely, we investigate the asymptotic behavior of the number of closed curves whose ρ1\rho_{1}-Hilbert length and ρ2\rho_{2}-Hilbert length are roughly the same. This question was first considered by Schwartz and Sharp in the context of hyperbolic surfaces in [51] (see also [24]), and more generally for negatively curved metrics in [22, 46]. The holonomy of a real convex projective structure is an example of Hitchin representation and our correlation theorem holds in this more general setting.

We now discuss our results in greater detail. A (marked real) convex projective structure on the surface SS is described by a strictly convex set Ωρ\Omega_{\rho} in the real projective plane which admits a cocompact action by a discrete subgroup of SL(3,)\mathrm{SL}(3,\mathbb{R}) isomorphic to Γ=π1(S)\Gamma=\pi_{1}(S). We denote by ρ:ΓSL(3,)\rho\colon\Gamma\to\mathrm{SL}(3,\mathbb{R}) the corresponding representation and by XρX_{\rho} the surface SS equipped with the convex projective structure. Goldman [27] proved that the space of convex projective surfaces (S)\mathfrak{C}(S) is an open cell of dimension 8χ(S)-8\chi(S). Hyperbolic structures on SS define, via the Klein model of the hyperbolic plane, a 3χ(S)-3\chi(S)-dimensional subspace of (S)\mathfrak{C}(S), called the Fuchsian locus, which is naturally identified with the Teichmüller space 𝒯(S)\mathcal{T}(S) of SS. As customary, we will blur the distinction between 𝒯(S)\mathcal{T}(S) and the Fuchsian locus.

Every convex projective structure ρ\rho induces a Hilbert length ρH\ell^{H}_{\rho} for non-trivial conjugacy classes of group elements in Γ\Gamma. Given [γ][Γ][\gamma]\in[\Gamma] a non-trivial conjugacy class, we define ρH([γ])\ell^{H}_{\rho}([\gamma]) as the length of the unique closed geodesic with respect to the Hilbert metric on XρX_{\rho} that corresponds to the free homotopy class of [γ][\gamma] on SS. The (marked) Hilbert length spectrum of ρ\rho is the function ρH:[Γ]>0\ell^{H}_{\rho}\colon[\Gamma]\to\mathbb{R}_{>0}. We investigate a slightly different notion of length. Denoting the topological entropy of (the Hilbert geodesic flow of) ρ\rho as

h(ρ)=lim supT1Tlog#{[γ][Γ]ρH([γ])T},h(\rho)=\limsup_{T\to\infty}\frac{1}{T}\log\#\{[\gamma]\in[\Gamma]\mid\ell_{\rho}^{H}([\gamma])\leq T\},

we define the renormalized Hilbert length spectrum of ρ\rho as LρH=h(ρ)ρHL^{H}_{\rho}=h(\rho)\ell^{H}_{\rho}.

Our first theorem concerns the correlation of the renormalized Hilbert length spectra of two different convex projective structures.

Theorem 1.1 (Correlation Theorem for Convex Real Projective Structures).

Fix a precision ε>0\varepsilon>0. Consider two convex projective structures ρ1\rho_{1} and ρ2\rho_{2} on a surface SS with distinct renormalized Hilbert length spectra Lρ1HLρ2HL_{\rho_{1}}^{H}\neq L_{\rho_{2}}^{H}. There exist constants C=C(ε,ρ1,ρ2)>0C=C(\varepsilon,\rho_{1},\rho_{2})>0 and M=M(ρ1,ρ2)(0,1)M=M(\rho_{1},\rho_{2})\in(0,1) such that

#{[γ][Γ]|Lρ1H([γ])(x,x+h(ρ1)ε),Lρ2H([γ])(x,x+h(ρ2)ε)}CeMxx3/2.\#\Big{\{}[\gamma]\in[\Gamma]\,\Big{|}\,L_{\rho_{1}}^{H}([\gamma])\in\big{(}x,x+h(\rho_{1})\varepsilon\big{)},\ L_{\rho_{2}}^{H}([\gamma])\in\big{(}x,x+h(\rho_{2})\varepsilon\big{)}\Big{\}}\sim C\frac{e^{Mx}}{x^{3/2}}.

where f(x)g(x)f(x)\sim g(x) means f(x)/g(x)1f(x)/g(x)\to 1 as xx\to\infty.

Remark 1.2.
  1. (1)

    There is an involution on the space of convex projective structures, called the contragredient involution, which sends ρ\rho to ρ=(ρ1)t\rho^{*}=(\rho^{-1})^{t}. Cooper-Delp and Kim [20, 33] show that Lρ1H=Lρ2HL_{\rho_{1}}^{H}=L_{\rho_{2}}^{H} if and only if ρ2\rho_{2} is ρ1\rho_{1} or ρ1\rho_{1}^{*}.

  2. (2)

    It follows from the generalized prime geodesic theorem [41, 44], that for any fixed ε\varepsilon, if ρ1\rho_{1} and ρ2\rho_{2} converge to convex projective structures with the same renormalized Hilbert length spectrum, then M(ρ1,ρ2)M(\rho_{1},\rho_{2}) converges to one and C(ε,ρ1,ρ2)C(\varepsilon,\rho_{1},\rho_{2}) diverges.

The topological entropy of any hyperbolic structure equals to one [31] and the renormalized Hilbert length coincides with the Hilbert length. The entropy varies continuously on (S)\mathfrak{C}(S) and Crampon [21] shows that it is strictly less than one away from the Fuchsian locus. Nie and Zhang [43, 55] prove that the entropy can be arbitrarily close to zero. In Example 5.2, we show that there exist a Fuchsian representation ρ1\rho_{1} and a representation ρ2(S)\rho_{2}\in\mathfrak{C}(S) with topological entropy different from one such that

limx#{[γ][Γ]|ρ1H([γ])(x,x+ε),ρ2H([γ])(x,x+ε)}=0.\lim_{x\to\infty}\#\Big{\{}[\gamma]\in[\Gamma]\,\Big{|}\,\ell_{\rho_{1}}^{H}([\gamma])\in\big{(}x,x+\varepsilon\big{)},\ \ell_{\rho_{2}}^{H}([\gamma])\in\big{(}x,x+\varepsilon\big{)}\Big{\}}=0.

In particular, for these representations the size of the set {[γ][Γ]|ρiH([γ])(x,x+ϵ),i=1,2}\Big{\{}[\gamma]\in[\Gamma]\,\Big{|}\,\ell_{\rho_{i}}^{H}([\gamma])\in(x,x+\epsilon),\ i=1,2\Big{\}} does not grow exponentially. Renormalized length spectra are natural objects from a dynamical point of view (see Remark 3.5 for a more detailed justification) and thus they play a key role in our discussion.

We refer to the exponent M(ρ1,ρ2)M(\rho_{1},\rho_{2}) from Theorem 1.1 as the correlation number of ρ1\rho_{1} and ρ2\rho_{2}. An important goal of this paper is to study the correlation number as we vary ρ1\rho_{1} and ρ2\rho_{2} in (S)\mathfrak{C}(S). One interesting question asked in [51] for the case of the Teichmüller space is whether the correlation number M(ρ1,ρ2)M(\rho_{1},\rho_{2}) is uniformly bounded away from zero as its arguments range over all hyperbolic structures. We answer this question in the negative in Section 4. We prove the following.

Theorem 1.3 (Decay of Correlation Number).

There exist sequences (ρn)n=1(\rho_{n})_{n=1}^{\infty} and (ηn)n=1(\eta_{n})_{n=1}^{\infty} in the Teichmüller space 𝒯(S)\mathcal{T}(S) such that the correlation number satisfies

limnM(ρn,ηn)=0.\lim_{n\to\infty}M(\rho_{n},\eta_{n})=0.

The sequences in Theorem 1.3 are given by pinching a hyperbolic structure along two different pants decompositions which are filling. Intuitively, these are two families of hyperbolic structures diverging from each other in the Teichmüller space thus suggesting a small correlation number when going to infinity. A key to prove Theorem 1.3 is a characterization of Sharp [53] of the correlation number in terms of the Manhattan curve [17]. In Theorem 4.2, we extend Sharp’s characterization of the correlation number to the case of two convex projective structures.

In Section 5, we study explicit examples of correlation number in the space of convex projective structures. In contrast to Theorem 1.3, we provide pairs of diverging sequences for which the correlation numbers are uniformly bounded below away from zero. These sequences (ρt)t0(\rho_{t})_{t\geq 0}, called cubic rays, are defined using holomorphic cubic differentials via Labourie-Loftin’s parameterization of (S)\mathfrak{C}(S) [35, 40]. More precisely, Labourie and Loftin describe a mapping class group equivariant homeomorphism between (S)\mathfrak{C}(S) and the vector bundle of holomorphic cubic differentials over 𝒯(S)\mathcal{T}(S). The sequences (ρt)t0(\rho_{t})_{t\geq 0} lie in fibers of (S)\mathfrak{C}(S) with base point ρ0𝒯(S)\rho_{0}\in\mathcal{T}(S) and correspond to rays (tq)t0(tq)_{t\geq 0} for qq a fixed ρ0\rho_{0}-holomorphic cubic differential. We will recall Labourie-Loftin’s parameterization of (S)\mathfrak{C}(S) and the definition of cubic rays more precisely in section 5.

Using work of Tholozan [54], we show in Lemma 5.1 that, for a cubic ray (ρt)t0(\rho_{t})_{t\geq 0}, the renormalized Hilbert length of ρt\rho_{t} is bi-Lipschitz to the one of ρ0\rho_{0} with Lipschitz constants independent of tt. We deduce that the correlation number of any two convex projective structures in different fibers is uniformly bounded from below by the correlation number of its base hyperbolic structures.

Theorem 1.4.

Let (ρt)t0(\rho_{t})_{t\geq 0}, (ηr)r0(\eta_{r})_{r\geq 0} be two cubic rays associated to two different hyperbolic structures ρ0η0\rho_{0}\neq\eta_{0}. Then, there exists a constant C>0C>0 such that for all t,r0t,r\geq 0,

M(ρt,ηr)CM(ρ0,η0).M(\rho_{t},\eta_{r})\geq CM(\rho_{0},\eta_{0}).

A similar statement holds for most pairs of cubic rays with the same base hyperbolic structure.

Theorem 1.5.

Let ρt\rho_{t} and ηt\eta_{t} be two cubic rays associated to two different holomorphic cubic differential q1q_{1} and q2q_{2} on a hyperbolic structure X0X_{0} such that q1,q2q_{1},q_{2} have unit L2L^{2}-norm with respect to X0X_{0} and q1q2q_{1}\neq-q_{2}. Then there exists a constant C>0C>0 such that for all t>0t>0,

M(ρt,ηt)C.M(\rho_{t},\eta_{t})\geq C.

In Section 5.3, which is for the most part independent of the rest of the paper, we show how Lemma 5.1 can be used to study renormalized Hilbert geodesic currents νρt\nu_{\rho_{t}} along a cubic ray (ρt)t>0(\rho_{t})_{t>0}. Geodesic currents are geometric measures on the space of complete geodesics on the universal cover of SS and each geodesic current ν\nu has a corresponding length spectrum ν:[Γ]0\ell_{\nu}\colon[\Gamma]\to\mathbb{R}_{\geq 0} with systole Sys(ν)=infc[Γ]ν(c)\mathrm{Sys}(\nu)=\inf_{c\in[\Gamma]}\ell_{\nu}(c). See §2.4 for a detailed discussion. It follows from [9, 15, 42] that for every convex projective structure ρ\rho, there exists a unique geodesic current υρ\upsilon_{\rho} whose length spectrum coincides with the renormalized Hilbert length spectrum of ρ\rho. We call υρ\upsilon_{\rho} the renormalized Hilbert geodesic current (also known as the renormalized Liouville current) of ρ\rho. We prove the following.

Theorem 1.6.

As tt goes to infinity, the renormalized Hilbert geodesic current (υρt)t>0(\upsilon_{\rho_{t}})_{t\in>0} along a cubic ray (ρt)t0(\rho_{t})_{t\geq 0} converges, up to subsequences, to a geodesic current υ\upsilon with Sys(υ)>0\mathrm{Sys}(\upsilon)>0.

This should be compared with what happens for hyperbolic structures: given a sequence of hyperbolic structures which leaves every compact subset of 𝒯(S)\mathcal{T}(S), up to rescaling and passing to subsequences, the corresponding sequence of geodesic currents converges to a geodesic current with vanishing systole. Burger, Iozzi, Parreau, and Pozzetti [18] show that this fact no longer holds for general sequences of Hilbert geodesic currents. Theorem 1.6 provides new examples of diverging sequences of convex projective structures whose associated geodesic currents converge (projectively and up to subsequences) to a geodesic current with positive systole. In light of [2, Theorem 1.13], Theorem 1.6 extends a theorem of Burger, Iozzi, Parreau, Pozzetti [18, Theorem 1.12] who prove that the Hilbert geodesic currents of a diverging sequence of convex projective structures of a triangle group converge (projectively and up to subsequences) to a geodesic current with positive systole.

Finally, in Section 6, we generalize the correlation theorem (Theorem 1.1) to Hitchin components for a large class of length functions. We will replace the Hilbert geodesic flow of a convex projective structure, which is Anosov, with more general metric Anosov translation flows.

Given a hyperbolic structure ρ𝒯(S)\rho\in\mathcal{T}(S), seen as a representation ρ:ΓPSL(2,)\rho\colon\Gamma\to\mathrm{PSL}(2,\mathbb{R}), we obtain a representation iρ:ΓPSL(d,)i\circ\rho\colon\Gamma\to\mathrm{PSL}(d,\mathbb{R}) by post-composing ρ\rho with the unique (up to conjugation) irreducible representation i:PSL(2,)PSL(d,)i\colon\mathrm{PSL}(2,\mathbb{R})\to\mathrm{PSL}(d,\mathbb{R}). The Teichmüller space 𝒯(S)\mathcal{T}(S) embeds in this way in the character variety of SS and PSL(d,)\mathrm{PSL}(d,\mathbb{R}). The connected component d(S)\mathcal{H}_{d}(S) of the character variety containing this image is known as the Hitchin component. Hitchin [30] showed that d(S)\mathcal{H}_{d}(S) is homeomorphic to an open cell of dimension (dimPSL(d,))χ(S)-(\dim\mathrm{PSL}(d,\mathbb{R}))\chi(S). Choi and Goldman [19] identify 3(S)\mathcal{H}_{3}(S) with the space (S)\mathfrak{C}(S) which will be our main focus from Section 2 to Section 5.

In order to state the general correlation theorem for Hitchin representations (Theorem 1.7), we need to introduce some Lie theoretical notations.

Let

𝔞={xdx1++xd=0} and 𝔞+={x𝔞x1xd}\mathfrak{a}=\{\vec{x}\in\mathbb{R}^{d}\mid x_{1}+\dots+x_{d}=0\}\qquad\text{ and }\qquad\mathfrak{a}^{+}=\{\vec{x}\in\mathfrak{a}\mid x_{1}\geq\dots\geq x_{d}\}

denote the (standard) Cartan subspace for PSL(d,)\mathrm{PSL}(d,\mathbb{R}) and the (standard) positive Weyl chamber, respectively. Let λ:PSL(d,)𝔞+\lambda\colon\mathrm{PSL}(d,\mathbb{R})\to\mathfrak{a}^{+} be the Jordan projection given by λ(g)=(logλ1(g),,logλd(g))\lambda(g)=(\log\lambda_{1}(g),\dots,\log\lambda_{d}(g)) consisting of the logarithms of the moduli of the eigenvalues of gg in nonincreasing order. We consider linear functionals in

Δ={c1α1++cd1αd1ci0,ici>0},\Delta=\left\{c_{1}\alpha_{1}+\dots+c_{d-1}\alpha_{d-1}\mid c_{i}\geq 0,\sum_{i}c_{i}>0\right\},

where αi:𝔞\alpha_{i}\colon\mathfrak{a}\to\mathbb{R} are the simple roots defined by αi(x)=xixi+1\alpha_{i}(\vec{x})=x_{i}-x_{i+1} with i=1,,d1i=1,\dots,d-1. Observe that if ϕΔ\phi\in\Delta, then ϕ(x)>0\phi(\vec{x})>0 for all x\vec{x} in the interior of 𝔞+\mathfrak{a}^{+}. The length function ρϕ\ell^{\phi}_{\rho} for ϕΔ\phi\in\Delta and ρd(S)\rho\in\mathcal{H}_{d}(S) is defined by ρϕ([γ])=ϕ(λ(ρ(γ)))\ell^{\phi}_{\rho}([\gamma])=\phi(\lambda(\rho(\gamma))). The length function is strictly positive because λ(ρ(γ))\lambda(\rho(\gamma)) is in the interior of 𝔞+\mathfrak{a}^{+} for all [γ][Γ][\gamma]\in[\Gamma] (see [23, 34]). The topological entropy hϕ(ρ)h^{\phi}(\rho) and renormalized ϕ\phi-length Lρϕ([γ])=hϕ(ρ)ρϕ([γ])L^{\phi}_{\rho}([\gamma])=h^{\phi}(\rho)\ell^{\phi}_{\rho}([\gamma]) are defined in similar manner as for convex real projective structures. Given a Hitchin representation ρ\rho, we denote by ρ\rho^{*} its contragredient given by ρ=(ρ1)t\rho^{*}=(\rho^{-1})^{t}.

We are now ready to state the correlation theorem for Hitchin representations.

Theorem 1.7.

Given a linear functional ϕΔ\phi\in\Delta and a fixed precision ε>0\varepsilon>0, for any two different Hitchin representations ρ1,ρ2:ΓPSL(d,)\rho_{1},\rho_{2}:\Gamma\to\mathrm{PSL}(d,\mathbb{R}) such that ρ2ρ1\rho_{2}\neq\rho_{1}^{*}, there exist constants C=C(ε,ρ1,ρ2,ϕ)>0C=C(\varepsilon,\rho_{1},\rho_{2},\phi)>0 and M=M(ρ1,ρ2,ϕ)(0,1)M=M(\rho_{1},\rho_{2},\phi)\in(0,1) such that

#{[γ][Γ]|Lρ1ϕ([γ])(x,x+hϕ(ρ1)ε),Lρ2ϕ([γ])(x,x+hϕ(ρ2)ε)}CeMxx3/2.\#\Big{\{}[\gamma]\in[\Gamma]\,\Big{|}\,L_{\rho_{1}}^{\phi}([\gamma])\in\big{(}x,x+h^{\phi}(\rho_{1})\varepsilon\big{)},\ L_{\rho_{2}}^{\phi}([\gamma])\in\big{(}x,x+h^{\phi}(\rho_{2})\varepsilon\big{)}\Big{\}}\sim C\frac{e^{Mx}}{x^{3/2}}.
Remark 1.8.
  1. (1)

    For ρd(S)\rho\in\mathcal{H}_{d}(S), when ϕ=αi\phi=\alpha_{i} and i=1,,d1i=1,\dots,d-1, we know that hαi(ρ)=1h^{\alpha_{i}}(\rho)=1 thanks to [47, Thm B]. Thus, Lραi=ραiL^{\alpha_{i}}_{\rho}=\ell^{\alpha_{i}}_{\rho} is the simple root length without any renormalization. The correlation theorem in this case can be seen as a natural generalization of Schwartz and Sharp’s correlation theorem for hyperbolic surfaces.

  2. (2)

    Note that Theorem 1.1 is a corollary of Theorem 1.7 when we set d=3d=3 and consider the positive root ϕ(x)=(α1+α2)(x)=x1x3\phi(\vec{x})=(\alpha_{1}+\alpha_{2})(\vec{x})=x_{1}-x_{3}.

It would be interesting to extend the results on the correlation number proved in sections 4 and 5 to the context Hitchin representations. In this spirit, we end the paper by raising Question 6.15 and Conjecture 6.16 which are motivated by Theorem 1.3 and Theorem 1.5, respectively.

Structure of the paper

In Sections 2 through 5, we focus on convex projective structures. In this case, the length function can be defined geometrically via the Hilbert distance and Benoist proved in [7] that the associated geodesic flow is Anosov. These two facts will simplify the exposition. The main results (Theorems 1.3, 1.4 and 1.5) in Sections 4 and 5 concern the behavior of the correlation number along geometrically defined sequences of convex projective structures. We establish the correlation theorem in full generality in Section 6, after recalling the precise definitions of Hitchin representations and their length functions and the theory of metric Anosov translation flows.

2. Preliminaries for convex real projective structures

Consider a connected, closed, oriented surface SS with genus g2g\geq 2 and denote by Γ\Gamma its fundamental group. In this preliminary section we focus on convex projective structures on SS. We will discuss in section 6 how parts of the material presented here hold for general Hitchin components.

The structure of Section 2 is as follows. In §2.1, we briefly recall the relevant geometric aspects of the theory of convex projective structures on surfaces. We refer to [7, 27] for further details and background. In §2.2, we collect dynamical properties of convex projective surfaces which will play an important role in sections 3 and 4. We briefly survey the theory of geodesic currents in §2.4 which will be used in the proof of Theorem 1.3 and in §5.3. Finally, in §2.5 we prove the independence lemma (Lemma 2.12) which plays a key role in the proof of Theorem 1.1.

2.1. Convex projective surfaces

A properly convex set Ω\Omega in 2\mathbb{RP}^{2} is a bounded open convex subset of an affine chart. A properly convex set whose boundary does not contain open line segments is strictly convex. We will exclusively focus on strictly convex sets in this paper. We equip a strictly convex set Ω\Omega with its Hilbert metric dΩd_{\Omega}. More precisely, if x,yΩx,y\in\Omega, the projective line xy¯\overline{xy} passing through xx and yy intersects the boundary of Ω\Omega in two points a,ba,b where a,x,y,ba,x,y,b appear in this order along xy¯\overline{xy}. The Hilbert distance between xx and yy is

dΩ(x,y)=12log[a,x,y,b]d_{\Omega}(x,y)=\frac{1}{2}\log[a,x,y,b]

where [a,x,y,b][a,x,y,b] denotes the crossratio of four points on a projective line. With the Hilbert metric, geodesics are segments of a projective line intersecting Ω\Omega. Typically, the Hilbert metric is not Riemannian, but it derives from a Finsler norm. Thus, one can study the unit tangent bundle T1ΩT^{1}\Omega of a strictly convex set Ω\Omega.

The main objects of interest are representations ρ:ΓSL(3,)\rho\colon\Gamma\to\mathrm{SL}(3,\mathbb{R}) such that ρ(Γ)\rho(\Gamma) preserves a properly convex set Ωρ\Omega_{\rho} on which it acts properly discontinuously with quotient homeomorphic to the closed surface SS. In this case, we say that ρ\rho is a (marked real) convex projective structure which divides Ωρ\Omega_{\rho} and denote by XρX_{\rho} the surface SS equipped with the convex projective structure ρ\rho. Since SS is a closed surface of negative Euler characteristic, Ωρ\Omega_{\rho} is strictly convex and if γΓ\gamma\in\Gamma is non-trivial, then the moduli λ1(ρ(γ))>λ2(ρ(γ))>λ3(ρ(γ))>0\lambda_{1}(\rho(\gamma))>\lambda_{2}(\rho(\gamma))>\lambda_{3}(\rho(\gamma))>0 of the eigenvalues of ρ(γ)\rho(\gamma) are distinct. (See for example [27, 3.2 Theorem] and references therein).

The Hilbert distance on Ωρ\Omega_{\rho} induces the Hilbert length spectrum ρH\ell_{\rho}^{H} for non-trivial conjugacy classes of group elements in Γ\Gamma. Algebraically, if [Γ][\Gamma] is the set of conjugacy classes of non-identity elements in Γ\Gamma, namely [γ][Γ][\gamma]\in[\Gamma] is the conjugacy class of γid\gamma\neq\text{id}, then

ρH([γ])=12logλ1(ρ(γ))λ3(ρ(γ)).\ell_{\rho}^{H}([\gamma])=\frac{1}{2}\log\frac{\lambda_{1}(\rho(\gamma))}{\lambda_{3}(\rho(\gamma))}.

Benoist [7] proved that if ρ\rho is a convex projective structure on SS, then (Ωρ,dΩρ)(\Omega_{\rho},d_{\Omega_{\rho}}) is Gromov hyperbolic, the Gromov boundary and the topological boundary of Ωρ\Omega_{\rho} coincide, and Ωρ\partial\Omega_{\rho} is of class C1+αC^{1+\alpha} for some α(0,1]\alpha\in(0,1]. For a(ny) point oΩρo\in\Omega_{\rho}, the orbit map τo\tau_{o} is a quasi-isometric embedding. It follows that the induced limit map between Gromov boundaries ξρ:ΓΩρ\xi_{\rho}\colon\partial\Gamma\to\partial\Omega_{\rho} is a ρ\rho-equivariant bi-Hölder homeomorphism.

If ρ(Γ)\rho(\Gamma) divides a strictly convex set Ωρ\Omega_{\rho} in 2\mathbb{RP}^{2}, then ρ(Γ)=(ρ(Γ)1)t\rho^{\ast}(\Gamma)=(\rho(\Gamma)^{-1})^{t} divides a (typically different) strictly convex set Ωρ\Omega_{\rho^{\ast}}. We refer to this operation on the space of convex projective structures as the contragredient involution. The contragredient involution preserves the Hilbert length because for all [γ][Γ][\gamma]\in[\Gamma]

ρH([γ])=12logλ1(ρ(γ))λ3(ρ(γ))=12log1λ3(ρ(γ))1λ1(ρ(γ))=ρH([γ]).\ell^{H}_{\rho}([\gamma])=\frac{1}{2}\log\frac{\lambda_{1}(\rho(\gamma))}{\lambda_{3}(\rho(\gamma))}=\frac{1}{2}\log\frac{\frac{1}{\lambda_{3}(\rho(\gamma))}}{\frac{1}{\lambda_{1}(\rho(\gamma))}}=\ell^{H}_{\rho^{\ast}}([\gamma]).

A standard computation using the irreducible representation PSL(2,)PSL(3,)\mathrm{PSL}(2,\mathbb{R})\to\mathrm{PSL}(3,\mathbb{R}) shows that if ρ\rho is a hyperbolic structure, then ρ=ρ\rho=\rho^{\ast}. The converse holds by [5, Thm 1.3]: if ρ\rho is a convex projective structure such that ρ=ρ\rho=\rho^{*}, then ρ\rho is a hyperbolic structure.

2.2. The Hilbert geodesic flow and reparametrization function

Suppose that ρ\rho is a convex projective surface. The Hilbert geodesic flow Φρ\Phi^{\rho} is defined on the unit tangent bundle of the surface T1XρT^{1}X_{\rho}. The image Φtρ(w)\Phi^{\rho}_{t}(w) of a point w=(x,v)w=(x,v) is obtained by following the unit speed geodesic for time tt leaving xx in the direction vv. When it is clear from context, we simply write Φρ\Phi^{\rho} as Φ\Phi.

The Hilbert geodesic flow Φ\Phi on T1XρT^{1}X_{\rho} is an example of a topologically mixing Anosov flow by [7, Prop. 3.3 and 5.6]. A standard reference for the theory of Anosov flows is [32, §6]. A key property for our discussion is that topologically mixing Anosov flows can be modeled by Markov partitions and symbolic dynamics in the sense of Bowen [11].

Given a positive Hölder continuous function f:T1Xρf\colon T^{1}X_{\rho}\to\mathbb{R}, one can define a Hölder reparametrization of the flow Φ\Phi by time change. We construct the flow Φf\Phi^{f} following [49, section 2]. First, we define κ:T1Xρ×\kappa:T^{1}X_{\rho}\times\mathbb{R}\to\mathbb{R} as

κ(x,t)=0tf(Φs(x))ds.\kappa(x,t)=\int_{0}^{t}f(\Phi_{s}(x))\mathrm{d}s.

Given the fact that ff is positive and T1XρT^{1}X_{\rho} is compact, the function κ(x,)\kappa(x,\cdot) is an increasing homeomorphism of \mathbb{R}. We therefore have an inverse α:T1Xρ×\alpha:T^{1}X_{\rho}\times\mathbb{R}\to\mathbb{R} that verifies

α(x,κ(x,t))=κ(x,α(x,t))=t\alpha(x,\kappa(x,t))=\kappa(x,\alpha(x,t))=t

for every xT1Xρ×x\in T^{1}X_{\rho}\times\mathbb{R}. The Hölder reparametrization of Φ\Phi by the Hölder continuous function ff is given by Φtf(x)=Φα(x,t)(x)\Phi^{f}_{t}(x)=\Phi_{\alpha(x,t)}(x). We say that ff is a reparametrization function for Φf\Phi^{f}. The new flow Φf={Φtf}t\Phi^{f}=\{\Phi_{t}^{f}\}_{t\in\mathbb{R}} shares the same set of periodic orbits of Φ\Phi. For any periodic orbit τ\tau of Φ\Phi with period λ(τ)\lambda(\tau), its period as a Φf\Phi^{f} periodic orbit is

(2.1) λ(f,τ)=0λ(τ)f(Φs(x))ds.\lambda(f,\tau)=\int_{0}^{\lambda(\tau)}f(\Phi_{s}(x))\mathrm{d}s.

Property (2.1) is a simple application of the definitions of α(x,t)\alpha(x,t) and κ(x,t)\kappa(x,t).

Remark 2.1.

Each oriented closed geodesic γ\gamma on a convex projective structure ρ\rho is associated with a periodic orbit τ\tau of Φ\Phi. On the other hand, an oriented closed geodesic γ\gamma corresponds to a free homotopy class [γ][Γ][\gamma]\in[\Gamma]. We adopt different perspectives depending on necessity in this paper while keeping in mind that they are the same object described from different points of view.

Remark 2.2.
  1. (1)

    One can check from the definition that Φ\Phi is a Hölder reparametrization of Φf\Phi^{f} with the Hölder reparametrization function given by 1/f1/f.

  2. (2)

    Set Ψ=Φf\Psi=\Phi^{f} and consider gg a positive Hölder reparametrization of Ψ\Psi, then Ψg\Psi^{g} is a Hölder reparametrization of Φ\Phi by the Hölder function gfg\cdot f.

Let ρ1\rho_{1} and ρ2\rho_{2} be two convex projective structures on a surface SS. The next lemma states that there exists a positive Hölder continuous reparametrization function fρ1ρ2:T1Xρ1f_{\rho_{1}}^{\rho_{2}}:T^{1}X_{\rho_{1}}\to\mathbb{R} encoding the Hilbert length spectrum of ρ2\rho_{2}.

Lemma 2.3.

Let ρ1\rho_{1} and ρ2\rho_{2} be convex projective structures on a surface SS. There exists a positive Hölder continuous function fρ1ρ2:T1Xρ1f_{\rho_{1}}^{\rho_{2}}:T^{1}X_{\rho_{1}}\to\mathbb{R} such that for every periodic orbit τ\tau corresponding to [γ][Γ][\gamma]\in[\Gamma] one has

λ(fρ1ρ2,τ)=ρ2H([γ]).\lambda\left(f_{\rho_{1}}^{\rho_{2}},\tau\right)=\ell_{\rho_{2}}^{H}([\gamma]).
Proof.

This is a standard argument which we include for the sake of completeness. Let us lift the picture to the universal cover. By [7, Equation (20)] and since the limit map of ρ1\rho_{1} is bi-Hölder, there exists a Hölder continuous ρ1\rho_{1}-equivariant homeomorphism χ:T1Ωρ13Ωρ1\chi:T^{1}\Omega_{\rho_{1}}\to\partial^{3}\Omega_{\rho_{1}} where 3Ωρ1\partial^{3}\Omega_{\rho_{1}} is the set of ordered triples of distinct points in Ωρ1\partial\Omega_{\rho_{1}}. Since the Hilbert geodesic flow is Anosov, it follows from [49, Theorem 3.2] (see also [16, Proposition 5.21]) that for any choice of an auxiliary hyperbolic surface ρ0\rho_{0}, there exists a Hölder continuous positive reparametrization gρ0ρ2:T1Ωρ032g^{\rho_{2}}_{\rho_{0}}\colon T^{1}\Omega_{\rho_{0}}\cong\partial^{3}\mathbb{H}^{2}\to\mathbb{R} of the geodesic flow of ρ0\rho_{0} with periods ρ2H([γ])\ell_{\rho_{2}}^{H}([\gamma]). Considering the Hölder continuous function ξ(3):3Ωρ132\xi^{(3)}\colon\partial^{3}\Omega_{\rho_{1}}\to\partial^{3}\mathbb{H}^{2} induced by the inverse of the limit map ξρ1\xi_{\rho_{1}} and the limit map ξρ0:Γ2\xi_{\rho_{0}}\colon\partial\Gamma\to\partial\mathbb{H}^{2}, we obtain the composition gρ0ρ2ξ(3)χ:T1Ωρ1g^{\rho_{2}}_{\rho_{0}}\circ\xi^{(3)}\circ\chi\colon T^{1}\Omega_{\rho_{1}}\to\mathbb{R} which is the lift of the desired reparametrization function fρ1ρ2f^{\rho_{2}}_{\rho_{1}}. The equality λ(fρ1ρ2,τ)=ρ2H([γ])\lambda\left(f_{\rho_{1}}^{\rho_{2}},\tau\right)=\ell_{\rho_{2}}^{H}([\gamma]) follows from equivariance of the limit maps. ∎

2.3. Thermodynamic formalism

In this subsection, we will introduce several important concepts from thermodynamic formalism in our context that will be needed later. Standard references for thermodynamic formalism and Markov codings are [12, 44].

For a continuous function f:T1Xρf:T^{1}X_{\rho}\to\mathbb{R}, we define its pressure with respect to Φ\Phi as

P(Φ,f)=lim supT1Tlog(τRTeλ(f,τ))P(\Phi,f)=\limsup\limits_{T\longrightarrow\infty}\frac{1}{T}\log\Big{(}\sum\limits_{\tau\in R_{T}}e^{\lambda(f,\tau)}\Big{)}

where RT:={τ periodic orbit of Φ | λ(τ)[T1,T]}R_{T}:=\{\tau\text{ }\text{periodic orbit of }\Phi\text{ }|\text{ }\lambda(\tau)\in[T-1,T]\}. One can check that the topological entropy h(ρ)h(\rho) is P(Φ,0)P(\Phi,0). For simplicity, we omit the geodesic flow Φ\Phi from the notation and write P()P(\cdot) for P(Φ,)P(\Phi,\cdot). The pressure can be characterized as follows.

Proposition 2.4 (Variational principle).

The pressure of a continuous function f:T1Xρf:T^{1}X_{\rho}\to\mathbb{R} satisfies

P(f)=supμΦ(h(μ)+fdμ)P(f)=\sup\limits_{\mu\in\mathcal{M}^{\Phi}}\Big{(}h(\mu)+\int f\mathrm{d}\mu\Big{)}

where Φ\mathcal{M}^{\Phi} is the space of Φ\Phi-invariant probability measures on T1XρT^{1}X_{\rho} and h(μ)=h(Φ,μ)h(\mu)=h(\Phi,\mu) denotes the measure-theoretic entropy of Φ\Phi with respect to μΦ\mu\in\mathcal{M}^{\Phi}.

A Φ\Phi-invariant probability measure μ\mu on T1XρT^{1}X_{\rho} is called an equilibrium state for ff if the supremum is attained at μ\mu.

Remark 2.5.
  1. (1)

    For a Hölder continuous function f:T1Xρf:T^{1}X_{\rho}\to\mathbb{R} there exists a unique equilibrium state μf\mu_{f} by [13, Theorem 3.3].

  2. (2)

    The equilibrium state μ0\mu_{0} for f=0f=0 is called a probability measure of maximal entropy or Bowen-Margulis measure, denoted as μΦ\mu_{\Phi}. The Hilbert geodesic flow Φ\Phi is topologically mixing and Anosov, thus admits a unique measure of maximal entropy on T1XρT^{1}X_{\rho}. The entropy of the measure of maximal entropy coincides with the topological entropy. See for instance [32, Section 20].

The following lemma, derived from Abramov’s formula [1], allows us to rescale a reparametrization function to be pressure zero.

Lemma 2.6 (Sambarino [49, Lemma 2.4], Bowen-Ruelle [13, Proposition 3.1]).

For a positive Hölder reparametrization function ff on T1XρT^{1}X_{\rho} and hh\in\mathbb{R}, the pressure function satisfies

P(hf)=0P(-hf)=0

if and only if h=h(Φf)h=h(\Phi^{f}), where

h(Φf)=lim supT1Tlog#{τ periodic orbitλ(f,τ)T}.h(\Phi^{f})=\limsup_{T\to\infty}\frac{1}{T}\log\#\{\tau\text{ periodic orbit}\mid\lambda(f,\tau)\leq T\}.

By definition, h(Φf)h(\Phi^{f}) is the topological entropy of the reparametrized flow Φf\Phi^{f}.

We will use Lemma 2.6 in the proofs of Theorems 1.1, Theorem 1.3 and Theorem 1.5.

Remark 2.7.

By construction, if f=fρ1ρ2f=f^{\rho_{2}}_{\rho_{1}} is the reparametrization function defined in Lemma 2.3, then the topological entropy of the flow Φf\Phi^{f} is equal to the topological entropy h(ρ2)h(\rho_{2}) of ρ2\rho_{2} as defined in the introduction.

Finally, we introduce (Livšic) cohomology. We say two Hölder continuous functions ff and gg are (Livšic) cohomologous if there exists a Hölder continuous function V:T1XρV\colon T^{1}X_{\rho}\to\mathbb{R} that is differentiable in the flow’s direction such that

f(x)g(x)=t|t=0V(Φt(x)).f(x)-g(x)=\frac{\partial}{\partial t}\bigg{|}_{t=0}V(\Phi_{t}(x)).
Remark 2.8.
  1. (1)

    (Livšic’s Theorem, [38]) Two Hölder continuous functions ff and gg are cohomologous on T1XρT^{1}X_{\rho} if and only if λ(f,τ)=λ(g,τ)\lambda(f,\tau)=\lambda(g,\tau) for any periodic orbit τ\tau of Φ\Phi. It follows that the pressure of a Hölder continuous function depends only on its cohomology class.

  2. (2)

    Two Hölder continuous functions ff and gg have the same equilibrium state on T1XρT^{1}X_{\rho} if and only if fgf-g is cohomologous to a constant CC. In this case, we have P(f)=P(g)+CP(f)=P(g)+C [32, Section 20].

2.4. Geodesic currents for convex projective surfaces

Fix an auxiliary hyperbolic structure mm on SS. A geodesic current is a Borel, locally-finite, π1(S)\pi_{1}(S)-invariant measure on the set of complete geodesics of the universal cover S~\widetilde{S}. An important example is the geodesic current δγ\delta_{\gamma} given by Dirac measures on the axes of the lifts of a closed geodesic γ\gamma in SS.

The space 𝒞(S)\mathcal{C}(S) of geodesic currents is a convex cone in an infinite dimensional vector space. Bonahon [8] extended the intersection pairing on closed curves to the space of geodesic currents, i.e. there exists a positive, symmetric, bilinear pairing

i:𝒞(S)×𝒞(S)0i\colon\mathcal{C}(S)\times\mathcal{C}(S)\to\mathbb{R}_{\geq 0}

such that i(δc,δd)i(\delta_{c},\delta_{d}) equals the intersection number of the closed geodesics cc and dd.

Extending work of Bonahon [9], in [15, 42] it was shown that for each convex projective surface ρ\rho there exists a Hilbert geodesic current νρ\nu_{\rho} such that for every [γ][Γ][\gamma]\in[\Gamma]

i(νρ,δγ)=ρH([γ]),i(\nu_{\rho},\delta_{\gamma})=\ell_{\rho}^{H}([\gamma]),

where γ\gamma denotes the unique closed geodesic in its free homotopy class [γ][\gamma]. Bonahon [9, Proposition 15] proves that the geodesic current νρ\nu_{\rho} of a hyperbolic structure ρ\rho has self-intersection i(νρ,νρ)=π2χ(S)i(\nu_{\rho},\nu_{\rho})=-\pi^{2}\chi(S). On the other hand, if ρ(S)\rho\in\mathfrak{C}(S) is not in the Fuchsian locus, then i(νρ,νρ)>π2χ(S)i(\nu_{\rho},\nu_{\rho})>-\pi^{2}\chi(S) by Corollary 5.3 in [15].

In general, given a geodesic current ν\nu, we can use the intersection number to define its length spectrum ν:[Γ]+\ell_{\nu}\colon[\Gamma]\to\mathbb{R}^{+} as ν([γ])=i(ν,δγ)\ell_{\nu}([\gamma])=i(\nu,\delta_{\gamma}). The systole of ν\nu is then Sys(ν):=inf[γ][Γ]ν([γ])\mathrm{Sys}(\nu):=\inf_{[\gamma]\in[\Gamma]}\ell_{\nu}([\gamma]). Corollary 1.5 in [18] shows that Sys:𝒞(S)[0,)\mathrm{Sys}\colon\mathcal{C}(S)\to[0,\infty) is a continuous function.

A geodesic current is period minimizing if for all T>0T>0 the set #{[γ][Γ]ν([γ])<T}\#\{[\gamma]\in[\Gamma]\mid\ell_{\nu}([\gamma])<T\} is finite. We define the exponential growth rate of a period minimizing geodesic current ν\nu by

h(ν)=limT1Tlog#{[γ][Γ]ν([γ])<T}.h(\nu)=\lim_{T\to\infty}\frac{1}{T}\log\#\{[\gamma]\in[\Gamma]\mid\ell_{\nu}([\gamma])<T\}.

The notation is motivated by the fact that if ρ\rho is a convex projective structure and νρ\nu_{\rho} is the corresponding Hilbert geodesic current, then h(νρ)h(\nu_{\rho}) is equal to the topological entropy h(ρ)h(\rho) of ρ\rho.

The systole and the exponential growth rate of a geodesic current are related by the following inequality, which will play an important role in the proof of Theorem 1.3.

Theorem 2.9 (Corollary 7.6 in [42]).

Let SS be a closed, connected, oriented surface of genus g2g\geq 2. There exists a constant C>0C>0 depending only on gg such that for every period minimizing geodesic current ν𝒞(S)\nu\in\mathcal{C}(S)

Sys(ν)h(ν)C.\text{Sys}(\nu)h(\nu)\leq C.

2.5. Independence of convex projective surfaces

We start by recalling the notion of (topologically) weakly mixing flows which motivates the concept of independence of representations. A flow φ\varphi on T1XρT^{1}X_{\rho} is weakly mixing if its periods do not generate a discrete subgroup of \mathbb{R}. In particular, we can ask whether the Hilbert geodesic flow Φ\Phi is weakly mixing. Equivalently, we ask whether there exists some non-zero real number aa\in\mathbb{R} such that aρH(γ)a\ell_{\rho}^{H}(\gamma)\in\mathbb{Z} for all [γ]Γ[\gamma]\in\Gamma. In the proof of Theorem 1.1, we will need a strengthening of this property for a pair of representations.

Two convex projective structures ρ1\rho_{1} and ρ2\rho_{2} are dependent if there exist a1,a2a_{1},a_{2}\in\mathbb{R}, not both equal to zero, such that a1ρ1H([γ])+a2ρ2H([γ])a_{1}\ell^{H}_{\rho_{1}}([\gamma])+a_{2}\ell^{H}_{\rho_{2}}([\gamma])\in\mathbb{Z} for all [γ][Γ][\gamma]\in[\Gamma]. Otherwise, ρ1\rho_{1} and ρ2\rho_{2} are independent over \mathbb{Z}. The next definition clarifies that this notion of independence is of a dynamical nature.

Definition 2.10.

Two positive Hölder continuous functions f1,f2:T1Xρf_{1},f_{2}\colon T^{1}X_{\rho}\to\mathbb{R} are dependent if there exists a1,a2a_{1},a_{2}\in\mathbb{R} not both equal to zero and a complex valued C1C^{1} function u:T1XρS1u:T^{1}X_{\rho}\to S^{1} such that a1f1+a2f2=12πiuua_{1}f_{1}+a_{2}f_{2}=\frac{1}{2\pi i}\frac{u^{\prime}}{u}. Here, uu^{\prime} denotes the derivative of uu along the flow, i.e. t|t=0uΦt\frac{\partial}{\partial t}\big{|}_{t=0}u\circ\Phi_{t}. Otherwise, f1f_{1} and f2f_{2} are independent. In particular, ff is said to be (in)dependent if ff and the constant function g1g\equiv 1 are (in)dependent.

Remark 2.11.

The integral over a closed orbit of uu\frac{u^{\prime}}{u} is an integer multiple of 2πi2\pi i. Thus, by integrating along closed orbits and using Equation (2.1) and Remark 2.2, we see that if ρ1\rho_{1} and ρ2\rho_{2} are independent, then the reparametrization function fρ1ρ2f_{\rho_{1}}^{\rho_{2}} on T1Xρ1T^{1}X_{\rho_{1}} is independent.

We now prove that convex projective surfaces with distinct Hilbert length spectra are independent.

Lemma 2.12 (Independence lemma).

Let ρ1\rho_{1} and ρ2\rho_{2} be convex projective structures with distinct Hilbert length spectra. If there exist a1,a2a_{1},a_{2}\in\mathbb{R} such that a1ρ1H([γ])+a2ρ2H([γ])a_{1}\ell^{H}_{\rho_{1}}([\gamma])+a_{2}\ell^{H}_{\rho_{2}}([\gamma])\in\mathbb{Z} for all [γ][Γ][\gamma]\in[\Gamma], then a1=a2=0a_{1}=a_{2}=0.

Proof.

Our proof follows from combining results of Benoist and an argument of Glorieux from [25].

We prove this statement by contradiction. Consider the product representation η=ρ1×ρ2:ΓG1×G2\eta=\rho_{1}\times\rho_{2}\colon\Gamma\to\mathrm{G}_{1}\times\mathrm{G}_{2} where Gi\mathrm{G}_{i} denotes the Zariski closure of ρi\rho_{i}. Benoist [5, Thm 1.3] proved that Gi\mathrm{G}_{i} is PSL(3,)\mathrm{PSL}(3,\mathbb{R}) or isomoprhic to PSO(1,2)\mathrm{PSO}(1,2). Either way, this Zariski closure is connected and simple, so G1×G2\mathrm{G}_{1}\times\mathrm{G}_{2} is semi-simple. Choose a Cartan subspace of G1×G2\mathrm{G}_{1}\times\mathrm{G}_{2} such that

𝔞{(x,y)3×3x1+x2+x3=0=y1+y2+y3}\mathfrak{a}\subseteq\{(\vec{x},\vec{y})\in\mathbb{R}^{3}\times\mathbb{R}^{3}\mid x_{1}+x_{2}+x_{3}=0=y_{1}+y_{2}+y_{3}\}

and a positive Weyl chamber 𝔞+\mathfrak{a}^{+} contained in {(x,y)𝔞x1x2x3 and y1y2y3}\{(\vec{x},\vec{y})\in\mathfrak{a}\mid x_{1}\geq x_{2}\geq x_{3}\text{ and }y_{1}\geq y_{2}\geq y_{3}\}. Denote by λ:G1×G2𝔞+\lambda\colon\mathrm{G}_{1}\times\mathrm{G}_{2}\to\mathfrak{a}^{+} the corresponding Jordan projection and by ϕa1,a2H\phi^{H}_{a_{1},a_{2}} the non-zero linear functional ϕa1,a2H(x,y)=a1(x1x3)+a2(y1y3)\phi^{H}_{a_{1},a_{2}}(\vec{x},\vec{y})=a_{1}(x_{1}-x_{3})+a_{2}(y_{1}-y_{3}) so that ϕa1,a2H(λ(η(γ)))=a1ρ1H([γ])+a2ρ2H([γ])\phi^{H}_{a_{1},a_{2}}(\lambda(\eta(\gamma)))=a_{1}\ell^{H}_{\rho_{1}}([\gamma])+a_{2}\ell^{H}_{\rho_{2}}([\gamma]).

Denote by HG1×G2\mathrm{H}\subseteq\mathrm{G}_{1}\times\mathrm{G}_{2} the Zariski closure of η(Γ)\eta(\Gamma). As a first step, observe that HG1×G2\mathrm{H}\neq\mathrm{G}_{1}\times\mathrm{G}_{2}. Otherwise, the Proposition on page 2 of [6] implies directly that λ(η(Γ))\lambda(\eta(\Gamma)) is dense (in the standard topology) in 𝔞\mathfrak{a}. We obtain a contradiction as we assumed ϕa1,a2H(λ(η(γ)))\phi^{H}_{a_{1},a_{2}}(\lambda(\eta(\gamma)))\in\mathbb{Z} for all γΓ\gamma\in\Gamma and ϕa1,a2H\phi^{H}_{a_{1},a_{2}} is continuous.

Let πi:HGi\pi_{i}\colon\mathrm{H}\to\mathrm{G}_{i} for i=1,2i=1,2 denote the projection maps and note that πi(η(Γ))=ρi(Γ)\pi_{i}(\eta(\Gamma))=\rho_{i}(\Gamma). In particular, πi\pi_{i} is surjective since H\mathrm{H} and Gi\mathrm{G}_{i} are the Zariski closures of η(Γ)\eta(\Gamma) and ρi(Γ)\rho_{i}(\Gamma), respectively. Denote by N1=π21(id)\mathrm{N}_{1}=\pi_{2}^{-1}(\text{id}) (resp. N2=π11(id)\mathrm{N}_{2}=\pi_{1}^{-1}(\text{id})) the kernel of π2\pi_{2} (resp. π1\pi_{1}) which is naturally identified with a normal subgroup of G1\mathrm{G}_{1} (resp. G2\mathrm{G}_{2}) (note the indices in the definition of Ni\mathrm{N}_{i}). Then, Goursat’s lemma [29, Thm 5.5.1] states that the image of H\mathrm{H} in G1/N1×G2/N2\mathrm{G}_{1}/\mathrm{N}_{1}\times\mathrm{G}_{2}/\mathrm{N}_{2} is the graph of an isomorphism G1/N1G2/N2\mathrm{G}_{1}/\mathrm{N}_{1}\cong\mathrm{G}_{2}/\mathrm{N}_{2}. Since G1\mathrm{G}_{1} is simple, then N1={e}\mathrm{N}_{1}=\{e\} or G1\mathrm{G}_{1}.

Case 1: Suppose N1=G1\mathrm{N}_{1}=\mathrm{G}_{1}. Then N2=G2\mathrm{N}_{2}=\mathrm{G}_{2} and H\mathrm{H} is the direct product G1×G2\mathrm{G}_{1}\times\mathrm{G}_{2}, which is a contradiction.

Case 2: Suppose N1={e}\mathrm{N}_{1}=\{e\}. Since G2\mathrm{G}_{2} is simple, N2={e}\mathrm{N}_{2}=\{e\} and G1G2\mathrm{G}_{1}\cong\mathrm{G}_{2}. In other words, H\mathrm{H} is the graph of an automorphism ι:G1G2\iota\colon\mathrm{G}_{1}\to\mathrm{G}_{2}. This induces an automorphism of the corresponding Lie algebras and, by the classification of their outer automorphisms given in [28] (see also [14, Theorem 11.9]), we deduce that ρ2\rho_{2} is conjugated to either ρ1\rho_{1} or ρ1\rho_{1}^{\ast}, which contradicts our hypothesis. ∎

Observe that Lemma 2.12 readily implies that the geodesic flow of a convex projective structure is weakly mixing. Otherwise there exist aa\in\mathbb{R}, a0a\neq 0 such that aρH([γ])a\ell^{H}_{\rho}([\gamma])\in\mathbb{Z} which directly contradicts Lemma 2.12 with a1=aa_{1}=a, a2=0a_{2}=0, ρ1=ρ\rho_{1}=\rho and ρ2(S)\rho_{2}\in\mathfrak{C}(S) different from ρ1\rho_{1} and ρ1\rho_{1}^{\ast}.

3. The Correlation Theorem

In this section, we study the length spectra of two convex real projective structures simultaneously.

This idea appeared first in [51] for studying correlation of hyperbolic structures. We adapt their argument to the context of convex real projective structures. Theorem 3.1, which was proved independently by Lalley and Sharp with slightly different conditions, gives the asymptotic formula for the number of closed orbits of an Axiom A flow under constraints. Anosov flows are important examples of Axiom A flows and Theorem 3.1 will be a crucial ingredient for our proof of Theorem 1.1.

Fix a convex projective surface ρ\rho. Let f:T1Xρf\colon T^{1}X_{\rho}\to\mathbb{R} be a Hölder continuous function and consider the function tP(tf)t\to P(tf) for tt\in\mathbb{R}, where PP denotes the pressure. This function is real analytic and strictly convex in tt when ff is not cohomologous to a constant [44, Prop 4.12]. Its derivative satisfies

(3.1) P(tf):=ddtP(tf)=f𝑑μtf,P^{\prime}(tf):=\frac{d}{dt}P(tf)=\int fd\mu_{tf},

where μtf\mu_{tf} is the equilibrium state for tftf. We denote by J(f)J(f) the open interval of values P(tf)P^{\prime}(tf). If aJ(f)a\in J(f), we let tat_{a}\in\mathbb{R} be the unique real number for which P(taf)=fdμtaf=aP^{\prime}(t_{a}f)=\int f\mathrm{d}\mu_{t_{a}f}=a. We ease notation and set μa=μtaf\mu_{a}=\mu_{t_{a}f}.

The following is the key result needed to establish our Theorem 1.1 (and Theorem 1.7).

Theorem 3.1 (Lalley [36, Theorem I], Sharp [52, Theorem 1]).

Let f:T1Xρf:T^{1}X_{\rho}\to\mathbb{R} be an independent Hölder continuous function and let aJ(f)a\in J(f). Then, for fixed ε>0\varepsilon>0, there is a constant C=C(f,ε)C=C(f,\varepsilon) such that

#{τ:λ(τ)(x,x+ε),λ(f,τ)(ax,ax+ε)}Cexp(h(μa)x)x3/2.\#\{\tau:\lambda(\tau)\in(x,x+\varepsilon),\ \lambda(f,\tau)\in(ax,ax+\varepsilon)\}\sim C\frac{\exp(h(\mu_{a})x)}{x^{3/2}}.
Proof.

Recall that the Hilbert geodesic flow on T1XρT^{1}X_{\rho} is Anosov and, as pointed out at the end of section 2, weakly mixing. Thus, we can apply Lalley and Sharp’s results which hold for all weakly mixing Axiom A flows. ∎

Remark 3.2.

The constant C=C(f,ε)>0C=C(f,\varepsilon)>0 has the same order of magnitude as ε2\varepsilon^{2} and is related to P′′(taf)P^{\prime\prime}(t_{a}f). See [36, Thm 5] and [51].

We introduce the concepts of pressure intersection and renormalized pressure intersection which we will need for the proof of Theorem 1.1.

Definition 3.3.

Let ρ1\rho_{1} and ρ2\rho_{2} be two convex projective structures and let f:T1Xρ1f\colon T^{1}X_{\rho_{1}}\to\mathbb{R} be a Hölder continuous reparametrization function. The pressure intersection of ρ1\rho_{1} and ρ2\rho_{2} is

𝐈(ρ1,ρ2):=fdμΦρ1\mathbf{I}(\rho_{1},\rho_{2}):=\int f\mathrm{d}\mu_{\Phi^{\rho_{1}}}

where μΦρ1\mu_{\Phi^{\rho_{1}}} is the measure of maximal entropy for Φρ1\Phi^{\rho_{1}}. The renormalized pressure intersection of ρ1\rho_{1} and ρ2\rho_{2} is

𝐉(ρ1,ρ2):=h(ρ2)h(ρ1)𝐈(ρ1,ρ2).\mathbf{J}(\rho_{1},\rho_{2}):=\frac{h(\rho_{2})}{h(\rho_{1})}\mathbf{I}(\rho_{1},\rho_{2}).

By Livšic’s Theorem, the definitions of pressure intersection and renormalized pressure intersection do not depend on the choice of reparametrization function (see Remark 2.8).

Proposition 3.4 ([14, Proposition 3.8]).

For every ρ1,ρ2(S)\rho_{1},\rho_{2}\in\mathfrak{C}(S), the renormalized pressure intersection is such that

𝐉(ρ1,ρ2)1\mathbf{J}(\rho_{1},\rho_{2})\geq 1

with equality if and only if Lρ1H=Lρ2HL^{H}_{\rho_{1}}=L^{H}_{\rho_{2}}.

Remark 3.5.

Given two distinct elements ρ1,ρ2\rho_{1},\rho_{2} in 𝒯(S)\mathcal{T}(S), we can always find some free homotopy classes [γ1][\gamma_{1}] and [γ2][\gamma_{2}] so that ρ1([γ1])>ρ2([γ1])\ell_{\rho_{1}}([\gamma_{1}])>\ell_{\rho_{2}}([\gamma_{1}]) and ρ1([γ2])<ρ2([γ2])\ell_{\rho_{1}}([\gamma_{2}])<\ell_{\rho_{2}}([\gamma_{2}]). However, Tholozan [54, Theorem B] shows that there exist representations ρ\rho in (S)\mathfrak{C}(S) which dominate a Fuchsian representation j𝒯(S)j\in\mathcal{T}(S) in the sense that ρH([γ])j([γ])\ell^{H}_{\rho}([\gamma])\geq\ell_{j}([\gamma]) for all [γ]π1(S)[\gamma]\in\pi_{1}(S) (see also Section 5). However, Proposition 3.4 implies that whenever ρ1\rho_{1} and ρ2\rho_{2} in (S)\mathfrak{C}(S) have distinct renormalized Hilbert length spectra we can always find some free homotopy classes [γ1][\gamma_{1}] and [γ2][\gamma_{2}] so that Lρ1H([γ1])>Lρ2H([γ1])L^{H}_{\rho_{1}}([\gamma_{1}])>L^{H}_{\rho_{2}}([\gamma_{1}]) and Lρ1H([γ2])<Lρ2H([γ2])L^{H}_{\rho_{1}}([\gamma_{2}])<L^{H}_{\rho_{2}}([\gamma_{2}]). This motivates our focus on the renormalized Hilbert length as in the proof of the correlation theorem 1.1 below.

Now we are ready to prove our main theorem stated in the introduction and repeated below for the convenience of the reader.

Theorem 1.1. Fix a precision ε>0\varepsilon>0. Consider two convex projective structures ρ1\rho_{1} and ρ2\rho_{2} on a surface SS with distinct renormalized Hilbert length spectra Lρ1HLρ2HL_{\rho_{1}}^{H}\neq L_{\rho_{2}}^{H}. There exist constants C=C(ε,ρ1,ρ2)>0C=C(\varepsilon,\rho_{1},\rho_{2})>0 and M=M(ρ1,ρ2)(0,1)M=M(\rho_{1},\rho_{2})\in(0,1) such that

#{[γ][Γ]Lρ1H([γ])(x,x+h(ρ1)ε),Lρ2H([γ])(x,x+h(ρ2)ε)}CeMxx3/2\#\Big{\{}[\gamma]\in[\Gamma]\mid L_{\rho_{1}}^{H}([\gamma])\in\big{(}x,x+h(\rho_{1})\varepsilon\big{)},\ L_{\rho_{2}}^{H}([\gamma])\in\big{(}x,x+h(\rho_{2})\varepsilon\big{)}\Big{\}}\sim C\frac{e^{Mx}}{x^{3/2}}

where f(x)g(x)f(x)\sim g(x) means f(x)/g(x)1f(x)/g(x)\to 1 as xx\to\infty.

Proof.

This proof is inspired by the proof for hyperbolic surfaces in [51]. Our first goal is to show that, for the reparametrization function fρ1ρ2f_{\rho_{1}}^{\rho_{2}} described in Lemma 2.3, the value aa in Theorem 3.1 can be chosen to be h(ρ1)h(ρ2)\frac{h(\rho_{1})}{h(\rho_{2})}. In order to ease notations, we set f=fρ1ρ2f=f_{\rho_{1}}^{\rho_{2}} and we write Φ\Phi for the Hilbert geodesic flow Φρ1\Phi^{\rho_{1}} on T1Xρ1T^{1}X_{\rho_{1}}.

By Lemma 2.6, we have that

0=P(h(ρ2)f)=h(μh(ρ2)f)h(ρ2)fdμh(ρ2)f0=P(-h(\rho_{2})f)=h\big{(}\mu_{-h(\rho_{2})f}\big{)}-\int h(\rho_{2})f\mathrm{d}\mu_{-h(\rho_{2})f}

where μh(ρ2)f\mu_{-h(\rho_{2})f} is the equilibrium state of h(ρ2)f-h(\rho_{2})f. Hence

h(ρ2)fdμh(ρ2)f=h(μh(ρ2)f)h(\rho_{2})\int f\mathrm{d}\mu_{-h(\rho_{2})f}=h(\mu_{-h(\rho_{2})f})

and

(3.2) fdμh(ρ2)f=h(μh(ρ2)f)h(ρ2)h(ρ1)h(ρ2).\int f\mathrm{d}\mu_{-h(\rho_{2})f}=\frac{h(\mu_{-h(\rho_{2})f})}{h(\rho_{2})}\leq\frac{h(\rho_{1})}{h(\rho_{2})}.

Notice the equality can be attained only when μh(ρ2)f=μΦ\mu_{-h(\rho_{2})f}=\mu_{\Phi}, where μΦ\mu_{\Phi} is the measure of maximal entropy for the geodesic flow Φ\Phi. This happens only if h(ρ2)f-h(\rho_{2})f is cohomologous to a constant which, by integrating, implies the length spectrum ρ1H\ell_{\rho_{1}}^{H} is a multiple of ρ2H\ell_{\rho_{2}}^{H}. This is impossible by Lemma 2.12 and therefore the inequality is strict.

On the other hand, by Propositon 3.4, we have that

1𝐉(ρ1,ρ2)\displaystyle 1\leq\mathbf{J}(\rho_{1},\rho_{2}) =h(ρ2)h(ρ1)𝐈(ρ1,ρ2)=h(ρ2)h(ρ1)fdμΦ.\displaystyle=\frac{h(\rho_{2})}{h(\rho_{1})}\mathbf{I}(\rho_{1},\rho_{2})=\frac{h(\rho_{2})}{h(\rho_{1})}\int f\mathrm{d}\mu_{\Phi}.

where μΦ\mu_{\Phi} is the Bowen-Margulis measure of Φ\Phi. This yields

(3.3) fdμΦh(ρ1)h(ρ2).\int f\mathrm{d}\mu_{\Phi}\geq\frac{h(\rho_{1})}{h(\rho_{2})}.

Moreover, the equality is attained only when h(ρ2)fh(\rho_{2})f is cohomologous to h(ρ1)h(\rho_{1}) which is impossible given our hypotheses.

Combining the inequalities (3.2) and (3.3), together with the fact that J(f)J(f) is an open interval, we conclude that (h(ρ1)h(ρ2)δ,h(ρ1)h(ρ2)+δ)J(f)\left(\frac{h(\rho_{1})}{h(\rho_{2})}-\delta,\frac{h(\rho_{1})}{h(\rho_{2})}+\delta\right)\subset J(f) for small δ>0\delta>0. In particular, because P(tf)P^{\prime}(tf) is a strictly increasing continuous function, there exists some ta0(h(ρ2),0)t_{a_{0}}\in(-h(\rho_{2}),0) such that a0=P(ta0f)=h(ρ1)h(ρ2)a_{0}=P^{\prime}(t_{a_{0}}f)=\frac{h(\rho_{1})}{h(\rho_{2})}, as desired.

We now show that 0<h(μa0)<h(ρ1)0<h(\mu_{a_{0}})<h(\rho_{1}). Because the Hilbert geodesic flow is Anosov, it has positive entropy with respect to any equilibrium state of a Hölder reparametrization, so 0<h(μa0)h(μΦ)=h(ρ1)0<h(\mu_{a_{0}})\leq h(\mu_{\Phi})=h(\rho_{1}) and the equality occurs if and only if μa=μΦ\mu_{a}=\mu_{\Phi} is the measure of maximal entropy. But as shown above,

fdμΦ>h(ρ1)h(ρ2)=fdμa0.\int f\mathrm{d}\mu_{\Phi}>\frac{h(\rho_{1})}{h(\rho_{2})}=\int f\mathrm{d}\mu_{a_{0}}.

This shows that μa0\mu_{a_{0}} can not be the Bowen-Margulis measure and hence h(μa0)<h(ρ1)h(\mu_{a_{0}})<h(\rho_{1}).

Thanks to Lemma 2.12 and Theorem 3.1, we can conclude that for ρ1\rho_{1}, ρ2\rho_{2} convex projective structures such that ρ2ρ1,ρ1\rho_{2}\neq\rho_{1},\rho_{1}^{\ast}, there exists C~=C~(ρ1,ρ2,ε)\widetilde{C}=\widetilde{C}(\rho_{1},\rho_{2},\varepsilon) such that

#{[γ][Γ]:ρ1H([γ])(y,y+ε),ρ2H([γ])(h(ρ1)h(ρ2)y,h(ρ1)h(ρ2)y+ε)}C~exp(h(μa0)y)y3/2.\#\left\{[\gamma]\in[\Gamma]\colon\ell_{\rho_{1}}^{H}([\gamma])\in(y,y+\varepsilon),\ \ell_{\rho_{2}}^{H}([\gamma])\in\left(\frac{h(\rho_{1})}{h(\rho_{2})}y,\frac{h(\rho_{1})}{h(\rho_{2})}y+\varepsilon\right)\right\}\sim\widetilde{C}\frac{\exp(h(\mu_{a_{0}})y)}{y^{3/2}}.

Setting x=h(ρ1)yx=h(\rho_{1})y and clearing denominators, we have the desired statement with C=h(ρ1)3/2C~C=h(\rho_{1})^{3/2}\widetilde{C} and M=h(μa0)h(ρ1)(0,1)M=\frac{h(\mu_{a_{0}})}{h(\rho_{1})}\in{(0,1)}. ∎

4. Correlation number, Manhattan curve, and Decay of Correlation number

In this section we focus on the correlation number M(ρ1,ρ2)M(\rho_{1},\rho_{2}) from Theorem 1.1. In Theorem 4.2 we express the correlation number in terms of Burger’s Mahnattan curve [17] generalizing the main result in [53]. In §4.2 we prove that the correlation number is not uniformly bounded away from zero in (S)\mathfrak{C}(S). Specifically, we provide two sequences of hyperbolic structures along which the correlation number goes to zero, thus answering a question from [51].

4.1. Manhattan curve and Correlation number

Let ρ1\rho_{1} and ρ2\rho_{2} be two convex projective structures. The Manhattan curve of ρ1,ρ2\rho_{1},\rho_{2} is the curve 𝒞(ρ1,ρ2)\mathcal{C}(\rho_{1},\rho_{2}) that bounds the convex set

{(a,b)2:[γ][Γ]e(aρ1H([γ])+bρ2H([γ]))<+}\bigg{\{}(a,b)\in\mathbb{R}^{2}:\sum_{[\gamma]\in[\Gamma]}e^{-\big{(}a\ell^{H}_{\rho_{1}}([\gamma])+b\ell^{H}_{\rho_{2}}([\gamma])\big{)}}<+\infty\bigg{\}}

Equivalently (see [53]), the Manhattan curve can be defined in terms of the pressure function and the reparametrization function f=fρ1ρ2:T1Xρ1f=f_{\rho_{1}}^{\rho_{2}}:T^{1}X_{\rho_{1}}\to\mathbb{R} from Lemma 2.3 as

𝒞(ρ1,ρ2)={(a,b)2:P(abf)=0}={(a,b)2:P(bf)=a}.\mathcal{C}(\rho_{1},\rho_{2})=\bigg{\{}(a,b)\in\mathbb{R}^{2}\colon P(-a-bf)=0\bigg{\}}=\bigg{\{}(a,b)\in\mathbb{R}^{2}\colon P(-bf)=a\bigg{\}}.

The next theorem collects the properties of the Manhattan curve we will need which were first discussed in the setting of representations into isometry groups of rank one symmetric spaces (see [17, Theorem 1]).

Theorem 4.1.

Let ρ1,ρ2(S)\rho_{1},\rho_{2}\in\mathfrak{C}(S) denote two convex projective surfaces.

  1. (1)

    The Manhattan curve is a real analytic convex curve passing through the points (h(ρ1),0)(h(\rho_{1}),0) and (0,h(ρ2))(0,h(\rho_{2})).

  2. (2)

    The normals to the Manhattan curve at the points (h(ρ1),0)(h(\rho_{1}),0) and (0,h(ρ2))(0,h(\rho_{2})) have slopes 𝐈(ρ1,ρ2)\mathbf{I}(\rho_{1},\rho_{2}) and 1/𝐈(ρ2,ρ1)1/\mathbf{I}(\rho_{2},\rho_{1}), respectively.

  3. (3)

    The Manhattan curve is strictly convex if and only if ρ2ρ1,ρ1\rho_{2}\neq\rho_{1},\rho_{1}^{\ast}.

Proof.

Real analyticity and convexity of the Manhattan curve follow from its definition and real analyticity of the pressure function. The fact that 𝒞(ρ1,ρ2)\mathcal{C}(\rho_{1},\rho_{2}) passes through (h(ρ1),0)(h(\rho_{1}),0) and (0,h(ρ2))(0,h(\rho_{2})) follows from Lemma 2.6 and the definition of topological entropy. To see that the normal to the Manhattan curve at (h(ρ1),0)(h(\rho_{1}),0) has slope 𝐈(ρ1,ρ2){\mathbf{I}}(\rho_{1},\rho_{2}), observe that the Manhattan curve is the graph 𝒞(ρ1,ρ2)\mathcal{C}(\rho_{1},\rho_{2}) of a real analytic function qq such that P(q(s)f)=sP(-q(s)f)=s. Then, differentiate the equality P(q(s)f)=sP(-q(s)f)=s using equation (3.1). The rest of item (2) follows by symmetry. Item (3) follows from item (2), Theorem 2.12 and Proposition 3.4. ∎

Sharp [53] expressed the correlation number M(ρ1,ρ2)M(\rho_{1},\rho_{2}) in terms of the Manhattan curve for hyperbolic structures. We establish an analogous result for convex real projective structures.

Theorem 4.2.

Let ρ1\rho_{1} and ρ2\rho_{2} be convex projective structures with distinct renormalized Hilbert length spectra. Then, their correlation number can be written as

M(ρ1,ρ2)=ah(ρ1)+bh(ρ2)M(\rho_{1},\rho_{2})=\frac{a}{h(\rho_{1})}+\frac{b}{h(\rho_{2})}

where (a,b)𝒞(ρ1,ρ2)(a,b)\in\mathcal{C}(\rho_{1},\rho_{2}) is the point on the Manhattan curve at which the tangent line is parallel to the line passing through (h(ρ1),0)(h(\rho_{1}),0) and (0,h(ρ2))(0,h(\rho_{2})). See Figure 1.

Refer to caption

0aabbh(ρ2)h(\rho_{2})h(ρ1)h(\rho_{1})

Figure 1. The Manhattan curve and the point (a,b)(a,b) described in Theorem 4.2
Proof.

By the proof of Theorem 1.1, the correlation number is such that h(ρ1)M(ρ1,ρ2)=h(μa0)h(\rho_{1})M(\rho_{1},\rho_{2})=h(\mu_{a_{0}}), where a0=fdμa0=h(ρ1)h(ρ2)a_{0}=\int f\mathrm{d}\mu_{a_{0}}=\frac{h(\rho_{1})}{h(\rho_{2})}. By definition,

h(μa0)=P(ta0f)ta0h(ρ1)h(ρ2).h(\mu_{a_{0}})=P(t_{a_{0}}f)-t_{a_{0}}\frac{h(\rho_{1})}{h(\rho_{2})}.

Note that 𝒞(ρ1,ρ2)\mathcal{C}(\rho_{1},\rho_{2}) is the graph of a real analytic function qq defined implicitly as P(q(s)f)=sP(-q(s)f)=s. Setting q(s)=ta0q(s)=-t_{a_{0}}, it follows that

M(ρ1,ρ2)=h(μa0)h(ρ1)=P(q(s)f)h(ρ1)+q(s)h(ρ2)=sh(ρ1)+q(s)h(ρ2).\displaystyle M(\rho_{1},\rho_{2})=\frac{h(\mu_{a_{0}})}{h(\rho_{1})}=\frac{P(-q(s)f)}{h(\rho_{1})}+\frac{q(s)}{h(\rho_{2})}=\frac{s}{h(\rho_{1})}+\frac{q(s)}{h(\rho_{2})}.

Moreover, observe that

1=ddsP(q(s)f)=(fdμq(s)f)q(s).1=\frac{d}{ds}P(-q(s)f)=\left(-\int f\mathrm{d}\mu_{-q(s)f}\right)q^{\prime}(s).

We conclude by recalling that f𝑑μta0f=h(ρ1)h(ρ2)\int fd\mu_{t_{a_{0}}f}=\frac{h(\rho_{1})}{h(\rho_{2})} and that the line passing through (h(ρ1),0)(h(\rho_{1}),0) and (0,h(ρ2))(0,h(\rho_{2})) has slope h(ρ2)h(ρ1)-\frac{h(\rho_{2})}{h(\rho_{1})}. ∎

Remark 4.3.

It follows from strict convexity of the Manhattan curve and Theorem 4.2 that M(ρ1,ρ2)(0,1)M(\rho_{1},\rho_{2})\in(0,1). This fact is independently proved in Theorem 1.1.

4.2. Decay of correlation number

This section is dedicated to the proof of Theorem 1.3 from the introduction, which we restate here for the convenience of the reader.

Theorem 1.3. There exist sequences (ρn)n=1(\rho_{n})_{n=1}^{\infty} and (ηn)n=1(\eta_{n})_{n=1}^{\infty} in the Teichmüller space 𝒯(S)\mathcal{T}(S) such that the correlation numbers M(ρn,ηn)M(\rho_{n},\eta_{n}) satisfy

limnM(ρn,ηn)=0.\lim_{n\to\infty}M(\rho_{n},\eta_{n})=0.
Proof.

We construct two special families of hyperbolic structures ρn\rho_{n}, ηn\eta_{n} and consider their corresponding geodesic currents νρn\nu_{\rho_{n}}, νηn\nu_{\eta_{n}} as in section 2.4. Our proof proceeds in two steps. First, we take geodesic currents νρn+νηn\nu_{\rho_{n}}+\nu_{\eta_{n}} given by the sum of the currents of ρn\rho_{n} and ηn\eta_{n} and show that their exponential growth rates satisfy limnh(νρn+νηn)=0\lim\limits_{n\to\infty}h(\nu_{\rho_{n}}+\nu_{\eta_{n}})=0. Then we show that this condition implies that the correlation number M(ρn,ηn)M(\rho_{n},\eta_{n}) goes to zero as well.

We consider two filling pair-of-pants decomposition (αi)(\alpha_{i}) and (βi)(\beta_{i}) on a hyperbolic structure ρ0\rho_{0}. A family of simple closed curves is filling if the complement of their union consists of topological discs. We take (ρn)n=1(\rho_{n})_{n=1}^{\infty} to be a sequence of hyperbolic structures obtained by pinching all αi\alpha_{i} on ρ0\rho_{0} so that the hyperbolic length ρn(αi)=ϵn\ell_{\rho_{n}}(\alpha_{i})=\epsilon_{n} with ϵn0\epsilon_{n}\to 0 when nn\to\infty. Similarly, we take (ηn)(\eta_{n}) to be another sequence of hyperbolic structures obtained by pinching all βi\beta_{i} on ρ0\rho_{0} so that the hyperbolic length ηn(βi)=ϵn\ell_{\eta_{n}}(\beta_{i})=\epsilon_{n}. Note in such cases, we have that ρnH=ρn\ell^{H}_{\rho_{n}}=\ell_{\rho_{n}} and ηnH=ηn\ell^{H}_{\eta_{n}}=\ell_{\eta_{n}} and that the topological entropy h(ρn)=h(ηn)=1h(\rho_{n})=h(\eta_{n})=1 for all nn. We now proceed to prove limnh(νρn+νηn)=0\lim\limits_{n\to\infty}h(\nu_{\rho_{n}}+\nu_{\eta_{n}})=0.

By definition, the systole of νρn+νηn\nu_{\rho_{n}}+\nu_{\eta_{n}} is equal to Ln=inf[γ][Γ]{ρnH([γ])+ηnH([γ])}L_{n}=\inf\limits_{[\gamma]\in[\Gamma]}\left\{\ell^{H}_{\rho_{n}}([\gamma])+\ell^{H}_{\eta_{n}}([\gamma])\right\}. Note that for a fixed nn, #{[γ][Γ]ρnH([γ])+ηnH([γ])<T}<\#\{[\gamma]\in[\Gamma]\mid\ell^{H}_{\rho_{n}}([\gamma])+\ell^{H}_{\eta_{n}}([\gamma])<T\}<\infty for all T>0T>0. In other words, the geodesic current νρn+νηn\nu_{\rho_{n}}+\nu_{\eta_{n}} is period minimizing and so we can apply Theorem 2.9 to find a constant CC depending only on the topology of SS such that

h(νρn+νηn)CLn.h(\nu_{\rho_{n}}+\nu_{\eta_{n}})\leq\frac{C}{L_{n}}.

Therefore to show limnh(νρn+νηn)=0\lim\limits_{n\to\infty}h(\nu_{\rho_{n}}+\nu_{\eta_{n}})=0, it suffices to show that limnLn=\lim\limits_{n\to\infty}L_{n}=\infty. Because of the filling condition, the geodesic representative of any [γ][Γ][\gamma]\in[\Gamma] must intersect either curves in (αi)(\alpha_{i}) or curves in (βi)(\beta_{i}). By the Collar Lemma, each geodesic representative of αi\alpha_{i} (resp. βi\beta_{i}) for ρn\rho_{n} (resp. ηn\eta_{n}) is enclosed in a standard collar neighborhood of width approximately log(1ϵn)\log\Big{(}\frac{1}{\epsilon_{n}}\Big{)}. In particular, every closed curve traverses a collar neighborhood of width approximately log(1ϵn)\log\Big{(}\frac{1}{\epsilon_{n}}\Big{)} for the hyperbolic metric ρn\rho_{n} or the hyperbolic metric ηn\eta_{n} which implies that limnLn=\lim\limits_{n\to\infty}L_{n}=\infty.

We now show that limnh(νρn+νηn)=0\lim\limits_{n\to\infty}h(\nu_{\rho_{n}}+\nu_{\eta_{n}})=0 implies that the correlation number goes to zero as well. Consider the reparametrization functions fρnηnf^{\eta_{n}}_{\rho_{n}} as in Lemma 2.3 and notice that λ(1+fρnηn,τ)=ρnH([γ])+ηnH([γ])\lambda(1+f^{\eta_{n}}_{\rho_{n}},\tau)=\ell^{H}_{\rho_{n}}([\gamma])+\ell^{H}_{\eta_{n}}([\gamma]) for every periodic orbit corresponding to [γ][Γ][\gamma]\in[\Gamma]. By Lemma 2.6, for all nn

P(h(νρn+νηn)h(νρn+νηn)fρnηn)=P(h(νρn+νηn)(1+fρnηn))=0.P(-h(\nu_{\rho_{n}}+\nu_{\eta_{n}})-h(\nu_{\rho_{n}}+\nu_{\eta_{n}})f^{\eta_{n}}_{\rho_{n}})=P(-h(\nu_{\rho_{n}}+\nu_{\eta_{n}})(1+f^{\eta_{n}}_{\rho_{n}}))=0.

We have that (h(νρn+νηn),h(νρn+νηn))𝒞(ρn,ηn)(h(\nu_{\rho_{n}}+\nu_{\eta_{n}}),h(\nu_{\rho_{n}}+\nu_{\eta_{n}}))\in\mathcal{C}(\rho_{n},\eta_{n}) thanks to the characterization of the Manhattan curve in terms of the pressure function. If the line y+x=2h(νρn+νηn)y+x=2h(\nu_{\rho_{n}}+\nu_{\eta_{n}}) is tangent to the Manhattan curve 𝒞(ρn,ηn)\mathcal{C}(\rho_{n},\eta_{n}), then by Theorem 4.2, we obtain M(ρn,ηn)=2h(νρn+νηn)M(\rho_{n},\eta_{n})=2h(\nu_{\rho_{n}}+\nu_{\eta_{n}}). If not, then the line y+x=2h(νρn+νηn)y+x=2h(\nu_{\rho_{n}}+\nu_{\eta_{n}}) must intersect 𝒞(ρn,ηn)\mathcal{C}(\rho_{n},\eta_{n}) at two points. See Figure 2. Now by the mean value theorem and the convexity of the Manhattan curve, we have from Theorem 4.2 that there exists 0<an,bn<2h(νρn+νηn)0<a_{n},b_{n}<2h(\nu_{\rho_{n}}+\nu_{\eta_{n}}) such that

0<M(ρn,ηn)=an+bn<2h(νρn+νηn).0<M(\rho_{n},\eta_{n})=a_{n}+b_{n}<2h(\nu_{\rho_{n}}+\nu_{\eta_{n}}).

This immediately implies that M(ρn,ηn)M(\rho_{n},\eta_{n}) goes to zero as nn goes to infinity. ∎

Refer to caption

0h(νρn+νηn)h(\nu_{\rho_{n}}+\nu_{\eta_{n}})2h(νρn+νηn)2h(\nu_{\rho_{n}}+\nu_{\eta_{n}})2h(νρn+νηn)2h(\nu_{\rho_{n}}+\nu_{\eta_{n}})h(νρn+νηn)h(\nu_{\rho_{n}}+\nu_{\eta_{n}})1111

Figure 2. The Manhattan curve and the point (h(νρn+νηn),h(νρn+νηn))(h(\nu_{\rho_{n}}+\nu_{\eta_{n}}),h(\nu_{\rho_{n}}+\nu_{\eta_{n}})) described in the proof of Theorem 1.3

5. The renormalized Hilbert length along cubic rays

5.1. Cubic rays and affine spheres

Building on work of Hitchin [30], Labourie [35] and Loftin [40] have independently parametrized (S)\mathfrak{C}(S) as the vector bundle of holomorphic cubic differentials over 𝒯(S)\mathcal{T}(S). We briefly recall this parametrization since we will use it in this subsection to study the correlation number. The reader can refer to [35] and [40] for a detailed construction. This parametrization shows that the space of pairs formed by a complex structure JJ on SS and a JJ-holomorphic cubic differential qq is in one-to-one correspondence with the space of convex real projective structures on SS.

Because of the classical correspondence between hyperbolic structures and complex structures, we will sometimes blur the difference between a Riemann surface XX and a hyperbolic structure XX. We recall that a holomorphic cubic differential qq is a holomorphic section of KKKK\otimes K\otimes K where KK is the canonical line bundle of a Riemann surface XX. Locally, a holomorphic cubic differential can be written in complex coordinate charts as q=q(z)dzdzdzq=q(z)dz\otimes dz\otimes dz (often written as q(z)dz3q(z)dz^{3}) with q(z)q(z) holomorphic. The cubic differential qq is invariant under change of coordinates, meaning that if we pick a different coordinate chart ww, then q(w)dw3=q(z)dz3q(w)dw^{3}=q(z)dz^{3}. The hyperbolic metric σ\sigma that corresponds to the complex structure of XX induces a Hermitian metric ,\langle\cdot,\cdot\rangle on the space of holomorphic cubic differentials on XX. The L2L^{2}-norm of a cubic differential qq is defined as

qX=Xq,qdσ.\|q\|_{X}=\int_{X}\langle q,q\rangle\mathrm{d}\sigma.

The correspondence between the space of pairs of complex structures on SS and holomorphic cubic differentials on associated Riemann surfaces with the set of convex real projective structures on SS goes through a geometrical object invariant under affine transformations, called the affine sphere. We briefly describe the relation. Consider a pair (J,q)(J,q), where JJ is a complex structure and qq is a holomorphic cubic differential on the Riemann surface XJX_{J} associated to JJ. Then one can obtain an affine sphere, which is a hypersurface in 3\mathbb{R}^{3} that is invariant under affine transformations, by solving a second order elliptic PDE called Wang’s equation (see [40, Section 4]). The affine sphere can be projected to 2\mathbb{RP}^{2} so to obtain the developing image Ωρ\Omega_{\rho} of some convex projective structure ρ\rho. Conversely, given a convex projective structure ρ\rho and its developing image Ωρ\Omega_{\rho}, we can construct an affine sphere by solving a Monge-Ampére equation (see [40, Section 7]). The affine sphere will provide a complex structure JJ and a holomorphic cubic differential qq on XJX_{J}.

Since hyperbolic structures correspond to complex structures, the above identification provides a way to parametrize the space of convex projective structures (S)\mathfrak{C}(S) as the vector bundle of holomorphic cubic differentials over 𝒯(S)\mathcal{T}(S). In particular, we can fix a nonzero cubic differential qq and consider the associated family of representations (ρt)t(\rho_{t})_{\scriptscriptstyle t\in\mathbb{R}} in (S)\mathfrak{C}(S) parametrized by (tq)t(tq)_{t\in\mathbb{R}}. When t0t\geq 0, we call such a family a cubic ray. In this section, we study properties of the renormalized Hilbert length spectrum along cubic rays in (S)\mathfrak{C}(S).

For the reminder of this section, when there is no ambiguity, we ease notation by writing ht=h(ρt)h_{t}=h(\rho_{t}) for the topological entropy, and LtH=LρtHL^{H}_{t}=L^{H}_{\rho_{t}} for the renormalized Hilbert length spectrum of a cubic ray (ρt)t0(\rho_{t})_{t\geq 0}. With these notations, ρ0\rho_{0} is the hyperbolic structure associated to X0X_{0}. The entropy h0h_{0} equals 11 and the renormalized Hilbert length spectrum L0HL^{H}_{0} is the X0X_{0}-length spectrum.

The following observation is important and follows immediately from work of Tholozan [54].

Lemma 5.1.

Let (ρt)t0(\rho_{t})_{t\geq 0} be a cubic ray. There exists D>1D>1 and t00t_{0}\geq 0 such that, for tt0t\geq t_{0},

1DL0HLtHDL0H.\displaystyle\frac{1}{D}L^{H}_{0}\leq L^{H}_{t}\leq DL^{H}_{0}.
Proof.

Tholozan [54, Theorem 3.9] proves that there exists B>1B>1 and t00t_{0}\geq 0 such that, for tt0t\geq t_{0},

(5.1) 1Bt1/30HtHBt1/30H.\frac{1}{B}t^{1/3}\ell^{H}_{0}\leq\ell^{H}_{t}\leq Bt^{1/3}\ell^{H}_{0}.

In turn, this implies that

#{[γ]|0H([γ])1Bt1/3T}#{[γ]tH([γ])T}#{[γ]|0H([γ])Bt1/3T}.\#\left\{[\gamma]\ \Big{|}\ \ell^{H}_{0}([\gamma])\leq\frac{1}{Bt^{1/3}}T\right\}\leq\#\{[\gamma]\mid\ell^{H}_{t}([\gamma])\leq T\}\leq\#\left\{[\gamma]\ \Big{|}\ \ell^{H}_{0}([\gamma])\leq\frac{B}{t^{1/3}}T\right\}.

Since h0=1h_{0}=1, by Definition of topological entropy, the above inequalities imply that

(5.2) 1Bt1/3htBt1/3.\frac{1}{Bt^{1/3}}\leq h_{t}\leq\frac{B}{t^{1/3}}.

The result follows by taking D=B2D=B^{2} and combining the inequalities (5.1) and (5.2). ∎

The following example points out a problem with comparing un-renormalized length spectra and partially motivates our study of the renormalized Hilbert length spectrum. In particular, Equation (5.3) below should be compared to the correlation theorem.

Example 5.2.

Fix ϵ>0\epsilon>0. Let ρ0\rho_{0} be a Fuchsian representation in (S)\mathfrak{C}(S) and consider a cubic ray (ρt)t0(S)(\rho_{t})_{t\geq 0}\subset\mathfrak{C}(S). Let t00t_{0}\geq 0 be as in Lemma 5.1 and consider any tt0t\geq t_{0}. Then

(5.3) limx#{[γ][Γ]|0H([γ])(x,x+ε),tH([γ])(x,x+ε)}=0.\lim_{x\to\infty}\#\Big{\{}[\gamma]\in[\Gamma]\,\Big{|}\,\ell_{0}^{H}([\gamma])\in\big{(}x,x+\varepsilon\big{)},\ \ell_{t}^{H}([\gamma])\in\big{(}x,x+\varepsilon\big{)}\Big{\}}=0.
Proof.

For tt0t\geq t_{0}, consider δ>0\delta>0 small enough so that ht<1<Dδh_{t}<1<D-\delta, where DD is as in Lemma 5.1. There exists MM sufficiently large such that Dxx+εDδ\frac{Dx}{x+\varepsilon}\geq D-\delta for all x>Mx>M. Therefore, for all x>Mx>M, if x<0H([γ])<x+ϵx<\ell_{0}^{H}([\gamma])<x+\epsilon, then

tH([γ])Dht0H([γ])>DDδxx+ε.\displaystyle\ell^{H}_{t}([\gamma])\geq\frac{D}{h_{t}}\ell^{H}_{0}([\gamma])>\frac{D}{D-\delta}x\geq x+\varepsilon.

This shows that {[γ][Γ]|0H([γ])(x,x+ε),tH([γ])(x,x+ε)}=\Big{\{}[\gamma]\in[\Gamma]\,\Big{|}\,\ell_{0}^{H}([\gamma])\in\big{(}x,x+\varepsilon\big{)},\ \ell_{t}^{H}([\gamma])\in\big{(}x,x+\varepsilon\big{)}\Big{\}}=\emptyset for all x>Mx>M, which implies that equation (5.3) holds. ∎

5.2. Cubic rays and correlation numbers

In contrast with Theorem 1.3, we show some instances in which the correlation numbers M(ρt,ηt)M(\rho_{t},\eta_{t}) arising from two different cubic rays are uniformly bounded away from zero. We first introduce a convenient lemma that will be used in the proof of Theorems 1.4 and 1.5.

Lemma 5.3.

Suppose ρ\rho and η\eta are different convex real projective structures such that ρη\rho^{*}\neq\eta. Let υρ\upsilon_{\rho} and υη\upsilon_{\eta} be the renormalized Hilbert currents associated to ρ\rho and η\eta, respectively. For s[0,1]s\in[0,1], let usu_{s} denote the geodesic current sυρ+(1s)υηs\upsilon_{\rho}+(1-s)\upsilon_{\eta}. Then there exists a unique s0(0,1)s_{0}\in(0,1) such that

M(ρ,η)=h(us0).M(\rho,\eta)=h(u_{s_{0}}).
Proof.

Since ηρ,ρ\eta\neq\rho,\rho^{*}, the Manhattan curve 𝒞(ρ,η)\mathcal{C}(\rho,\eta) is strictly convex. The line xh(ρ)+yh(η)=1\frac{x}{h(\rho)}+\frac{y}{h(\eta)}=1 intersects the Manhattan curve in the first quadrant only at the points (h(ρ),0)(h(\rho),0) and (0,h(η))(0,h(\eta)). It follows that for every s[0,1]s\in[0,1] the straight line connecting (sh(ρ),(1s)h(η))(sh(\rho),(1-s)h(\eta)) and the origin must intersect 𝒞(ρ,η)\mathcal{C}(\rho,\eta). By Lemma 2.6,

P(h(us)sh(ρ)h(us)(1s)h(η)fρη)=P(h(us)(sh(ρ)+(1s)h(η)fρη))=0,P(-h(u_{s})sh(\rho)-h(u_{s})(1-s)h(\eta)f^{\eta}_{\rho})=P(-h(u_{s})(sh(\rho)+(1-s)h(\eta)f^{\eta}_{\rho}))=0,

where fρηf^{\eta}_{\rho} is the reparametrization function from Lemma 2.3. It follows from the definition of the Manhattan curve that every point on 𝒞(ρ,η)\mathcal{C}(\rho,\eta) in the first quadrant can be written in the form of (h(us)sh(ρ),h(us)(1s)h(η))(h(u_{s})sh(\rho),h(u_{s})(1-s)h(\eta)) for some s[0,1]s\in[0,1]. By Theorem 4.2, there exists s0(0,1)s_{0}\in(0,1) such that

M(ρ,η)=s0h(us0)h(ρ)h(ρ)+(1s0)h(us0)h(η)h(η)=h(us0).M(\rho,\eta)=\frac{s_{0}h(u_{s_{0}})h(\rho)}{h(\rho)}+\frac{(1-s_{0})h(u_{s_{0}})h(\eta)}{h(\eta)}=h(u_{s_{0}}).\qed

We first study the case of two cubic rays which lie in two different fibers of the vector bundle (S)𝒯(S)\mathfrak{C}(S)\to\mathcal{T}(S) given by the Labourie-Loftin parametrization of the space of convex projective surfaces.

Theorem 1.4. Let (ρt)t0(\rho_{t})_{t\geq 0}, (ηr)r0(\eta_{r})_{r\geq 0} be two cubic rays associated to two different hyperbolic structures ρ0η0\rho_{0}\neq\eta_{0}. Then, there exists a constant C>0C>0 such that for all t,r0t,r\geq 0

M(ρt,ηr)CM(ρ0,η0).M(\rho_{t},\eta_{r})\geq CM(\rho_{0},\eta_{0}).
Proof.

Fix t,r0t,r\geq 0 and write ρ=ρt\rho=\rho_{t} and η=ηr\eta=\eta_{r}, for simplicity. Note that by hypothesis LρHLηHL^{H}_{\rho}\neq L^{H}_{\eta} and Lρ0HLη0HL^{H}_{\rho_{0}}\neq L^{H}_{\eta_{0}}. For s[0,1]s\in[0,1], let usu_{s} denote the geodesic current given by sυρ+(1s)υηs\upsilon_{\rho}+(1-s)\upsilon_{\eta} where υρ\upsilon_{\rho} and υη\upsilon_{\eta} are the renormalized Hilbert geodesic currents of ρ\rho and η\eta, respectively. Denote by h(us)h(u_{s}) the exponential growth rate for the geodesic current usu_{s}.

By Lemma 5.1, there exist constants D1,D2D_{1},D_{2} depending on ρ0\rho_{0} and η0\eta_{0}, respectively such that

sLρH+(1s)LηHsD1Lρ0H+(1s)D2Lη0Hmax{D1,D2}(sLρ0H+(1s)Lη0H).sL^{H}_{\rho}+(1-s)L^{H}_{\eta}\leq sD_{1}L^{H}_{\rho_{0}}+(1-s)D_{2}L^{H}_{\eta_{0}}\leq\max\{D_{1},D_{2}\}(sL^{H}_{\rho_{0}}+(1-s)L^{H}_{\eta_{0}}).

Set C=1max{D1,D2}C=\frac{1}{\max\{D_{1},D_{2}\}} and ws=sυρ0+(1s)υη0w_{s}=s\upsilon_{\rho_{0}}+(1-s)\upsilon_{\eta_{0}}, where υρ0\upsilon_{\rho_{0}} and υη0\upsilon_{\eta_{0}} are the renormalized Hilbert geodesic currents of ρ0\rho_{0} and η0\eta_{0}, respectively. Then,

h(us)Ch(ws).\displaystyle h(u_{s})\geq Ch(w_{s}).

By Lemma 2.6, the intersection between the Manhattan curve 𝒞(ρ0,η0)\mathcal{C}(\rho_{0},\eta_{0}) and the line passing through the origin and the point (s,(1s))(s,(1-s)) has coordinates (sh(ws),(1s)h(ws))(sh(w_{s}),(1-s)h(w_{s})) and lies on the line y+x=h(ws)y+x=h(w_{s}). Then, by Theorem 4.2,

h(ws)M(ρ0,η0).h(w_{s})\geq M(\rho_{0},\eta_{0}).

See Figure 3. Finally, by Lemma 5.3 there exists s0(0,1)s_{0}\in(0,1) such that

M(ρ,η)=h(us0)h(ws0)CM(ρ0,η0)M(\rho,\eta)=h(u_{s_{0}})\geq h(w_{s_{0}})\geq CM(\rho_{0},\eta_{0})

which concludes the proof. ∎

Refer to caption

01111sh(ws)sh(w_{s})(1s)h(ws)(1-s)h(w_{s})y+x=M(ρ0,η0)y+x=M(\rho_{0},\eta_{0})y+x=h(ws)y+x=h(w_{s})

Figure 3. The correlation number of ρ0\rho_{0} and η0\eta_{0} is less or equal to the exponential growth rate of the geodesic current wsw_{s}

Finally, we prove Theorem 1.5 from the introduction.

Theorem 1.5. Let ρt\rho_{t} and ηt\eta_{t} be two cubic rays associated to two different holomorphic cubic differentials q1q_{1} and q2q_{2} on a hyperbolic structure X0X_{0} such that q1,q2q_{1},q_{2} have unit L2L^{2}-norm with respect to X0X_{0} and q1q2q_{1}\neq-q_{2}. Then the correlation number M(ρt,ηt)M(\rho_{t},\eta_{t}) is uniformly bounded away from zero as tt goes to infinity.

Proof.

We note that ρ0=η0\rho_{0}=\eta_{0} and Lρ0H=Lη0H=L0HL_{\rho_{0}}^{H}=L_{\eta_{0}}^{H}=L_{0}^{H}. Recall from Lemma 5.1, there exists D1,D2>1D_{1},D_{2}>1 such that for tt large,

1D1L0HLρtHD1L0H and 1D2L0HLηtHD2L0H.\displaystyle\frac{1}{D_{1}}L^{H}_{0}\leq L^{H}_{\rho_{t}}\leq D_{1}L^{H}_{0}\qquad\text{ and }\qquad\frac{1}{D_{2}}L^{H}_{0}\leq L^{H}_{\eta_{t}}\leq D_{2}L^{H}_{0}.

Consider the renormalized Hilbert currents υρt\upsilon_{\rho_{t}} and υηt\upsilon_{\eta_{t}} for ρt\rho_{t} and ηt\eta_{t}, respectively. Let s[0,1]s\in[0,1] and set the geodesic current ut,s=sυρt+(1s)υηtu_{t,s}=s\upsilon_{\rho_{t}}+(1-s)\upsilon_{\eta_{t}}. As a first step, we show that the entropy of the geodesic current ut,s=sυρt+(1s)υηtu_{t,s}=s\upsilon_{\rho_{t}}+(1-s)\upsilon_{\eta_{t}} is uniformly bounded away from zero. We have from Lemma 5.1 for tt large, for all [γ][Γ][\gamma]\in[\Gamma],

sLρtH([γ])+(1s)LηtH([γ])(sD1+(1s)D2)L0H([γ])max{D1,D2}L0H([γ]).sL_{\rho_{t}}^{H}([\gamma])+(1-s)L^{H}_{\eta_{t}}([\gamma])\leq(sD_{1}+(1-s)D_{2})L_{0}^{H}([\gamma])\leq\max\{D_{1},D_{2}\}L^{H}_{0}([\gamma]).

Thus, we obtain h(ut,s)1Dh(u_{t,s})\geq\frac{1}{D} where D=max{D1,D2}D=\max\{D_{1},D_{2}\}.

Next we want to show this implies the correlation number M(ρt,ηt)M(\rho_{t},\eta_{t}) is bounded away from zero. Because q2q1q_{2}\neq-q_{1}, we know ρtηt\rho_{t}^{*}\neq\eta_{t} for t>0t>0 (see [39, section 5 and section 8]). For each t>0t>0, by Lemma 5.3, we can find some s(0,1)s\in(0,1) such that

M(ρt,ηt)=h(ut,s).M(\rho_{t},\eta_{t})=h(u_{t,s}).

We conclude that M(ρt,ηt)1DM(\rho_{t},\eta_{t})\geq\frac{1}{D} for tt large, as desired. ∎

5.3. Cubic rays and geodesic currents

In this section we observe two properties of renormalized Hilbert currents υρt=htνρt\upsilon_{\rho_{t}}=h_{t}\nu_{\rho_{t}} along cubic rays (ρt)t0(\rho_{t})_{t\geq 0} which readily follow from Lemma 5.1. Let us start by showing that the self-intersection of the renormalized Hilbert geodesic current is uniformly bounded along cubic rays.

Proposition 5.4.

There exists C>1C>1 such that for all tt\in\mathbb{R}

1Ci(υρt,υρt)C\frac{1}{C}\leq i(\upsilon_{\rho_{t}},\upsilon_{\rho_{t}})\leq C

where υρt\upsilon_{\rho_{t}} is the renormalized Hilbert geodesic current of ρt\rho_{t}.

Proof.

Recall from Section 2.4 that Bonahon proved that i(υρ0,υρ0)=π2χ(S)i(\upsilon_{\rho_{0}},\upsilon_{\rho_{0}})=-\pi^{2}\chi(S), where χ(S)\chi(S) is the Euler characteristic of SS. Corollary 5.2 in [15] states that

(inf[γ][Γ]LtH([γ])L0H([γ]))2i(υρt,υρt)i(υρ0,υρ0)(sup[γ][Γ]LtH([γ])L0H([γ]))2\left(\inf_{[\gamma]\in[\Gamma]}\frac{L^{H}_{t}([\gamma])}{L^{H}_{0}([\gamma])}\right)^{2}\leq\frac{i(\upsilon_{\rho_{t}},\upsilon_{\rho_{t}})}{i(\upsilon_{\rho_{0}},\upsilon_{\rho_{0}})}\leq\left(\sup_{[\gamma]\in[\Gamma]}\frac{L^{H}_{t}([\gamma])}{L^{H}_{0}([\gamma])}\right)^{2}

so

π2χ(S)(inf[γ][Γ]LtH([γ])L0H([γ]))2i(υρt,υρt)π2χ(S)(sup[γ][Γ]LtH([γ])L0H([γ]))2-\pi^{2}\chi(S)\left(\inf_{[\gamma]\in[\Gamma]}\frac{L^{H}_{t}([\gamma])}{L^{H}_{0}([\gamma])}\right)^{2}\leq i(\upsilon_{\rho_{t}},\upsilon_{\rho_{t}})\leq-\pi^{2}\chi(S)\left(\sup_{[\gamma]\in[\Gamma]}\frac{L^{H}_{t}([\gamma])}{L^{H}_{0}([\gamma])}\right)^{2}

We conclude by applying Lemma 5.1 for tt large. ∎

Bonahon [9] showed that Thurston’s compactification of the Teichmüller space can be understood via geodesic currents. Explicitly, given a diverging sequence of hyperbolic structures mtm_{t}, there exists λt>0\lambda_{t}>0 and a geodesic current α\alpha with i(α,α)=0i(\alpha,\alpha)=0, a measured lamination, such that λtνmt\lambda_{t}\nu_{m_{t}} converges up to subsequences to α\alpha. In this case, the systole of α\alpha vanishes, i.e.

Sys(α)=inf[γ][Γ]i(α,δγ)=0.\mathrm{Sys}(\alpha)=\inf_{[\gamma]\in[\Gamma]}i(\alpha,\delta_{\gamma})=0.

Burger, Iozzi, Parreau, and Pozzetti [18] show that there exist diverging sequences of Hilbert geodesic currents which converge projectively to a current α\alpha with Sys(α)>0\mathrm{Sys}(\alpha)>0. We use Lemma 5.1 to show that this happens along cubic rays.

Theorem 1.6. As tt goes to infinity, the renormalized Hilbert geodesic current υρt\upsilon_{\rho_{t}} along a cubic ray converges, up to passing to a subsequence, to a geodesic current υ\upsilon with Sys(υ)>0\mathrm{Sys}(\upsilon)>0.

Proof.

Suppose (αi)(\alpha_{i}) and (βj)(\beta_{j}) are filling pair-of-pants decompositions, i.e. they are pants decompositions such that the complement of their union is a collection of topological discs. Set u=γ(αi)(βj)δγu=\sum_{\gamma\in(\alpha_{i})\cup(\beta_{j})}\delta_{\gamma} and let M=maxγ(αi)(βj)L0H([γ])M=\max_{\gamma\in(\alpha_{i})\cup(\beta_{j})}L^{H}_{0}([\gamma]). For every tt\in\mathbb{R}, we have

i(υρt,u)=γ(αi)(βj)LtH([γ]).i(\upsilon_{\rho_{t}},u)=\sum_{\gamma\in(\alpha_{i})\cup(\beta_{j})}L^{H}_{t}([\gamma]).

By Lemma 5.1 for tt large υt\upsilon_{t} lies in {ν𝒞(S)i(ν,u)(6g6)DM}\left\{\nu\in\mathcal{C}(S)\mid i\left(\nu,u\right)\leq(6g-6)DM\right\} for some D>1D>1. This set is compact by [9, Proposition 4] and by linearity of the intersection number. Thus, υt\upsilon_{t} converges, up to passing to a subsequence, to υ𝒞(S)\upsilon\in\mathcal{C}(S). Applying Lemma 5.1 again, we see that for all tt large Sys(υρt)\mathrm{Sys}(\upsilon_{\rho_{t}}) is greater or equal to D1Sys(υρ0)>0D^{-1}\mathrm{Sys}(\upsilon_{\rho_{0}})>0. By continuity, the systole Sys(υ)\mathrm{Sys}(\upsilon) is strictly positive. ∎

6. Generalization to Hitchin representations

In this section, we illustrate how to generalize the main results of sections 3 and 4 to the context of Hitchin representations.

We start with introducing Hitchin components and Hitchin representations. Given ρ𝒯(S)\rho\in\mathcal{T}(S), we can postcompose the corresponding holonomy representation ρ:ΓPSL(2,)\rho\colon\Gamma\to\mathrm{PSL}(2,\mathbb{R}) with the unique (up to conjugation) irreducible representation i:PSL(2,)PSL(d,)i\colon\mathrm{PSL}(2,\mathbb{R})\to\mathrm{PSL}(d,\mathbb{R}) given by the action of APSL(2,)A\in\mathrm{PSL}(2,\mathbb{R}) on the space of degree d1d-1 homogeneous polynomials in two variables. The Hitchin component d(S)\mathcal{H}_{d}(S) is the connected component of the character variety 𝖧𝗈𝗆(Γ,PSL(d,))//PSL(d,)\mathsf{Hom}(\Gamma,\mathrm{PSL}(d,\mathbb{R}))/\!/\mathrm{PSL}(d,\mathbb{R}) containing iρi\circ\rho. The copy of the Teichmüller space 𝒯(S)\mathcal{T}(S) embedded in the Hitchin component is its Fuchsian locus. Hitchin proves in [30], using Higgs bundles techniques, that d(S)\mathcal{H}_{d}(S) is homeomorphic to an open Euclidean ball of dimension (2g2)dim(PSL(d,))(2g-2)\cdot\dim(\mathrm{PSL}(d,\mathbb{R})). When d=2d=2, the Hitchin component 2(S)\mathcal{H}_{2}(S) coincides with the Teichmüller space 𝒯(S)\mathcal{T}(S). Choi and Goldman [19] identify the Hitchin component 3(S)\mathcal{H}_{3}(S) with the space of convex projective structures on the surface SS, which is the main focus of Sections 2 through 5.

In this section, we establish the correlation theorem 1.7 for pairs of Hitchin representations for any d3d\geq 3. In this setting, the length spectra will not be defined geometrically as for the Hilbert length spectrum of a convex projective structure, but they will be interpreted as periods of the Busemann cocycle, which was first introduced by Quint [48]. Then, we follow a construction of Sambarino [49, 50] to replace the geodesic flow on the unit tangent bundle of a convex projective surface with Sambarino’s translation flows. These flows are not necessarily Anosov, but fit in more general framework of metric Anosov flows. Nevertheless, the main results coming from Thermodynamic formalism needed for this paper still hold in this setting ([45, Section 3]).

6.1. Length functions, Busemann cocycles and entropy

In this section, we define length functions for Hitchin representations and the Busemann cocycles associated to them. We refer to [48, 49] for a more detailed construction.

We need to recall some Lie theory. Let G=PSL(d,)\mathrm{G}=\mathrm{PSL}(d,\mathbb{R}) and consider the standard choices of Cartan subspace

𝔞={xdx1++xd=0}\mathfrak{a}=\{\vec{x}\in\mathbb{R}^{d}\mid x_{1}+\dots+x_{d}=0\}

and positive Weyl chamber 𝔞+={x𝔞x1xd}\mathfrak{a}^{+}=\{\vec{x}\in\mathfrak{a}\mid x_{1}\geq\dots\geq x_{d}\}. Let λ:G𝔞+\lambda\colon\mathrm{G}\to\mathfrak{a}^{+} be the Jordan projection

λ(g)=(logλ1(g),,logλd(g))\lambda(g)=(\log\lambda_{1}(g),\dots,\log\lambda_{d}(g))

consisting of the logarithms of the moduli of the eigenvalues of gg in nonincreasing order. We will use the following fundamental property of the Jordan projection of the image of a Hitchin representation.

Theorem 6.1 (Fock-Goncharov [23], Labourie [34]).

Every Hitchin representation ρd(S)\rho\in\mathcal{H}_{d}(S) is discrete and faithful and for every [γ][Γ][\gamma]\in[\Gamma],

λ1(ρ(γ))>>λd(ρ(γ))>0.\lambda_{1}(\rho(\gamma))>\dots>\lambda_{d}(\rho(\gamma))>0.

The limit cone ρ\mathcal{L}_{\rho} of a Hitchin representation ρd(S)\rho\in\mathcal{H}_{d}(S), introduced in [6], is the closed cone of 𝔞+\mathfrak{a}^{+} generated by λ(ρ(γ))\lambda(\rho(\gamma)). This cone contains all the rays spanned by positive multiples of λ(ρ(γ))\lambda(\rho(\gamma)) for all γΓ\gamma\in\Gamma. The (positive) dual cone of ρ\mathcal{L}_{\rho} is defined as ρ={ϕ𝔞:ϕ|ρ0}\mathcal{L}_{\rho}^{*}=\{\phi\in\mathfrak{a}^{*}:\phi|_{\mathcal{L}_{\rho}}\geq 0\}. For every i=1,,d1i=1,\dots,d-1, the ii-th simple root αi:𝔞\alpha_{i}\colon\mathfrak{a}\to\mathbb{R}, defined by αi(x)=xixi+1\alpha_{i}(\vec{x})=x_{i}-x_{i+1}, is an important example of element in the interior of ρ\mathcal{L}_{\rho}^{*}. We will focus on linear functionals in the set

Δ={c1α1++cd1αd1ci0,ici>0}\Delta=\left\{c_{1}\alpha_{1}+\dots+c_{d-1}\alpha_{d-1}\mid c_{i}\geq 0,\ \sum_{i}c_{i}>0\right\}

which is contained in the interior of ρ\mathcal{L}_{\rho}^{*} by Theorem 6.1.

Given ϕρ\phi\in\mathcal{L}_{\rho}^{*}, the ϕ\phi-length of [γ][Γ][\gamma]\in[\Gamma] is defined as

ρϕ([γ])=ϕ(λ(ρ(γ)))>0,\ell^{\phi}_{\rho}([\gamma])=\phi(\lambda(\rho(\gamma)))>0,

and its exponential growth rate is

hϕ(ρ)=lim supT1T#log{[γ][Γ]ρϕ([γ])T}.h^{\phi}(\rho)={\limsup_{T\to\infty}\frac{1}{T}\#\log\{[\gamma]\in[\Gamma]\mid\ell^{\phi}_{\rho}([\gamma])\leq T\}.}

An important property of linear functionals in Δ\Delta is the following.

Lemma 6.2.

If ϕΔ\phi\in\Delta, then hϕ(ρ)(0,)h^{\phi}(\rho)\in(0,\infty).

Proof.

Since we know hαi(ρ)=1h^{\alpha_{i}}(\rho)=1 from [44, Theorem B], by Lemma 2.7 in [47], a linear functional φρ\varphi\in\mathcal{L}_{\rho}^{*} has finite and positive entropy if and only if it belongs to the interior of ρ\mathcal{L}_{\rho}^{*}. Since elements in Δ\Delta are contained in the interior of ρ\mathcal{L}_{\rho}^{*}, the entropy hϕ(ρ)h^{\phi}(\rho) is positive and finite for every ϕΔ\phi\in\Delta. ∎

The ϕ\phi-length function can be realized via the period of the ϕ\phi-Busemann cocycle. To define ϕ\phi-Busemann cocycles, we need to introduce the Frenet curve ξρ\xi_{\rho} for a Hitchin representation ρ\rho. Denote by 𝖥d\mathsf{F}_{d} the space of complete flags in d\mathbb{R}^{d}. Fock and Goncharov [23] and Labourie [34] show that for every Hitchin representation ρ\rho, there exists a unique (up to conjugation) ρ\rho-equivariant bi-Hölder continuous Frenet curve ξρ:Γ𝖥d\xi_{\rho}\colon\partial\Gamma\to\mathsf{F}_{d} which is transverse, positive, and it satisfies certain contraction properties. Moreover, the Frenet curve is dynamics preserving: if γ+Γ\gamma^{+}\in\partial\Gamma is the attracting fixed point of γΓ\gamma\in\Gamma, then ξρ(γ+)\xi_{\rho}(\gamma^{+}) is the attracting eigenflag of ρ(γ)\rho(\gamma). We refer to [23, Thm 1.15] and [34, Thm 4.1] for precise statements.

We are now ready to define ϕ\phi-Busemann cocycles for ϕΔ\phi\in\Delta. Fix F0𝖥dF_{0}\in\mathsf{F}_{d} and observe that for every F𝖥dF\in\mathsf{F}_{d}, there exists kSO(d)k\in\mathrm{SO}(d) such that F=kF0F=kF_{0}. Every MSL(d,)M\in\mathrm{SL}(d,\mathbb{R}) can be written as M=L(expσ(M))UM=L(\exp{\sigma(M)})U, for some LSO(d)L\in\mathrm{SO}(d), σ(M)𝔞\sigma(M)\in\mathfrak{a} and UU is unipotent and upper triangular. This is known as the Iwasawa decomposition of MM. The Iwasawa cocycle B:SL(d,)×𝖥d𝔞B\colon\mathrm{SL}(d,\mathbb{R})\times\mathsf{F}_{d}\to\mathfrak{a} is defined as B(A,F)=σ(Ak)B(A,F)=\sigma(Ak).

The vector valued Busemann cocycle Bρ:Γ×Γ𝔞B_{\rho}\colon\Gamma\times\partial\Gamma\to\mathfrak{a} of a Hitchin representation ρ\rho with Frenet curve ξρ\xi_{\rho} is

Bρ(γ,x)=B(ρ(γ),ξρ(x)).B_{\rho}(\gamma,x)=B(\rho(\gamma),\xi_{\rho}(x)).

For every linear functional ϕΔ\phi\in\Delta we set Bρϕ(γ,x)=ϕ(Bρ(γ,x))B^{\phi}_{\rho}(\gamma,x)=\phi(B_{\rho}(\gamma,x)). Lemma 7.5 in [49] directly shows that the cocycle BρϕB^{\phi}_{\rho} encodes the ϕ\phi-length via the equality Bρϕ(γ,γ+)=ρϕ([γ])B_{\rho}^{\phi}(\gamma,\gamma^{+})=\ell^{\phi}_{\rho}([\gamma]), for every γΓ\gamma\in\Gamma.

The ϕ\phi-Busemann cocycle is an example of a Hölder cocycle.

Definition 6.3.

A Hölder cocycle is a map c:Γ×Γc\colon\Gamma\times\partial\Gamma\to\mathbb{R} such that

c(γη,x)=c(γ,ηx)+c(η,x)c(\gamma\eta,x)=c(\gamma,\eta x)+c(\eta,x)

for any γ,ηΓ\gamma,\eta\in\Gamma and xΓx\in\partial\Gamma and there exists α(0,1)\alpha\in(0,1) such that c(γ,)c(\gamma,\cdot) is α\alpha-Hölder for all γΓ\gamma\in\Gamma.

For a general Hölder cocycle, let c(γ)=c(γ,γ+)\ell_{c}(\gamma)=c(\gamma,\gamma^{+}) denote the cc-length (also known as period) of γΓ\gamma\in\Gamma. Two Hölder cocycles are cohomologous if they have the same periods. We can define the exponential growth rate hch_{c} of a general Hölder cocycle c:Γ×Γc:\Gamma\times\partial\Gamma\to\mathbb{R} as

hc=lim supT1T#log{[γ][Γ]c(γ)T}.h_{c}=\limsup_{T\to\infty}\frac{1}{T}\#\log\{[\gamma]\in[\Gamma]\mid\ell_{c}(\gamma)\leq T\}.

Note that the exponential growth rate hϕ(ρ)h^{\phi}(\rho) of the linear functional ϕΔ\phi\in\Delta is the same as the exponential growth rate hBρϕh_{B^{\phi}_{\rho}} of the cocycle BρϕB^{\phi}_{\rho}. The number hϕ(ρ)h^{\phi}(\rho) is also called the topological entropy of a Hitchin representation ρ\rho with respect to the ϕ\phi-length. Indeed, hϕ(ρ)h^{\phi}(\rho) is the topological entropy of a translation flow as we recall in the next section.

6.2. Translation flows and metric Anosov flows

In this subsection, we recall how to equip Hitchin representations with translation flows which will allow us to use thermodynamic formalism tools to study them. We start by recalling the construction of Sambarino’s translation flow [49] which is analogous to Hopf’s parametrization of the geodesic flow of a negatively curved manifold.

Let 2Γ={(x,y)Γ×Γ:xy}\partial^{2}\Gamma=\{(x,y)\in\partial\Gamma\times\partial\Gamma:x\neq y\}. The translation flow {φt}t\{\varphi_{t}\}_{t\in\mathbb{R}} on 2Γ×\partial^{2}\Gamma\times\mathbb{R} is

φt(x,y,s)=(x,y,st).\varphi_{t}(x,y,s)=(x,y,s-t).

Given a Hölder cocycle c:Γ×Γc\colon\Gamma\times\partial\Gamma\to\mathbb{R}, the group Γ\Gamma acts on 2Γ×\partial^{2}\Gamma\times\mathbb{R} by

γ(x,y,t)=(γx,γy,tc(γ,y)).\gamma(x,y,t)=(\gamma x,\gamma y,t-c(\gamma,y)).

The translation flow {φt}t\{\varphi_{t}\}_{t\in\mathbb{R}} then descends to the quotient McM_{c} of 2Γ×\partial^{2}\Gamma\times\mathbb{R} by the Γ\Gamma-action via cc. Sambarino proves a reparametrization theorem for the translation flow φt\varphi_{t}.

First, recall that two (Hölder) continuous flows ψt\psi_{t} and ψt\psi_{t}^{\prime} on compact metric spaces XX and YY, respectively, are (Hölder) conjugated if there exists a (bi-Hölder) homeomorphism h:XYh\colon X\to Y such that ψt=hψth1\psi_{t}^{\prime}=h\circ\psi_{t}\circ h^{-1} for every tt\in\mathbb{R}.

Theorem 6.4 (Sambarino [49, Thm 3.2]).

Given a Hölder cocycle cc with non-negative periods and hc(0,)h_{c}\in(0,\infty), the action defined by the cocycle cc on 2Γ×\partial^{2}\Gamma\times\mathbb{R} is proper and co-compact. Moreover, the translation flow φt\varphi_{t} on the quotient space McM_{c} is Hölder conjugated to a Hölder reparametrization of the geodesic flow on T1X0T^{1}X_{0}, where X0X_{0} is a(ny) hyperbolic structure on SS. The translation flow on McM_{c} is topologically mixing and its topological entropy equals hch_{c}.

From now on we will always assume that the cocycle cc is of the form BρϕB^{\phi}_{\rho} for some Hitchin representation ρ\rho and ϕΔ\phi\in\Delta. In this case, the hypotheses of Theorem 6.4 are satisfied. We will still denote the translation flow associated to BρϕB^{\phi}_{\rho} by φt\varphi_{t} and write Mρϕ=MBρϕM_{\rho}^{\phi}=M_{B^{\phi}_{\rho}}.

We want to use Thermodynamic formalism methods for the translation flow φt\varphi_{t} and we briefly explain why we can do so in this setting. While it is well known that a Hölder reparametrization of an Anosov flow is an Anosov flow [3, Page 122], Hölder conjugacy does not necessarily preserve the Anosov property [26]. In order to construct symbolic codings and use results from the Thermodynamic formalism for Hitchin representations, we will work in the more general setting of metric Anosov flows. Metric Anosov flows (or Smale flows) were introduced by Pollicott in [45] to generalize classical results for Anosov flows and Axiom A flows. In particular, he constructed a symbolic coding for metric Anosov flows [45].

Our definition of metric Anosov flow, which is better suited to our purposes, differs slightly from the original definition in [45]. Recall for any continuous flow ψt\psi_{t} on a compact metric space XX, the local stable set of a point xXx\in X is defined for ϵ>0\epsilon>0 as

Wϵs(x)={yX:d(ψtx,ψty)ϵ, t0 and d(ψtx,ψty)0 as t}W^{s}_{\epsilon}(x)=\{y\in X:d(\psi_{t}x,\psi_{t}y)\leq\epsilon,\,\text{ }\forall t\geq 0\textnormal{ and }d(\psi_{t}x,\psi_{t}y)\to 0\textnormal{ as }t\to\infty\}

The local unstable set of a point xx for ϵ>0\epsilon>0 is

Wϵu(x)={yX:d(ψtx,ψty)ϵ, t0 and d(ψtx,ψty)0 as t}W^{u}_{\epsilon}(x)=\{y\in X:d(\psi_{-t}x,\psi_{-t}y)\leq\epsilon,\,\text{ }\forall t\geq 0\textnormal{ and }d(\psi_{-t}x,\psi_{-t}y)\to 0\textnormal{ as }t\to\infty\}
Definition 6.5.

A continuous flow ψ\psi on a compact metric space XX is metric Anosov if

  1. (1)

    There exist positive C,λ,ϵC,\lambda,\epsilon and α(0,1]\alpha\in(0,1] such that

    d(ψtx,ψty)Ceλtd(x,y)α when yWϵs(x) and t0d(\psi_{t}x,\psi_{t}y)\leq Ce^{-\lambda t}d(x,y)^{\alpha}\text{ when $y\in W^{s}_{\epsilon}(x)$ and $t\geq 0$}

    and

    d(ψtx,ψty)Ceλtd(x,y)α when yWϵu(x) and t0.d(\psi_{-t}x,\psi_{-t}y)\leq Ce^{-\lambda t}d(x,y)^{\alpha}\text{ when $y\in W^{u}_{\epsilon}(x)$ and $t\geq 0$}.
  2. (2)

    There exists δ>0\delta>0 and a continuous map υ\upsilon on the set {(x,y)X×X:d(x,y)δ}\{(x,y)\in X\times X:d(x,y)\leq\delta\} such that υ=υ(x,y)\upsilon=\upsilon(x,y) is the unique value for which Wϵu(ψυx)Wϵs(y)W^{u}_{\epsilon}(\psi_{\upsilon}x)\cap W^{s}_{\epsilon}(y) is nonempty. The set Wϵu(ψυx)Wϵs(y)W^{u}_{\epsilon}(\psi_{\upsilon}x)\cap W^{s}_{\epsilon}(y) consists of a single point, denoted as x,y\braket{x,y}.

Remark 6.6.

Metric Anosov flows in Definition 6.5 have a symbolic coding. Indeed, these flows verify one of the key properties ([10, Lemma 1.5]) used by Bowen to build symbolic codings for Axiom A flows. From this, and after adapting Bowen’s arguments to accommodate for the exponent α\alpha in the definition, we observe that the metric Anosov flows in Definition 6.5 satisfy the expansivity, tracing and specification properties [10, Prop 1.6 and Section 2]. These are also the crucial properties used by Pollicott in his construction of symbolic codings for Smale flows [45].

Proposition 6.7.

Let ψ1\psi^{1} be a Hölder continuous metric Anosov flow on a compact metric space XX. If ψ2\psi^{2} is a flow on a compact metric space YY and ψ2\psi^{2} is Hölder conjugate to ψ1\psi^{1}, then ψ2\psi^{2} is a Hölder continuous metric Anosov flow.

Proof.

We show here that a Hölder conjugacy preserve metric Anosov properties. By hypothesis, there exists a bi-Hölder homeomorphism h:XYh:X\to Y with Hölder exponent α0(0,1]\alpha_{0}\in(0,1] (for both hh and h1h^{-1}) such that hψt1=ψt2hh\circ\psi_{t}^{1}=\psi_{t}^{2}\circ h. We want to show that ψt2\psi_{t}^{2} also satisfies the metric Anosov property.

Given x2Yx_{2}\in Y, we want to find suitable parameters so that conditions (1) and (2) in Definition 6.5 hold. We show condition (1) for local stable sets. Suppose x2=h(x1)x_{2}=h(x_{1}). Because ψt1\psi_{t}^{1} is metric Anosov, we can find ε1,C1,λ1,δ1\varepsilon_{1},C_{1},\lambda_{1},\delta_{1} positive and α1(0,1]\alpha_{1}\in(0,1] so that

Wϵ1s(x1)={y1X:dX(ψt1x1,ψt1y1)ϵ1 and dX(ψt1x1,ψt1y1)C1eλ1tdX(x1,y1)α1, t0}.W^{s}_{\epsilon_{1}}(x_{1})=\{y_{1}\in X:d_{X}(\psi^{1}_{t}x_{1},\psi^{1}_{t}y_{1})\leq\epsilon_{1}\text{ and }d_{X}(\psi^{1}_{t}x_{1},\psi^{1}_{t}y_{1})\leq C_{1}e^{-\lambda_{1}t}d_{X}(x_{1},y_{1})^{\alpha_{1}},\text{ }\forall t\geq 0\}.

Note

dY(ψt2h(x1),ψt2h(y1))\displaystyle d_{Y}(\psi^{2}_{t}h(x_{1}),\psi^{2}_{t}h(y_{1})) =dY(h(ψt1x1),h(ψt1y1))ChdX(ψt1x1,ψt1y1)α0\displaystyle=d_{Y}(h(\psi_{t}^{1}x_{1}),h(\psi_{t}^{1}y_{1}))\leq C_{h}d_{X}(\psi^{1}_{t}x_{1},\psi^{1}_{t}y_{1})^{\alpha_{0}}
Ch(C1eλ1tdX(x1,y1)α1)α0C2eλ1α0tdY(h(x1),h(y1))α02α1\displaystyle\leq C_{h}(C_{1}e^{-\lambda_{1}t}d_{X}(x_{1},y_{1})^{\alpha_{1}})^{\alpha_{0}}\leq C_{2}e^{-\lambda_{1}\alpha_{0}t}d_{Y}(h(x_{1}),h(y_{1}))^{\alpha_{0}^{2}\alpha_{1}}

Here we have used the Hölder properties of both hh and h1h^{-1} and ChC_{h} is a constant from the Hölder property of hh. This implies that we can find ε2>0\varepsilon_{2}>0 so that Wε2s(x2)W^{s}_{\varepsilon_{2}}(x_{2}) satisfies condition (1) in Definition 6.5 for ψt2\psi^{2}_{t} with positive C2,λ2=λ1α0C_{2},\lambda_{2}=\lambda_{1}\alpha_{0} and α2=α02α1\alpha_{2}=\alpha_{0}^{2}\alpha_{1}. The argument is similar for local unstable sets. Furthermore, given x2,y2Yx_{2},y_{2}\in Y close enough, condition (2) in Definition 6.5 can be verified by taking x2,y2=h(h1(x2),h1(y2))\langle x_{2},y_{2}\rangle=h(\langle h^{-1}(x_{2}),h^{-1}(y_{2})\rangle). We conclude that ψt2\psi^{2}_{t} is metric Anosov. ∎

We now show that the translation flow is a metric Anosov flow.

Proposition 6.8.

The translation flow φt\varphi_{t} on MρϕM^{\phi}_{\rho} is a metric Anosov flow which admits a symbolic coding with Hölder continuous roof functions.

Proof.

Let X0X_{0} be a base hyperbolic surface. The geodesic flow on T1X0T^{1}X_{0} is a (metric) Anosov flow which admits a symbolic coding with Hölder roof function. From Theorem 6.4, we know that the translation flow φt\varphi_{t} associated to the cocycle BρϕB_{\rho}^{\phi} defined on the quotient space MρϕM^{\phi}_{\rho} is Hölder conjugate to a Hölder reparametrization of the geodesic flow ψt\psi_{t} on T1X0T^{1}X_{0}. In other words, there exists a Hölder continuous function f:T1X00f\colon T^{1}X_{0}\to\mathbb{R}_{\geq 0} such that (h1φh)t=ψtf(h^{-1}\circ\varphi\circ h)_{t}=\psi^{f}_{t} for all tt\in\mathbb{R}. Since ψtf\psi^{f}_{t} is an Anosov flow (see [3, Page 122]), ϕt\phi_{t} is a metric Anosov flow by Proposition 6.7.

The coding is therefore preserved. The roof function remains Hölder by either Hölder conjugacy or Hölder reparametrization as it is given by an opportune composition of Hölder functions. The translation flow is therefore a metric Anosov flow on a compact metric space that admits a symbolic coding with Hölder roof function. ∎

6.3. Reparametrization functions and independence lemma

After constructing flows for Hitchin representations, the next thing that we want to do is to define the reparametrization function in the setting of Hitchin representations. In Section 2.2, given a positive Hölder continuous function ff on the unit tangent bundle T1XρT^{1}X_{\rho} of a convex real projective structure, we defined a reparametrization of the flow by time change. This can be done more generally for any Hölder continuous flow on a compact metric space XX. In particular, given two Hitchin representations ρ1\rho_{1} and ρ2\rho_{2} in d(S)\mathcal{H}_{d}(S), we will show the existence of a positive Hölder continuous reparametrization function fρ1ρ2:Mρ1ϕ>0f_{\rho_{1}}^{\rho_{2}}:M^{\phi}_{\rho_{1}}\to\mathbb{R}_{>0} that encodes the ϕ\phi-length spectrum of ρ2\rho_{2}.

We start with a lemma that relates the positivity of the Hölder reparametrization function to the positivity of its entropy. The entropy of a Hölder function ff on a compact metric space XX equipped with a metric Anosov flow is defined as

hf=lim supT1T#log{τ periodicτfT}.h_{f}=\limsup_{T\to\infty}\frac{1}{T}\#\log\left\{\tau\text{ periodic}\mid\int_{\tau}f\leq T\right\}.
Lemma 6.9 (Ledrappier [37, Lemma 1], Sambarino [49, Lemma 3.8]).

Let f:Xf:X\to\mathbb{R} be a Hölder continuous function with non-negative periods on a compact metric space XX equipped with a topological transitive metric Anosov flow. The following are equivalent:

  1. (1)

    the function ff is cohomologous to a positive Hölder continuous function.

  2. (2)

    there exists κ>0\kappa>0 such that τf>κp(τ)\int_{\tau}f>\kappa p(\tau) where p(τ)p(\tau) is the period of τ\tau.

  3. (3)

    the entropy hf(0,)h_{f}\in(0,\infty).

We also need the following theorem of Ledrappier [37] which establishes the correspondence between cohomologous Hölder cocycles and cohomologous Hölder continuous functions on T1X0T^{1}X_{0}. We will state this theorem for vector valued Hölder cocycles. Their definition is analogous to Definition 6.3.

Theorem 6.10 (Ledrappier [37, page 105]).

Let VV be a finite dimensional vector space. For each Hölder cocycle c:Γ×ΓVc:\Gamma\times\partial\Gamma\to V, there exists a Hölder continuous map Fc:T1X0VF_{c}:T^{1}X_{0}\to V, such that for every γΓ{e}\gamma\in\Gamma-\{e\}

c(γ)=[γ]Fc.\ell_{c}(\gamma)=\int_{[\gamma]}F_{c}.

This map cFcc\to F_{c} induces a bijection between the set of cohomology classes of VV-valued Hölder cocycles and the set of cohomology classes of Hölder maps from T1X0T^{1}X_{0} to VV.

For a Hitchin representation ρ\rho, this tells us that we can find a Hölder continuous map gρ:T1X0𝔞g_{\rho}:T^{1}X_{0}\to\mathfrak{a} with periods equal to the periods of the vector valued Busemann cocycle BρB_{\rho}. Since hϕgρ=hBρϕ=hϕ(ρ)h_{\phi\circ g_{\rho}}=h_{B^{\phi}_{\rho}}=h^{\phi}(\rho) is finite and positive, by Lemma 6.9, the reparametrization ϕgρ\phi\circ g_{\rho} on T1X0T^{1}X_{0} is cohomologous to a positive Hölder continuous function.

Now we are ready to state our lemma about the existence of positive reparametrization functions, which should be compared to Lemma 2.3.

Lemma 6.11.

Let ρ1\rho_{1}and ρ2\rho_{2} be two different representations in d(S)\mathcal{H}_{d}(S). There exists a positive Hölder continuous function fρ1ρ2:Mρ1ϕ>0f_{\rho_{1}}^{\rho_{2}}:M^{\phi}_{\rho_{1}}\to\mathbb{R}_{>0} such that for every periodic orbit τ\tau corresponding to [γ][Γ][\gamma]\in[\Gamma]

λ(fρ1ρ2,τ)=ρ2ϕ([γ]).\lambda\left(f_{\rho_{1}}^{\rho_{2}},\tau\right)=\ell_{\rho_{2}}^{\phi}([\gamma]).
Proof.

We denote the translation flow for MρiϕM^{\phi}_{\rho_{i}} with respect to the Busemann cocycle BρiϕB_{\rho_{i}}^{\phi} as φti\varphi^{i}_{t} for i=1,2i=1,2. Now since BρiϕB^{\phi}_{\rho_{i}} has positive entropy, by Theorem 6.4, both φti\varphi^{i}_{t} are Hölder conjugate to Hölder reparametrizations of the geodesic flow on T1X0T^{1}X_{0}, where X0X_{0} is an auxiliary hyperbolic surface. The reparametrization functions for φti\varphi^{i}_{t} on T1X0T^{1}X_{0} can be chosen to be positive by the discussion after Theorem 6.10. Generalizing Remark 2.2, we conclude that the flow φt2\varphi^{2}_{t} is Hölder conjugate to a Hölder reparametrization of φt1\varphi^{1}_{t} on Mρ1ϕM^{\phi}_{\rho_{1}}. Therefore, there exists a Hölder function f:Mρ1ϕf\colon M^{\phi}_{\rho_{1}}\to\mathbb{R} such that the flow φt2\varphi^{2}_{t} is Hölder conjugate to (φt1)f\big{(}\varphi^{1}_{t}\big{)}^{f} where the flow (φt1)f\big{(}\varphi^{1}_{t}\big{)}^{f} is a reparametrization of φt1\varphi^{1}_{t} by ff. Because hf=hBρϕh_{f}=h_{B^{\phi}_{\rho}} is positive and finite, we can always choose f=fρ1ρ2f=f_{\rho_{1}}^{\rho_{2}} in its cohomology class to be a positive function by Lemma 6.9. One easily checks that λ(f,τ)=ρ2ϕ([γ])\lambda\left(f,\tau\right)=\ell_{\rho_{2}}^{\phi}([\gamma]). ∎

Once we have defined reparametrization functions, we see that all definitions, remarks and results in Section 2.3 regarding thermodynamic formalism readily generalize to this context by simply changing the domain from T1XρT^{1}X_{\rho} to MρϕM_{\rho}^{\phi}.

6.4. Correlation and Manhattan curve theorem revisited

First, we state the independence lemma for Hitchin representations which generalizes Lemma 2.12.

Lemma 6.12 (Independence lemma).

Consider ϕΔ\phi\in\Delta and Hitchin representations ρ1,ρ2d(S)\rho_{1},\rho_{2}\in\mathcal{H}_{d}(S) such that ρ2ρ1\rho_{2}\neq\rho_{1} or ρ1\rho_{1}^{\ast}. If there exist a1,a2a_{1},a_{2}\in\mathbb{R} such that a1ρ1ϕ([γ])+a2ρ2ϕ([γ])a_{1}\ell^{\phi}_{\rho_{1}}([\gamma])+a_{2}\ell^{\phi}_{\rho_{2}}([\gamma])\in\mathbb{Z} for all [γ][Γ][\gamma]\in[\Gamma], then a1=a2=0a_{1}=a_{2}=0.

Proof.

The Zariski closure Gi\mathrm{G}_{i} of ρi(Γ)\rho_{i}(\Gamma) is simple and connected by a result of Guichard (see [14, Theorem 11.7]). Then, we argue by contradiction as in the proof of Lemma 2.12. ∎

Recall that a flow is weakly mixing if its periods do not generate a discrete subgroup of \mathbb{R}. In particular, the independence lemma implies that the translation flow φt\varphi_{t} is weakly mixing.

We are now ready to state our correlation theorem for Hitchin representations. Recall that we denote the renormalized ϕ\phi length spectrum of ρ\rho by Lρϕ=hϕ(ρ)ρϕL^{\phi}_{\rho}=h^{\phi}(\rho)\ell^{\phi}_{\rho},

Theorem 1.7. Given a linear functional ϕΔ\phi\in\Delta and a fixed precision ε>0\varepsilon>0, for any two different Hitchin representations ρ1,ρ2:ΓPSL(d,)\rho_{1},\rho_{2}\colon\Gamma\to\mathrm{PSL}(d,\mathbb{R}) such that ρ2ρ1\rho_{2}\neq\rho_{1}^{*}, there exist constants C=C(ε,ρ1,ρ2,ϕ)>0C=C(\varepsilon,\rho_{1},\rho_{2},\phi)>0 and M=M(ρ1,ρ2,ϕ)(0,1)M=M(\rho_{1},\rho_{2},\phi)\in(0,1) such that

#{[γ][Γ]|Lρ1ϕ([γ])(x,x+hϕ(ρ1)ε),Lρ2ϕ([γ])(x,x+hϕ(ρ2)ε)}CeMxx3/2.\#\Big{\{}[\gamma]\in[\Gamma]\,\Big{|}\,L_{\rho_{1}}^{\phi}([\gamma])\in\big{(}x,x+h^{\phi}(\rho_{1})\varepsilon\big{)},\ L_{\rho_{2}}^{\phi}([\gamma])\in\big{(}x,x+h^{\phi}(\rho_{2})\varepsilon\big{)}\Big{\}}\sim C\frac{e^{Mx}}{x^{3/2}}.
Proof.

By Proposition 6.8 and Lemma 6.12, the translation flow associated to the linear functional ϕ\phi is a weakly mixing metric Anosov flow on Mρ1ϕM^{\phi}_{\rho_{1}} that admits a symbolic coding with Hölder continuous roof function. Moreover, ρ1\rho_{1} and ρ2\rho_{2} are independent thanks to the independence lemma 6.12. Thus we can apply Lalley and Sharp’s Theorem 3.1 which hold in the context of flows with a symbolic coding with Hölder roof function.

Recall J(fρ1ρ2)J(f_{\rho_{1}}^{\rho_{2}}) is the open interval of values P(tfρ1ρ2)P^{\prime}(tf_{\rho_{1}}^{\rho_{2}}) for tt\in\mathbb{R}. We then want to verify hϕ(ρ1)hϕ(ρ2)J(fρ1ρ2)\frac{h^{\phi}(\rho_{1})}{h^{\phi}(\rho_{2})}\in J(f_{\rho_{1}}^{\rho_{2}}). The proof of this fact follows from the same argument as in the proof for Theorem 1.1. ∎

Remark 6.13.

Fix ϕΔ\phi\in\Delta. Sambarino’s orbit counting theorem [49, Thm 7.8] implies that the correlation number M=M(ρ1,ρ2,ϕ)M=M(\rho_{1},\rho_{2},\phi) converges to one and C(ε,ρ1,ρ2,ϕ)C(\varepsilon,\rho_{1},\rho_{2},\phi) diverges if ρ1\rho_{1} and ρ2\rho_{2} converges to a Hitchin representation with the same ϕ\phi-length spectrum.

For two Hitchin representations ρ1\rho_{1}, ρ2\rho_{2} with ρ2ρ1,ρ1\rho_{2}\neq\rho_{1},\rho_{1}^{*} and ϕΔ\phi\in\Delta consider the Manhattan curve

𝒞ϕ(ρ1,ρ2)={(a,b)2:P(abf)=0}\mathcal{C}^{\phi}(\rho_{1},\rho_{2})=\{(a,b)\in\mathbb{R}^{2}\colon P(-a-bf)=0\}

where ff is the reparametrization function from Lemma 6.11. We obtain a characterization of the correlation number analogous to Theorem 4.2.

Theorem 6.14.

Fix ϕΔ\phi\in\Delta, and let ρ1\rho_{1} and ρ2\rho_{2} be Hitchin representations in d(S)\mathcal{H}_{d}(S) such that ρ2ρ1,ρ1\rho_{2}\neq\rho_{1},\rho_{1}^{*}. Their correlation number can be written as

M(ρ1,ρ2,ϕ)=ahϕ(ρ1)+bhϕ(ρ2)M(\rho_{1},\rho_{2},\phi)=\frac{a}{h^{\phi}(\rho_{1})}+\frac{b}{h^{\phi}(\rho_{2})}

where (a,b)𝒞ϕ(ρ1,ρ2)(a,b)\in\mathcal{C}^{\phi}(\rho_{1},\rho_{2}) is the point on the Manhattan curve at which the tangent line is parallel to the line passing through (hϕ(ρ1),0)(h^{\phi}(\rho_{1}),0) and (0,hϕ(ρ2))(0,h^{\phi}(\rho_{2})).

We conclude by raising two questions which are motivated by Theorem 1.3 and Theorem 1.5.

Recall that Potrie and Sambarino [47, Thm B] showed that for every Hitchin representation ρ\rho, the simple root lengths are such that hαi(ρ)=1h^{\alpha_{i}}(\rho)=1 for i=1,,d1i=1,\dots,d-1. Moreover, the simple root lengths restrict to the hyperbolic length on the Fuchsian locus 𝒯(S)d(S)\mathcal{T}(S)\subset\mathcal{H}_{d}(S). Thus, Theorem 1.7 in this case is a particularly natural generalization of the correlation theorem for hyperbolic surfaces [51]. Theorem 1.3 exhibits examples of sequences for which the αi\alpha_{i}-correlation number decays. We ask whether there exist similar examples which lie outside the Fuchsian locus.

Question 6.15.

For d3d\geq 3 and i=1,,d1i=1,\cdots,d-1, do there exist sequences (ρn)n=1(\rho_{n})_{n=1}^{\infty} and (ηn)n=1(\eta_{n})_{n=1}^{\infty} in d(S)\mathcal{H}_{d}(S) which leave every compact neighborhood of the Fuchsian locus and such that the correlation numbers satisfy limnM(ρn,ηn,αi)=0\lim\limits_{n\to\infty}M(\rho_{n},\eta_{n},\alpha_{i})=0?

Finally, we raise a conjecture motivated by Theorem 1.5. Similarly to cubic rays introduced in section 5, in general for the Hitchin component d(S)\mathcal{H}_{d}(S), one can consider a dd-th order holomorphic differential qq over X0X_{0} and its associated family of representations (ρt)t(\rho_{t})_{\scriptscriptstyle t\in\mathbb{R}} in d(S)\mathcal{H}_{d}(S) parametrized by (tq)t(tq)_{t\in\mathbb{R}}. These representations are given by holonomies of cyclic Higgs bundles. We refer to [4] for a detailed definition and discussion. When t0t\geq 0, we say this family of representations (ρt)t0(\rho_{t})_{t\geq 0} is a ray associated to qq.

Conjecture 6.16.

Let (ρt)t0(\rho_{t})_{t\geq 0} and (ηt)t0(\eta_{t})_{t\geq 0} be two rays associated to two different dd-th order holomorphic differentials q1q_{1} and q2q_{2} on a hyperbolic structure X0X_{0} such that q1,q2q_{1},q_{2} have unit L2L^{2}-norm with respect to X0X_{0} and q1q2q_{1}\neq-q_{2}. Then the correlation number M(ρt,ηt,i=1d1αi)M(\rho_{t},\eta_{t},\sum_{i=1}^{d-1}\alpha_{i}) is uniformly bounded away from zero as tt goes to infinity.

Acknowledgements

We would like to thank Harrison Bray, Richard Canary, León Carvajales and Michael Wolf for several insightful conversations and suggestions. We are very grateful to Richard Canary and an anonymous referee for their comments on early versions of this manuscript. GM acknowledges partial support by the American Mathematical Society and the Simons Foundation. XD wants to thank Rice math department for providing support during the preparation of this paper.

References

  • [1] L. M. Abramov. On the entropy of a flow. Dokl. Akad. Nauk SSSR, 128:873–875, 1959.
  • [2] D. Alessandrini, G.-S. Lee, and F. Schaffhauser. Hitchin components for orbifolds. Preprint arXiv:1811.05366, to appear on Journal of the European Mathematical Society, pages 1–40, 2018.
  • [3] D. V. Anosov and Y. G. Sinai. Some Smooth Ergodic Systems. Russian Mathematical Surveys, 22(5):103–167, Oct. 1967.
  • [4] D. Baraglia. Cyclic Higgs bundles and the affine Toda equations. Geom. Dedicata, 174:25–42, 2015.
  • [5] Y. Benoist. Automorphismes des cônes convexes. Invent. Math., 141(1):149–193, 2000.
  • [6] Y. Benoist. Propriétés asymptotiques des groupes linéaires. II. In Analysis on homogeneous spaces and representation theory of Lie groups, Okayama–Kyoto (1997), volume 26 of Adv. Stud. Pure Math., pages 33–48. Math. Soc. Japan, Tokyo, 2000.
  • [7] Y. Benoist. Convexes divisibles. I. In Algebraic groups and arithmetic, pages 339–374. Tata Inst. Fund. Res., Mumbai, 2004.
  • [8] F. Bonahon. Bouts des variétés hyperboliques de dimension 33. Ann. of Math. (2), 124(1):71–158, 1986.
  • [9] F. Bonahon. The geometry of Teichmüller space via geodesic currents. Invent. Math., 92(1):139–162, 1988.
  • [10] R. Bowen. Periodic orbits for hyperbolic flows. Amer. J. Math., 94:1–30, 1972.
  • [11] R. Bowen. Symbolic dynamics for hyperbolic flows. Amer. J. Math., 95:429–460, 1973.
  • [12] R. Bowen. Equilibrium states and the ergodic theory of Anosov diffeomorphisms, volume 470 of Lecture Notes in Mathematics. Springer-Verlag, Berlin, revised edition, 2008. With a preface by David Ruelle, Edited by Jean-René Chazottes.
  • [13] R. Bowen and D. Ruelle. The ergodic theory of Axiom A flows. Invent. Math., 29(3):181–202, 1975.
  • [14] M. Bridgeman, R. Canary, F. Labourie, and A. Sambarino. The pressure metric for Anosov representations. Geom. Funct. Anal., 25(4):1089–1179, 2015.
  • [15] M. Bridgeman, R. Canary, F. Labourie, and A. Sambarino. Simple root flows for Hitchin representations. Geom. Dedicata, 192:57–86, 2018.
  • [16] M. Bridgeman, R. Canary, and A. Sambarino. An introduction to pressure metrics for higher Teichmüller spaces. Ergodic Theory Dynam. Systems, 38(6):2001–2035, 2018.
  • [17] M. Burger. Intersection, the Manhattan curve, and Patterson-Sullivan theory in rank 22. Internat. Math. Res. Notices, (7):217–225, 1993.
  • [18] M. Burger, A. Iozzi, A. Parreau, and M. B. Pozzetti. Currents, Systoles, and Compactifications of Character Varieties. Preprint arXiv:1902.07680, to appear in PLMS, pages 1–36, 2019.
  • [19] S. Choi and W. M. Goldman. Convex real projective structures on closed surfaces are closed. Proc. Amer. Math. Soc., 118(2):657–661, 1993.
  • [20] D. Cooper and K. Delp. The marked length spectrum of a projective manifold or orbifold. Proc. Amer. Math. Soc., 138(9):3361–3376, 2010.
  • [21] M. Crampon. Entropies of strictly convex projective manifolds. J. Mod. Dyn., 3(4):511–547, 2009.
  • [22] F. Dal’bo. Remarques sur le spectre des longueurs d’une surface et comptages. Bol. Soc. Brasil. Mat. (N.S.), 30(2):199–221, 1999.
  • [23] V. Fock and A. Goncharov. Moduli spaces of local systems and higher Teichmüller theory. Publ. Math. Inst. Hautes Études Sci., (103):1–211, 2006.
  • [24] O. Glorieux. Counting closed geodesics in globally hyperbolic maximal compact AdS 3-manifolds. Geom. Dedicata, 188:63–101, 2017.
  • [25] O. Glorieux. The embedding of the space of negatively curved surfaces in geodesic currents. Preprint arXiv:1904.02558, pages 1–8, 2019.
  • [26] A. Gogolev. Diffeomorphisms Hölder conjugate to Anosov diffeomorphisms. Ergodic Theory Dynam. Systems, 30(2):441–456, 2010.
  • [27] W. M. Goldman. Convex real projective structures on compact surfaces. J. Differential Geom., 31(3):791–845, 1990.
  • [28] H. Gündoğan. The component group of the automorphism group of a simple Lie algebra and the splitting of the corresponding short exact sequence. J. Lie Theory, 20(4):709–737, 2010.
  • [29] M. Hall, Jr. The theory of groups. The Macmillan Co., New York, N.Y., 1959.
  • [30] N. J. Hitchin. Lie groups and Teichmüller space. Topology, 31(3):449–473, 1992.
  • [31] H. Huber. Zur analytischen Theorie hyperbolischer Raumformen und Bewegungsgruppen. II. Math. Ann., 142:385–398, 1960/61.
  • [32] A. Katok and B. Hasselblatt. Introduction to the modern theory of dynamical systems. Cambridge University Press, 1995.
  • [33] I. Kim. Rigidity and deformation spaces of strictly convex real projective structures on compact manifolds. J. Differential Geom., 58(2):189–218, 2001.
  • [34] F. Labourie. Anosov flows, surface groups and curves in projective space. Invent. Math., 165(1):51–114, 2006.
  • [35] F. Labourie. Flat projective structures on surfaces and cubic holomorphic differentials. Pure Appl. Math. Q., 3(4, Special Issue: In honor of Grigory Margulis. Part 1):1057–1099, 2007.
  • [36] S. P. Lalley. Distribution of periodic orbits of symbolic and Axiom A flows. Adv. in Appl. Math., 8(2):154–193, 1987.
  • [37] F. Ledrappier. Structure au bord des variétés à courbure négative. Séminaire de théorie spectrale et géométrie, 13:97–122, 1994-1995.
  • [38] A. N. Livšic. Cohomology of dynamical systems. Izv. Akad. Nauk SSSR Ser. Mat., 36:1296–1320, 1972.
  • [39] J. Loftin. Survey on affine spheres. In Handbook of geometric analysis, No. 2, volume 13 of Adv. Lect. Math. (ALM), pages 161–191. Int. Press, Somerville, MA, 2010.
  • [40] J. C. Loftin. Affine spheres and convex n\mathbb{RP}^{n}-manifolds. Amer. J. Math., 123(2):255–274, 2001.
  • [41] G. A. Margulis. On some aspects of the theory of Anosov systems. Springer Monographs in Mathematics. Springer-Verlag, Berlin, 2004. With a survey by Richard Sharp: Periodic orbits of hyperbolic flows, Translated from the Russian by Valentina Vladimirovna Szulikowska.
  • [42] G. Martone and T. Zhang. Positively ratioed representations. Comment. Math. Helv., 94(2):273–345, 2019.
  • [43] X. Nie. On the Hilbert geometry of simplicial Tits sets. Ann. Inst. Fourier (Grenoble), 65(3):1005–1030, 2015.
  • [44] W. Parry and M. Pollicott. Zeta functions and the periodic orbit structure of hyperbolic dynamics. Astérisque, (187-188):268, 1990.
  • [45] M. Pollicott. Symbolic dynamics for Smale flows. Amer. J. Math., 109(1):183–200, 1987.
  • [46] M. Pollicott and R. Sharp. Correlations of length spectra for negatively curved manifolds. Comm. Math. Phys., 319(2):515–533, 2013.
  • [47] R. Potrie and A. Sambarino. Eigenvalues and entropy of a Hitchin representation. Invent. Math., 209(3):885–925, 2017.
  • [48] J.-F. Quint. Mesures de Patterson-Sullivan en rang supérieur. Geom. Funct. Anal., 12(4):776–809, 2002.
  • [49] A. Sambarino. Quantitative properties of convex representations. Comment. Math. Helv., 89(2):443–488, 2014.
  • [50] A. Sambarino. The orbital counting problem for hyperconvex representations. Ann. Inst. Fourier (Grenoble), 65(4):1755–1797, 2015.
  • [51] R. Schwartz and R. Sharp. The correlation of length spectra of two hyperbolic surfaces. Comm. Math. Phys., 153(2):423–430, 1993.
  • [52] R. Sharp. Prime orbit theorems with multi-dimensional constraints for Axiom A flows. Monatsh. Math., 114(3-4):261–304, 1992.
  • [53] R. Sharp. The Manhattan curve and the correlation of length spectra on hyperbolic surfaces. Math. Z., 228(4):745–750, 1998.
  • [54] N. Tholozan. Volume entropy of Hilbert metrics and length spectrum of Hitchin representations into PSL(3,){\mathrm{P}SL}(3,\mathbb{R}). Duke Math. J., 166(7):1377–1403, 2017.
  • [55] T. Zhang. The degeneration of convex 2\mathbb{R}\mathbb{P}^{2} structures on surfaces. Proc. Lond. Math. Soc. (3), 111(5):967–1012, 2015.