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

Exponential mixing and essential spectral gaps for Anosov subgroups

Michael Chow Department of Mathematics, Yale University, New Haven, CT 06511, USA [email protected]  and  Pratyush Sarkar Department of Mathematics, UC San Diego, La Jolla, CA 92093, USA [email protected]
(Date: 24 décembre 2024)
Abstract.

Let Γ\Gamma be a Zariski dense Θ\Theta-Anosov subgroup of a connected semisimple real algebraic group for some nonempty subset of simple roots Θ\Theta. In the Anosov setting, there is a natural compact metric space 𝒳\mathcal{X} equipped with a family of translation flows {at𝗏}t\{a^{\mathsf{v}}_{t}\}_{t\in\mathbb{R}}, parameterized by vectors 𝗏\mathsf{v} in the interior of the Θ\Theta-limit cone Θ\mathcal{L}_{\Theta} of Γ\Gamma, and they are Hölder conjugate to Hölder reparametrizations of the Gromov geodesic flow. We prove that for all 𝗏\mathsf{v} outside an exceptional cone intΘ\mathscr{E}\subset\operatorname{int}\mathcal{L}_{\Theta}, which is a smooth image of the walls of the Weyl chamber, the translation flow is exponentially mixing with respect to the Bowen–Margulis–Sullivan measure associated to 𝗏\mathsf{v}. Moreover, the exponential rate is uniform for a compact set of such 𝗏\mathsf{v}. We also obtain an essential spectral gap formulated in terms of the Selberg zeta function and a prime orbit theorem with a power saving error term. Our proof relies on Lie theoretic techniques to prove the crucial local non-integrability condition (LNIC) for the translation flows and thereby implement Dolgopyat’s method in a uniform fashion. The exceptional cone \mathscr{E} arises from the failure of LNIC for those vectors.

1. Introduction

The main result of this paper is exponential mixing of translation flows associated to Anosov subgroups. Anosov subgroups, first introduced by Labourie [36] and later generalized by Guichard–Wienhard [25], includes many interesting geometric examples such as the images of Hitchin representations [36] into PSLn()\operatorname{PSL}_{n}(\mathbb{R}), strongly convex cocompact projective representations into PGLn()\operatorname{PGL}_{n}(\mathbb{R}) [19], maximal representations into PSp2n()\mathrm{PSp}_{2n}(\mathbb{R}) and PO(2,n)\mathrm{PO}(2,n) [10, 11], and Barbot representations into PSLn()\operatorname{PSL}_{n}(\mathbb{R}) [2]. In recent times, there has been a lot activity in understanding the dynamics related to Anosov subgroups (see for instance [60, 9, 39, 40, 23, 12, 18, 61, 31], etc.). Mixing properties are of particular interest because existing techniques allow one to readily derive various counting and equidistribution results. What is more, exponential mixing allows one to effectivize these results, i.e., obtain precise error terms.

Let us first provide some basic setup for the main theorem. Let GG be a connected semisimple real algebraic group with 𝔤:=Lie(G)\mathfrak{g}:=\operatorname{Lie}(G), AA be a maximal real split torus of GG and fix a closed positive Weyl chamber 𝔞+𝔞:=Lie(A)\mathfrak{a}^{+}\subset\mathfrak{a}:=\operatorname{Lie}(A). Let Π𝔞\Pi\subset\mathfrak{a}^{*} denote the set of all simple roots for (𝔤,𝔞+)(\mathfrak{g},\mathfrak{a}^{+}) and let 𝗂:𝔞𝔞\mathsf{i}:\mathfrak{a}\to\mathfrak{a} denote the opposition involution preserving 𝔞+\mathfrak{a}^{+} and acting on Π\Pi by precomposition. Fix a nonempty subset ΘΠ\Theta\subset\Pi. Let PΘP_{\Theta} denote the standard parabolic subgroup of GG corresponding to Θ\Theta and PΘ=SΘAΘNΘP_{\Theta}=S_{\Theta}A_{\Theta}N_{\Theta} be its Langlands decomposition where SΘS_{\Theta} centralizes AΘ:=exp(𝔞Θ)A_{\Theta}:=\exp(\mathfrak{a}_{\Theta}) and 𝔞Θ:=αΠΘkerα𝔞\mathfrak{a}_{\Theta}:=\bigcap_{\alpha\in\Pi-\Theta}\ker\alpha\subset\mathfrak{a}. The Θ\Theta-Furstenberg boundary is defined as Θ:=G/PΘ\mathcal{F}_{\Theta}:=G/P_{\Theta} and we denote the unique open GG-orbit in Θ×𝗂Θ\mathcal{F}_{\Theta}\times\mathcal{F}_{\mathsf{i}\Theta} by Θ(2)\mathcal{F}_{\Theta}^{(2)}.

Let Γ<G\Gamma<G be a torsion-free Zariski dense Θ\Theta-Anosov subgroup, that is, Γ\Gamma is a torsion-free Gromov hyperbolic group with Gromov boundary Γ\partial\Gamma and continuous Γ\Gamma-equivariant maps ζΘ:ΓΘ\zeta_{\Theta}:\partial\Gamma\to\mathcal{F}_{\Theta} and ζ𝗂Θ:Γ𝗂Θ\zeta_{\mathsf{i}\Theta}:\partial\Gamma\to\mathcal{F}_{\mathsf{i}\Theta} such that (ζΘ(x),ζ𝗂Θ(y))Θ(2)(\zeta_{\Theta}(x),\zeta_{\mathsf{i}\Theta}(y))\in\mathcal{F}_{\Theta}^{(2)} for all distinct x,yΓx,y\in\partial\Gamma. The images ΛΘ:=ζΘ(Γ)\Lambda_{\Theta}:=\zeta_{\Theta}(\partial\Gamma) and Λ𝗂Θ:=ζ𝗂Θ(Γ)\Lambda_{\mathsf{i}\Theta}:=\zeta_{\mathsf{i}\Theta}(\partial\Gamma) are the unique Γ\Gamma-minimal subsets of Θ\mathcal{F}_{\Theta} and 𝗂Θ\mathcal{F}_{\mathsf{i}\Theta}. Let ΛΘ(2)=(ΛΘ×Λ𝗂Θ)Θ(2)\Lambda_{\Theta}^{(2)}=(\Lambda_{\Theta}\times\Lambda_{\mathsf{i}\Theta})\cap\mathcal{F}_{\Theta}^{(2)}. Unless Θ=Π\Theta=\Pi (in which case SΠS_{\Pi} is compact), Γ\Gamma does not necessarily act on G/SΘΘ(2)×𝔞ΘG/S_{\Theta}\cong\mathcal{F}_{\Theta}^{(2)}\times\mathfrak{a}_{\Theta} properly discontinuously (see [24, §1.5] for an extensive discussion). However, the Γ\Gamma-action restricted to the Γ\Gamma-invariant subset ΛΘ(2)×𝔞Θ\Lambda_{\Theta}^{(2)}\times\mathfrak{a}_{\Theta} is always properly discontinuous.

Given 𝗏𝔞Θ\mathsf{v}\in\mathfrak{a}_{\Theta}, the translation action of 𝗏\mathbb{R}\mathsf{v} on the 𝔞Θ\mathfrak{a}_{\Theta}-coordinate of ΛΘ(2)×𝔞Θ\Lambda_{\Theta}^{(2)}\times\mathfrak{a}_{\Theta} induces the one-parameter diagonal flow on Ω:=Γ\(ΛΘ(2)×𝔞Θ)\Omega:=\Gamma\backslash\bigl{(}\Lambda_{\Theta}^{(2)}\times\mathfrak{a}_{\Theta}\bigr{)} which we denote by {at𝗏}t\{a_{t\mathsf{v}}\}_{t\in\mathbb{R}}. We further assume that 𝗏intΘ\mathsf{v}\in\operatorname{int}\mathcal{L}_{\Theta} where Θ\mathcal{L}_{\Theta} denotes Θ\Theta-limit cone of Γ\Gamma (see Definition 2.6). In this case, the dynamical system (Ω,m𝗏BMS,{at𝗏}t)(\Omega,m^{\mathrm{BMS}}_{\mathsf{v}},\{a_{t\mathsf{v}}\}_{t\in\mathbb{R}}) is known to be locally mixing where m𝗏BMSm^{\mathrm{BMS}}_{\mathsf{v}} is the Bowen–Margulis–Sullivan measure associated to 𝗏\mathsf{v}. That is, there exists κ𝗏>0\kappa_{\mathsf{v}}>0 such that for all ϕ1,ϕ2Cc(Ω)\phi_{1},\phi_{2}\in C_{\mathrm{c}}(\Omega), we have

limt+t#Θ12Ωϕ1(at𝗏x)ϕ2(x)𝑑m𝗏BMS(x)=κ𝗏m𝗏BMS(ϕ1)m𝗏BMS(ϕ2).\displaystyle\lim_{t\to+\infty}t^{\frac{\#\Theta-1}{2}}\int_{\Omega}\phi_{1}(a_{t\mathsf{v}}x)\phi_{2}(x)\,dm^{\mathrm{BMS}}_{\mathsf{v}}(x)=\kappa_{\mathsf{v}}m^{\mathrm{BMS}}_{\mathsf{v}}(\phi_{1})m^{\mathrm{BMS}}_{\mathsf{v}}(\phi_{2}).

When Θ=Π\Theta=\Pi, Thirion [69] first proved the above for Schottky subgroups of SLn()\operatorname{SL}_{n}(\mathbb{R}) and this was later generalized by Sambarino [60] to fundamental groups of compact negatively curved manifolds. Generalizing their work, the authors [18] proved local mixing on Γ\G\Gamma\backslash G (which is a principal SΠS_{\Pi}-bundle over Γ\(Π(2)×𝔞)\Gamma\backslash\bigl{(}\mathcal{F}_{\Pi}^{(2)}\times\mathfrak{a}\bigr{)}) for general Π\Pi-Anosov subgroups. The arguments in [18] easily generalize for arbitrary Θ\Theta without the SΠS_{\Pi}-valued holonomy (cf. [61, Appendix B] for a general theorem and a proof sketch amending the one in [60] using strategies similar to [18]). It is an interesting (and still open) question as to what the rate of mixing above should be in general. It is known to the authors that Dolgopyat’s method cannot be carried out as the non-concentration property or its generalization (cf. [64, 17]) fails miserably whenever GG is of higher rank; and hence, the rate is expected to be subexponential when #Θ>1\#\Theta>1. See the results below for the case #Θ=1\#\Theta=1.

On the other hand, the situation is completely different for the dynamical system (𝒳,m𝒳𝗏,{at𝗏}t)(\mathcal{X},m_{\mathcal{X}}^{\mathsf{v}},\{a^{\mathsf{v}}_{t}\}_{t\in\mathbb{R}}). Here, 𝒳\mathcal{X} is a compact metric space over which π:Ω𝒳\pi:\Omega\to\mathcal{X} is a trivial #Θ1\mathbb{R}^{\#\Theta-1}-bundle, m𝒳𝗏m_{\mathcal{X}}^{\mathsf{v}} is a probability measure on 𝒳\mathcal{X} such that we have the product structure m𝗏BMS=m𝒳𝗏Leb#Θ1m^{\mathrm{BMS}}_{\mathsf{v}}=m_{\mathcal{X}}^{\mathsf{v}}\otimes\mathrm{Leb}_{\mathbb{R}^{\#\Theta-1}}, and {at𝗏}t\{a^{\mathsf{v}}_{t}\}_{t\in\mathbb{R}} is the translation flow on 𝒳\mathcal{X} induced by {at𝗏}t\{a_{t\mathsf{v}}\}_{t\in\mathbb{R}} via the bundle projection (see Subsection 2.4 for details). So far in the literature on Anosov subgroups, translation flows have been studied more often as an auxiliary tool rather than a dynamical system of intrinsic interest. However, translation flows are natural dynamical systems which mimic (and in fact generalize) the geodesic flow for convex cocompact hyperbolic manifolds and they are Hölder conjugate to Hölder reparametrizations of the Gromov geodesic flow associated to Γ\Gamma. In this vein, we prove that translation flows are exponentially mixing for generic vectors 𝗏intΘ\mathsf{v}\in\operatorname{int}\mathcal{L}_{\Theta}. Indeed, we obtain a detailed characterization of the exponential mixing property for the whole family of translation flows.

The set of generic 𝗏intΘ\mathsf{v}\in\operatorname{int}\mathcal{L}_{\Theta} for which we can prove exponential mixing can be described in terms of a fundamental object associated to Γ\Gamma called its Θ\Theta-growth indicator ψΘ:Θ[0,+)\psi_{\Theta}:\mathcal{L}_{\Theta}\to[0,+\infty) (see Definition 2.7) which is the higher rank generalization of critical exponent in rank one. We say 𝗏intΘ\mathsf{v}\in\operatorname{int}\mathcal{L}_{\Theta} is ψΘ\psi_{\Theta}-regular if ψΘ(𝗏)𝔞Θ+\nabla\psi_{\Theta}(\mathsf{v})\notin\partial\mathfrak{a}_{\Theta}^{+} and we define the exceptional cone

={𝗏intΘ:𝗏 is not ψΘ-regular}.\displaystyle\mathscr{E}=\{\mathsf{v}\in\operatorname{int}\mathcal{L}_{\Theta}:\mathsf{v}\text{ is not }\psi_{\Theta}\text{-regular}\}.

It follows from the properties of ψΘ\psi_{\Theta} (see Theorem 2.11(2)(4) and Proposition B.1) that \mathscr{E} is the image of 𝔞Θ+\partial\mathfrak{a}_{\Theta}^{+} under a diffeomorphism on 𝔞Θ\mathfrak{a}_{\Theta} and bounds an open strictly convex cone.

We are now ready to state the main theorem regarding exponential mixing of the translation flows associated to ψΘ\psi_{\Theta}-regular 𝗏intΘ\mathsf{v}\in\operatorname{int}\mathcal{L}_{\Theta}. We remark that the translation flows are homothety equivariant, i.e., atc𝗏=act𝗏a^{c\mathsf{v}}_{t}=a^{\mathsf{v}}_{ct} for all tt\in\mathbb{R} and c>0c>0. For α(0,1]\alpha\in(0,1], we denote by C0,α(𝒳)C^{0,\alpha}(\mathcal{X}) the space of α\alpha-Hölder continuous functions on 𝒳\mathcal{X} endowed with the α\alpha-Hölder norm C0,α\|\cdot\|_{C^{0,\alpha}}.

Theorem 1.1.

Let α(0,1]\alpha\in(0,1]. There exist

  • a continuous piecewise-smooth function ηΘ,α:Θ{0}[0,+)\eta_{\Theta,\alpha}:\mathcal{L}_{\Theta}-\{0\}\to[0,+\infty) which is positive except possibly on (Θ){0}(\partial\mathcal{L}_{\Theta}\cup\mathscr{E})-\{0\},

  • a smooth positive function C:intΘC_{\bullet}:\operatorname{int}\mathcal{L}_{\Theta}\to\mathbb{R} (independent of α\alpha),

which are both homothety-invariant, such that for all 𝗏intΘ\mathsf{v}\in\operatorname{int}\mathcal{L}_{\Theta}, and ϕ1,ϕ2C0,α(𝒳)\phi_{1},\phi_{2}\in C^{0,\alpha}(\mathcal{X}), and t>0t>0, we have

|𝒳ϕ1(at𝗏x)ϕ2(x)𝑑m𝒳𝗏(x)m𝒳𝗏(ϕ1)m𝒳𝗏(ϕ2)|C𝗏e𝗏ηΘ,α(𝗏)tϕ1C0,αϕ2C0,α.\displaystyle\left|\int_{\mathcal{X}}\phi_{1}(a^{\mathsf{v}}_{t}x)\phi_{2}(x)\,dm_{\mathcal{X}}^{\mathsf{v}}(x)-m_{\mathcal{X}}^{\mathsf{v}}(\phi_{1})m_{\mathcal{X}}^{\mathsf{v}}(\phi_{2})\right|\leq C_{\mathsf{v}}e^{-\|\mathsf{v}\|\eta_{\Theta,\alpha}(\mathsf{v})t}\|\phi_{1}\|_{C^{0,\alpha}}\|\phi_{2}\|_{C^{0,\alpha}}.
Remark 1.2.

An imprecise version of Theorem 1.1 had been claimed in a talk by the second named author at Institut des Hautes Études Scientifiques in June 2023 as it was well-known to the authors and much of this paper had been written at the time. However, a sticking point delayed the authors in the distribution of the manuscript. See Remark 4.9.

We obtain the following simple corollary when #Θ=1\#\Theta=1 because in this case dim(𝔞Θ)=1\dim(\mathfrak{a}_{\Theta})=1 and so the family of translation flows collapses to a trivial one-dimensional family {{act}t}c>0\{\{a_{ct}\}_{t\in\mathbb{R}}\}_{c>0} coinciding with rescalings of the one-parameter diagonal flow. This theorem already generalizes the case that Γ\Gamma is a projective Anosov subgroup of PSLn()\operatorname{PSL}_{n}(\mathbb{R}) which now follows from the combination of the works of Delarue–Montclair–Sanders [20] and Stoyanov [67] (see the discussion below) and strengthens the local mixing result above. It also generalizes the case that GG is of rank one which gives the geodesic flow for convex cocompact locally symmetric spaces and is contained in the work of Stoyanov [67] (see [17] of the authors for the frame flow).

Theorem 1.3.

Suppose #Θ=1\#\Theta=1. Let α(0,1]\alpha\in(0,1]. Then, there exist ηα>0\eta_{\alpha}>0 and C>0C>0 (independent of α\alpha) such that for all ϕ1,ϕ2C0,α(𝒳)\phi_{1},\phi_{2}\in C^{0,\alpha}(\mathcal{X}) and t>0t>0, we have

|𝒳ϕ1(atx)ϕ2(x)𝑑m𝒳(x)m𝒳(ϕ1)m𝒳(ϕ2)|Ceηαtϕ1C0,αϕ2C0,α.\displaystyle\left|\int_{\mathcal{X}}\phi_{1}(a_{t}x)\phi_{2}(x)\,dm_{\mathcal{X}}(x)-m_{\mathcal{X}}(\phi_{1})m_{\mathcal{X}}(\phi_{2})\right|\leq Ce^{-\eta_{\alpha}t}\|\phi_{1}\|_{C^{0,\alpha}}\|\phi_{2}\|_{C^{0,\alpha}}.

We also obtain the following corollary regarding uniform exponential mixing.

Theorem 1.4.

Let α(0,1]\alpha\in(0,1]. Suppose 𝒦intΘ\mathcal{K}\subset\operatorname{int}\mathcal{L}_{\Theta}-\mathscr{E} is a compact subset. Then, there exist ηα,𝒦>0\eta_{\alpha,\mathcal{K}}>0 and C𝒦>0C_{\mathcal{K}}>0 (independent of α\alpha) such that for all 𝗏𝒦\mathsf{v}\in\mathcal{K}, and ϕ1,ϕ2C0,α(𝒳)\phi_{1},\phi_{2}\in C^{0,\alpha}(\mathcal{X}), and t>0t>0, we have

|𝒳ϕ1(at𝗏x)ϕ2(x)𝑑m𝒳𝗏(x)m𝒳𝗏(ϕ1)m𝒳𝗏(ϕ2)|C𝒦eηα,𝒦tϕ1C0,αϕ2C0,α.\displaystyle\left|\int_{\mathcal{X}}\phi_{1}(a^{\mathsf{v}}_{t}x)\phi_{2}(x)\,dm_{\mathcal{X}}^{\mathsf{v}}(x)-m_{\mathcal{X}}^{\mathsf{v}}(\phi_{1})m_{\mathcal{X}}^{\mathsf{v}}(\phi_{2})\right|\leq C_{\mathcal{K}}e^{-\eta_{\alpha,\mathcal{K}}t}\|\phi_{1}\|_{C^{0,\alpha}}\|\phi_{2}\|_{C^{0,\alpha}}.

Using the spectral bounds of transfer operators in Theorem 4.7, we actually prove the more detailed theorem below from which the above theorems follow. To describe the behavior of the correlation function especially near the boundary of the limit cone Θ\partial\mathcal{L}_{\Theta}, we introduce the map κΘ\kappa_{\Theta}. Equip 𝔞\mathfrak{a} with the inner product and norm induced by the Killing form on 𝔤\mathfrak{g}. Define κΘ:intΘ\kappa_{\Theta}:\operatorname{int}\mathcal{L}_{\Theta}\to\mathbb{R} by

κΘ(𝗏):=ψΘ(𝗏)ψΘ(𝗏)EΘ(ψΘ(𝗏))for all 𝗏intΘ\displaystyle\kappa_{\Theta}(\mathsf{v}):=\frac{\psi_{\Theta}(\mathsf{v})}{\|\nabla\psi_{\Theta}(\mathsf{v})\|}E_{\Theta}(\|\nabla\psi_{\Theta}(\mathsf{v})\|)\qquad\text{for all $\mathsf{v}\in\operatorname{int}\mathcal{L}_{\Theta}$}

where EΘ:>0>0E_{\Theta}:\mathbb{R}_{>0}\to\mathbb{R}_{>0} is the bounded elementary function (in the formal sense) with vanishing limit at ++\infty provided by Theorem 4.1 which can be computed explicitly from the proofs in the paper. Note that κΘ\kappa_{\Theta} is homogeneous of degree 11 because so is ψΘ\psi_{\Theta}. The complex numbers {λk(𝗏)}k=1k𝗏\{\lambda_{k}(\mathsf{v})\}_{k=1}^{k_{\mathsf{v}}} which appear are the Pollicott–Ruelle resonances for the translation flow associated to 𝗏intΘ\mathsf{v}\in\operatorname{int}\mathcal{L}_{\Theta}. The concept of Pollicott–Ruelle resonances originated in [47, 48, 56, 55].

Theorem 1.5.

Let α(0,1]\alpha\in(0,1]. There exist

  • a continuous piecewise-smooth function ηΘ,α:Θ{0}[0,+)\eta_{\Theta,\alpha}:\mathcal{L}_{\Theta}-\{0\}\to[0,+\infty) which is positive except possibly on \mathscr{E},

  • a smooth positive function C:intΘC_{\bullet}:\operatorname{int}\mathcal{L}_{\Theta}\to\mathbb{R} (independent of α\alpha),

such that for all 𝗏intΘ\mathsf{v}\in\operatorname{int}\mathcal{L}_{\Theta}, there exist k𝗏k_{\mathsf{v}}\in\mathbb{N} and

  • a finite set {λk(𝗏)}k=1k𝗏(1,0)×i(1,1)\{\lambda_{k}(\mathsf{v})\}_{k=1}^{k_{\mathsf{v}}}\subset(-1,0)\times i(-1,1) which come in conjugate pairs,

  • a finite set of finite-rank positive bilinear forms {𝗏,k}k=1k𝗏\{\mathcal{B}_{\mathsf{v},k}\}_{k=1}^{k_{\mathsf{v}}},

which are all homothety-invariant, such that for all ϕ1,ϕ2C0,α(𝒳)\phi_{1},\phi_{2}\in C^{0,\alpha}(\mathcal{X}) and t>0t>0, we have

|𝒳ϕ1(at𝗏x)ϕ2(x)𝑑m𝒳𝗏(x)(m𝒳𝗏(ϕ1)m𝒳𝗏(ϕ2)+k=1k𝗏eλk(𝗏)t𝗏,k(ϕ1,ϕ2))|C𝗏eκΘ(𝗏)ηΘ,α(𝗏)tϕ1C0,αϕ2C0,α.\left|\int_{\mathcal{X}}\phi_{1}(a^{\mathsf{v}}_{t}x)\phi_{2}(x)\,dm_{\mathcal{X}}^{\mathsf{v}}(x)-\left(m_{\mathcal{X}}^{\mathsf{v}}(\phi_{1})m_{\mathcal{X}}^{\mathsf{v}}(\phi_{2})+\sum_{k=1}^{k_{\mathsf{v}}}e^{\lambda_{k}(\mathsf{v})t}\mathcal{B}_{\mathsf{v},k}(\phi_{1},\phi_{2})\right)\right|\\ \leq C_{\mathsf{v}}e^{-\kappa_{\Theta}(\mathsf{v})\eta_{\Theta,\alpha}(\mathsf{v})t}\|\phi_{1}\|_{C^{0,\alpha}}\|\phi_{2}\|_{C^{0,\alpha}}.

1.1. Applications

Again using the powerful Theorem 4.7, we also obtain the following applications.

For all 𝗏intΘ\mathsf{v}\in\operatorname{int}\mathcal{L}_{\Theta} and T>0T>0, define

𝒫𝗏:=\displaystyle\mathcal{P}_{\mathsf{v}}:={} {γ:γ is a primitive closed {at𝗏}t-orbit in 𝒳},\displaystyle\{\gamma:\text{$\gamma$ is a primitive closed $\{a^{\mathsf{v}}_{t}\}_{t\in\mathbb{R}}$-orbit in $\mathcal{X}$}\},
𝒫𝗏(T):=\displaystyle\mathcal{P}_{\mathsf{v}}(T):={} {γ𝒫𝗏:𝗏(γ)T},\displaystyle\{\gamma\in\mathcal{P}_{\mathsf{v}}:\ell_{\mathsf{v}}(\gamma)\leq T\},

where 𝗏(γ)\ell_{\mathsf{v}}(\gamma) denotes the primitive period of γ\gamma. We may drop the subscript for the former and write 𝒫\mathcal{P} since the set of closed orbits coincide for all translation flows. Moreover, there is a canonical bijective correspondence

𝒫[Γprim],\displaystyle\mathcal{P}\leftrightarrow[\Gamma_{\mathrm{prim}}],

where the latter is the set of conjugacy classes of Γ\Gamma corresponding to the subset of primitive loxodromic elements ΓprimΓ\Gamma_{\mathrm{prim}}\subset\Gamma. Note that in our setting, all nontorsion elements of Γ\Gamma are loxodromic (see Theorem 2.11). Under the above bijective correspondence, we also have

𝒫𝗏(T){[γ][Γprim]:ψ𝗏(λ(γ))T}\displaystyle\mathcal{P}_{\mathsf{v}}(T)\leftrightarrow\{[\gamma]\in[\Gamma_{\mathrm{prim}}]:\psi_{\mathsf{v}}(\lambda(\gamma))\leq T\}

where we have used the notation ψ𝗏:=δ𝗏1ψΘ(𝗏),\psi_{\mathsf{v}}:=\langle\delta_{\mathsf{v}}^{-1}\nabla\psi_{\Theta}(\mathsf{v}),\cdot\rangle and δ𝗏:=ψΘ(𝗏)\delta_{\mathsf{v}}:=\psi_{\Theta}(\mathsf{v}), which is the topological entropy of the translation flow {at𝗏}t\{a^{\mathsf{v}}_{t}\}_{t\in\mathbb{R}}, from Subsection 2.4.

For all 𝗏intΘ\mathsf{v}\in\operatorname{int}\mathcal{L}_{\Theta}, recall the Selberg zeta function Z𝗏:{ξ:(ξ)>δ𝗏}Z_{\mathsf{v}}:\{\xi\in\mathbb{C}:\Re(\xi)>\delta_{\mathsf{v}}\}\to\mathbb{C} [65] defined by

Z𝗏(ξ)=k=0γ𝒫(1e(ξ+k)𝗏(γ))for all (ξ)>δ𝗏\displaystyle Z_{\mathsf{v}}(\xi)=\prod_{k=0}^{\infty}\prod_{\gamma\in\mathcal{P}}\bigl{(}1-e^{-(\xi+k)\ell_{\mathsf{v}}(\gamma)}\bigr{)}\qquad\text{for all $\Re(\xi)>\delta_{\mathsf{v}}$}

which converges absolutely using the noneffective prime orbit theorem in [16, Eq. (1.8)] (cf. [26, Chapter 2, Definition 4.1] and the subsequent remark). By the work of Pollicott [48], it then extends meromorphically to the domain {ξ:(ξ)>δ𝗏η𝗏}\{\xi\in\mathbb{C}:\Re(\xi)>\delta_{\mathsf{v}}-\eta_{\mathsf{v}}\} for some η𝗏>0\eta_{\mathsf{v}}>0 with a simple zero at δ𝗏\delta_{\mathsf{v}}. He also gives a simple explicit formula for η𝗏\eta_{\mathsf{v}} in terms of the expansion/contraction rate from the Anosov property and the topological entropy from which we see that η𝗏κΘ(𝗏)\eta_{\mathsf{v}}\asymp\kappa_{\Theta}(\mathsf{v}) on some conical neighborhood of Θ\partial\mathcal{L}_{\Theta}.

Theorem 1.6.

There exist a homothety-invariant continuous piecewise-smooth function ηΘ:Θ{0}[0,+)\eta_{\Theta}:\mathcal{L}_{\Theta}-\{0\}\to[0,+\infty) which is positive except possibly on \mathscr{E} such that for all 𝗏intΘ\mathsf{v}\in\operatorname{int}\mathcal{L}_{\Theta}, the Selberg zeta function Z𝗏Z_{\mathsf{v}} has only finitely many zeros on {ξ:(ξ)>δ𝗏κΘ(𝗏)ηΘ(𝗏)}\{\xi\in\mathbb{C}:\Re(\xi)>\delta_{\mathsf{v}}-\kappa_{\Theta}(\mathsf{v})\eta_{\Theta}(\mathsf{v})\}.

We say that the dynamical system (𝒳,m𝒳𝗏,{at𝗏}t)(\mathcal{X},m_{\mathcal{X}}^{\mathsf{v}},\{a^{\mathsf{v}}_{t}\}_{t\in\mathbb{R}}) has an essential spectral gap of κΘ(𝗏)ηΘ(𝗏)>0\kappa_{\Theta}(\mathsf{v})\eta_{\Theta}(\mathsf{v})>0. We emphasize that κΘ\kappa_{\Theta} is an explicit function which dictates the behavior of the decay of the essential spectral gap near the boundary of the limit cone Θ\partial\mathcal{L}_{\Theta} whereas we have no information on the zeros of the Selberg zeta function Z𝗏Z_{\mathsf{v}}.

Remark 1.7.

The above could also be formulated equivalently in terms of the Ruelle zeta function ζ𝗏:{ξ:(ξ)>δ𝗏}\zeta_{\mathsf{v}}:\{\xi\in\mathbb{C}:\Re(\xi)>\delta_{\mathsf{v}}\}\to\mathbb{C} defined by ζ𝗏(ξ):=Z𝗏(ξ+1)Z𝗏(ξ)=γ𝒫(1eξ𝗏(γ))1\zeta_{\mathsf{v}}(\xi):=\frac{Z_{\mathsf{v}}(\xi+1)}{Z_{\mathsf{v}}(\xi)}=\prod_{\gamma\in\mathcal{P}}\bigl{(}1-e^{-\xi\ell_{\mathsf{v}}(\gamma)}\bigr{)}^{-1} for all (ξ)>δ𝗏\Re(\xi)>\delta_{\mathsf{v}} and then extended meromorphically as above.

Recall the offset logarithmic integral function Li:(2,+)\operatorname{Li}:(2,+\infty)\to\mathbb{R} defined by Li(x)=2x1log(t)𝑑t\operatorname{Li}(x)=\int_{2}^{x}\frac{1}{\log(t)}\,dt for all x>2x>2. We have the following prime orbit theorem with a power saving error term.

Theorem 1.8.

There exist

  • a continuous piecewise-smooth function ηΘ:Θ{0}[0,+)\eta_{\Theta}:\mathcal{L}_{\Theta}-\{0\}\to[0,+\infty) which is positive except possibly on (Θ){0}(\partial\mathcal{L}_{\Theta}\cup\mathscr{E})-\{0\},

  • a smooth positive function C:intΘC_{\bullet}:\operatorname{int}\mathcal{L}_{\Theta}\to\mathbb{R},

which are both homothety-invariant, such that for all 𝗏intΘ\mathsf{v}\in\operatorname{int}\mathcal{L}_{\Theta} and T>0T>0, we have

|#𝒫𝗏(T)Li(eδ𝗏T)|C𝗏e(δ𝗏𝗏ηΘ(𝗏))T.\displaystyle\bigl{|}\#\mathcal{P}_{\mathsf{v}}(T)-\operatorname{Li}\bigl{(}e^{\delta_{\mathsf{v}}T}\bigr{)}\bigr{|}\leq C_{\mathsf{v}}e^{(\delta_{\mathsf{v}}-\|\mathsf{v}\|\eta_{\Theta}(\mathsf{v}))T}.

Both the theorems above generalize those in [51, 20] for the case that Γ\Gamma is a projective Anosov subgroup of PSLn()\operatorname{PSL}_{n}(\mathbb{R}) (see the discussion below). The above theorem also effectivizes the prime orbit theorem of Sambarino [59, Theorem 7.8] and its generalization by Chow–Fromm [16, Eq. (1.8)] (cf. [16, Corollary 1.4]). In fact, another approach to proving it is by effectivizing the work of [16] with Theorem 1.1 as input.

1.2. Related works

We discuss the related works of Pollicott–Sharp [51] and Delarue–Montclair–Sanders [20]. Both of these works are in the projective Anosov setting.

In [51], the results are for the case that Γ\Gamma is a Hitchin surface subgroup of PSLn()\operatorname{PSL}_{n}(\mathbb{R}), i.e., the image of a Hitchin representation of a surface group into PSLn()\operatorname{PSL}_{n}(\mathbb{R}), viewed as a projective Anosov subgroup (in which case #Θ=1\#\Theta=1). The authors use a very different kind of coding from that of ours to introduce transfer operators. They then use it to study the Selberg zeta function and obtain both an essential spectral gap and a prime orbit theorem with a power saving error term. This was generalized in [20] to the general projective Anosov setting.

In [20], the results are for the case that Γ\Gamma is a projective Anosov subgroup of PSLn()\operatorname{PSL}_{n}(\mathbb{R}) (in which case #Θ=1\#\Theta=1). The authors obtain a nice result which shows the existence of a Γ\Gamma-invariant and flow-invariant open subset ~Θ(2)×\tilde{\mathcal{M}}\subset\mathcal{F}_{\Theta}^{(2)}\times\mathbb{R} containing ΛΘ(2)×\Lambda_{\Theta}^{(2)}\times\mathbb{R} where the Γ\Gamma-action is properly discontinuous and the translation flow is a contact Axiom A flow on :=Γ\~\mathcal{M}:=\Gamma\backslash\tilde{\mathcal{M}} with 𝒳\mathcal{X} as its basic set [20, Theorem A]. With this construction in hand, exponential mixing follows at once via Stoyanov’s work [67]. They also obtain related theorems already mentioned above. The main common difficulty to both [20] and our paper is that the actions ΓΘ(2)×\Gamma\curvearrowright\mathcal{F}_{\Theta}^{(2)}\times\mathbb{R} which naturally appear are not necessarily properly discontinuous (see Appendix A). This poses an obstruction to using the smooth structure on GG to prove a local non-integrability condition (LNIC) and also to invoking [67]. Delarue–Montclair–Sanders overcome this by constructing ~\tilde{\mathcal{M}}. We take a more direct approach and use weaker properties (which is often advantageous) and do not invoke [67] (see Subsection 1.3 for more details).

Thus, a key point in our paper compared to the aforementioned works is that we handle the vastly more general setting of arbitrary Θ\Theta-Anosov subgroups which turns out to have major differences compared to the setting of projective Anosov subgroups. In the general setting, there is a nontrivial family of translation flows parametrized by vectors in intΘ\operatorname{int}\mathcal{L}_{\Theta} and we discover that we must delete an exceptional set \mathscr{E} for the expected results, i.e., exponential mixing, essential spectral gap, and prime orbit theorem with a power saving error term hold for the generic vectors in intΘ\operatorname{int}\mathcal{L}_{\Theta}-\mathscr{E}. We also obtain the precise nature of the dependence on the vectors in intΘ\operatorname{int}\mathcal{L}_{\Theta}.

1.3. On the proof of Theorem 1.1

For general Θ\Theta-Anosov subgroups, we have a family of translation flows {{at𝗏}t}𝗏intΘ\{\{a^{\mathsf{v}}_{t}\}_{t\in\mathbb{R}}\}_{\mathsf{v}\in\operatorname{int}\mathcal{L}_{\Theta}} on 𝒳\mathcal{X} and we begin by proving that they are metric Anosov. We note that when ΘΠ\Theta\neq\Pi, the space ΛΘ(2)×\Lambda_{\Theta}^{(2)}\times\mathbb{R} does not naturally inherit a Γ\Gamma-invariant metric from the Riemannian metric on GG and so the metric on 𝒳\mathcal{X} needs careful treatment. Moreover, we also show that the strong (un)stable foliations in 𝒳\mathcal{X} of a translation flow are projections of the horospherical foliations of GG (Theorem 3.2). Our proofs of these properties, following [18], use the fact that one can always construct a projective Anosov representation of Γ\Gamma using the Plücker representation and the results of Bridgeman–Canary–Labourie–Sambarino [9] on the metric Anosov property in this case. It is likely that the arguments in [9] can be extended to general Anosov subgroups to give more direct proofs of these properties but we do not do so for expedition reasons. An immediate consequence of the metric Anosov property is that there exist Markov sections for the translation flows which are compatible with the corresponding strong (un)stable foliations. The constructed framework is an important tool which do not appear in previous works and for which reason our approach is fruitful in the general Θ\Theta-Anosov setting.

Thanks to the Markov sections, we may freely introduce symbolic dynamics and thermodynamic formalism, and use transfer operators. More specifically, we define transfer operators ξτ𝗏:C(U,)C(U,)\mathcal{L}_{\xi\tau^{\mathsf{v}}}:C\bigl{(}U,\mathbb{C}\bigr{)}\to C\bigl{(}U,\mathbb{C}\bigr{)} defined by

ξτ𝗏(H)(u)=uσ1(u)eξτ𝗏(u)H(u)\displaystyle\mathcal{L}_{\xi\tau^{\mathsf{v}}}(H)(u)=\sum_{u^{\prime}\in\sigma^{-1}(u)}e^{\xi\tau^{\mathsf{v}}(u^{\prime})}H(u^{\prime})

where τ𝗏:U\tau^{\mathsf{v}}:U\to\mathbb{R} is the first return time map associated to the translation flow {at𝗏}t\{a^{\mathsf{v}}_{t}\}_{t\in\mathbb{R}} on 𝒳\mathcal{X}. We wish to then execute Dolgopyat’s method. Stoyanov [67] has done related work for Axiom A flows but we do not prove this or other strong properties and rely on weaker properties instead. In any case, we need to prove the crucial local non-integrability condition (LNIC) (Proposition 5.11). We wish to use Lie theoretic arguments for this. However, as alluded to in Subsection 1.2, we need to carefully deal with some topological issues due to the fractal nature of the limit set ΛΘ\Lambda_{\Theta} and the Γ\Gamma-action on Θ(2)×\mathcal{F}_{\Theta}^{(2)}\times\mathbb{R} not being properly discontinuous in many cases. Following [64], the relationship between the strong (un)stable foliations and horospherical foliations allows us to lift and extend the rectangles of the Markov section to open sets and use auxiliary smooth extensions of the Poincaré map and the first return time map. We then generalize the techniques in [64, 18] to prove first a smooth version and then a reverse Lipschitz version of LNIC on the original Markov section which is of fractal nature. In the process, it becomes clear that the aforementioned derivation holds for all 𝗏\mathsf{v} but those in a certain exceptional set intΘ\mathscr{E}\subset\operatorname{int}\mathcal{L}_{\Theta}. Fortunately, Dolgopyat’s method is robust enough that the above ingredients suffice to be able to run the classical arguments (for all 𝗏intΘ\mathsf{v}\in\operatorname{int}\mathcal{L}_{\Theta}-\mathscr{E}).

Throughout the paper, we also take a more unified viewpoint and carefully investigate the dependence of the family of translation flows on the parameter 𝗏intΘ\mathsf{v}\in\operatorname{int}\mathcal{L}_{\Theta}. This requires additional important properties. The first one is the existence of a compatible family of Markov sections for all 𝗏intΘ\mathsf{v}\in\operatorname{int}\mathcal{L}_{\Theta}. This allows us to identify the corresponding rectangles with each other and therefore use the same domain UU to define the transfer operators and only vary the first return time maps τ𝗏\tau^{\mathsf{v}} which is smooth in 𝗏intΘ\mathsf{v}\in\operatorname{int}\mathcal{L}_{\Theta}. We also show some bounds for τ𝗏\tau^{\mathsf{v}} which is used to carry through the dependence on 𝗏\mathsf{v} throughout the proofs.

1.4. Organization of the paper

Section 2 covers background on semisimple real algebraic groups, Anosov subgroups, and translation flows. In Section 3, we prove that the translation flow is metric Anosov for a suitable metric and describe the foliations in terms of the Θ\Theta-horospherical subgroups to establish Markov sections compatible with the Lie structure. Section 4 contains the formulation of a key theorem for a Lipschitz version of Dolgopyat’s method from which exponential mixing for Lipschitz continuous functions can be derived. In Section 5 we prove the crucial LNIC. Finally, in Section 6, we prove the aforementioned key theorem. We also include some appendices after that which contain proofs of some basic facts.

Acknowledgements

We thank Hee Oh for her encouragements. We thank Rafael Potrie for useful correspondence regarding metric Anosov flows. We also thank Institut des Hautes Études Scientifiques and the Fields Institute for their hospitality and facilitating conversations. Sarkar acknowledges support by an AMS-Simons Travel Grant.

2. Anosov subgroups and their translation flows

2.1. Lie theoretic preliminaries

Let GG be a connected semisimple real algebraic group with identity element ee and Lie algebra 𝔤\mathfrak{g}. Let B:𝔤×𝔤B:\mathfrak{g}\times\mathfrak{g}\to\mathbb{R} denote the Killing form and θ:𝔤𝔤\theta:\mathfrak{g}\to\mathfrak{g} be a Cartan involution, i.e., the symmetric bilinear form Bθ:𝔤×𝔤B_{\theta}:\mathfrak{g}\times\mathfrak{g}\to\mathbb{R} defined by Bθ(x,y)=B(x,θ(y))B_{\theta}(x,y)=-B(x,\theta(y)) for all x,y𝔤x,y\in\mathfrak{g} is positive definite. We write 1,2=Bθ(1,2)\langle*_{1},*_{2}\rangle=B_{\theta}(*_{1},*_{2}) for the inner product and \|\cdot\| for the induced norm. Let 𝔤=𝔨𝔭\mathfrak{g}=\mathfrak{k}\oplus\mathfrak{p} be the associated eigenspace decomposition corresponding to the eigenvalues +1+1 and 1-1 of θ\theta respectively. Then K=exp(𝔨)<GK=\exp(\mathfrak{k})<G is a maximal compact subgroup. Let 𝔞𝔭\mathfrak{a}\subset\mathfrak{p} be a maximal abelian subalgebra and let A=exp𝔞A=\exp\mathfrak{a}. Identifying 𝔞𝔞\mathfrak{a}\cong\mathfrak{a}^{*} via the Killing form, let Φ𝔞\Phi\subset\mathfrak{a}^{*} be the restricted root system of (𝔤,𝔞)(\mathfrak{g},\mathfrak{a}) and Φ±Φ\Phi^{\pm}\subset\Phi be a choice of sets of positive and negative roots. Let 𝔞+𝔞\mathfrak{a}^{+}\subset\mathfrak{a} be the corresponding closed positive Weyl chamber. We have the associated restricted root space decomposition

𝔤=𝔞𝔪αΦ𝔤α,\displaystyle\mathfrak{g}=\mathfrak{a}\oplus\mathfrak{m}\oplus\bigoplus_{\alpha\in\Phi}\mathfrak{g}_{\alpha},

where 𝔪\mathfrak{m} is the centralizer of 𝔞\mathfrak{a} in 𝔨\mathfrak{k}.

Let ΠΦ+\Pi\subset\Phi^{+} denote the set of all simple roots. We denote by Θ\langle\Theta\rangle the root subsystem generated by any subset ΘΠ\Theta\subset\Pi (it is empty if so is Θ\Theta). Fix a nonempty subset ΘΠ\Theta\subset\Pi. Define Lie subalgebras 𝔞Θ\mathfrak{a}_{\Theta} and 𝔫Θ±\mathfrak{n}_{\Theta}^{\pm} of 𝔤\mathfrak{g} by

𝔞Θ=αΠΘkerα,\displaystyle\mathfrak{a}_{\Theta}=\bigcap_{\alpha\in\Pi-\Theta}\ker\alpha, 𝔫Θ±=αΦ+ΠΘ𝔤α\displaystyle\mathfrak{n}^{\pm}_{\Theta}=\bigoplus_{\alpha\in\Phi^{+}-\langle\Pi-\Theta\rangle}\mathfrak{g}_{\mp\alpha}

where for the former, we take the empty intersection to be 𝔞\mathfrak{a}. Let AΘ=exp(𝔞Θ)A_{\Theta}=\exp(\mathfrak{a}_{\Theta}), 𝔞Θ+=𝔞Θ𝔞+\mathfrak{a}^{+}_{\Theta}=\mathfrak{a}_{\Theta}\cap\mathfrak{a}^{+} and AΘ+=exp(𝔞Θ+)AΘA^{+}_{\Theta}=\exp(\mathfrak{a}^{+}_{\Theta})\subset A_{\Theta}. We denote by int𝔞Θ+\operatorname{int}\mathfrak{a}^{+}_{\Theta} the interior of 𝔞Θ+\mathfrak{a}^{+}_{\Theta} in the topology of 𝔞Θ\mathfrak{a}_{\Theta}. Let 𝒲=NK(A)/ZK(A)\mathcal{W}=N_{K}(A)/Z_{K}(A) be the Weyl group which is a finite group. The adjoint action Ad\operatorname{Ad} induces an action 𝒲𝔞\mathcal{W}\curvearrowright\mathfrak{a} which we also denote by Ad\operatorname{Ad}. Let pΘ:𝔞𝔞Θp_{\Theta}:\mathfrak{a}\to\mathfrak{a}_{\Theta} denote the orthogonal projection map which is in fact the unique projection map invariant under precomposition by all elements of the Weyl group 𝒲\mathcal{W} which act trivially on 𝔞Θ\mathfrak{a}_{\Theta}. Let w0𝒲w_{0}\in\mathcal{W} be the longest element with respect to the generating set consisting of root reflections. Then, Adw0(𝔞+)=𝔞+\operatorname{Ad}_{w_{0}}(\mathfrak{a}^{+})=-\mathfrak{a}^{+} and the map 𝗂=Adw0:𝔞𝔞\mathsf{i}=-\operatorname{Ad}_{w_{0}}:\mathfrak{a}\to\mathfrak{a} is called the opposition involution of 𝔞\mathfrak{a}. The opposition involution preserves 𝔞+\mathfrak{a}^{+} and acts on Π\Pi by precomposition.

For a fixed vint𝔞Θ+v\in\operatorname{int}\mathfrak{a}_{\Theta}^{+}, the expanding/contracting Θ\Theta-horospherical subgroups are

NΘ±\displaystyle N^{\pm}_{\Theta} =exp(𝔫Θ±)={hG:limt±exp(tv)hexp(tv)=e}.\displaystyle=\exp(\mathfrak{n}^{\pm}_{\Theta})=\left\{h\in G:\lim_{t\to\pm\infty}\exp(tv)h\exp(-tv)=e\right\}.

The definition is independent of the choice of vint𝔞Θ+v\in\operatorname{int}\mathfrak{a}_{\Theta}^{+} and NΘ±=w0N𝗂Θw01N_{\Theta}^{\pm}=w_{0}N_{\mathsf{i}\Theta}^{\mp}w_{0}^{-1}.

Let PΘP_{\Theta} denote the standard parabolic subgroup associated to Θ\Theta, i.e., PΘP_{\Theta} is the normalizer of NΘN^{-}_{\Theta} in GG. The Lie algebra of PΘP_{\Theta} is given by 𝔭Θ=𝔞𝔪αΦ+ΠΘ𝔤α\mathfrak{p}_{\Theta}=\mathfrak{a}\oplus\mathfrak{m}\oplus\bigoplus_{\alpha\in\Phi^{+}\cup\langle\Pi-\Theta\rangle}\mathfrak{g}_{\alpha}. We denote the Θ\Theta-Furstenberg boundary of GG by

Θ=G/PΘ.\displaystyle\mathcal{F}_{\Theta}=G/P_{\Theta}.

Note that we also have an action 𝒲Θ\mathcal{W}\curvearrowright\mathcal{F}_{\Theta} induced by the left translation action. For all gGg\in G, we denote

g+\displaystyle g^{+} :=g[PΘ]Θ,\displaystyle:=g[P_{\Theta}]\in\mathcal{F}_{\Theta}, g\displaystyle g^{-} :=gw0[P𝗂Θ]𝗂Θ.\displaystyle:=gw_{0}[P_{\mathsf{i}\Theta}]\in\mathcal{F}_{\mathsf{i}\Theta}.

There is a unique open GG-orbit of Θ×𝗂Θ\mathcal{F}_{\Theta}\times\mathcal{F}_{\mathsf{i}\Theta} given by

Θ(2):=G(e+,e).\displaystyle\mathcal{F}_{\Theta}^{(2)}:=G\cdot(e^{+},e^{-}).

The Levi subgroup LΘL_{\Theta} is the centralizer of AΘA_{\Theta} in GG. It satisfies PΘ=LΘNΘP_{\Theta}=L_{\Theta}N_{\Theta}^{-}, w0P𝗂Θw01=LΘNΘ+w_{0}P_{\mathsf{i}\Theta}w_{0}^{-1}=L_{\Theta}N_{\Theta}^{+}, LΘ=PΘw0P𝗂Θw01L_{\Theta}=P_{\Theta}\cap w_{0}P_{\mathsf{i}\Theta}w_{0}^{-1} and LΘ=AΘSΘL_{\Theta}=A_{\Theta}S_{\Theta} where SΘS_{\Theta} is an almost direct product of a connected semisimple real algebraic group and a compact center. In all of the above and below notations, when Θ=Π\Theta=\Pi, we omit the subscript/superscript Π\Pi.

Definition 2.1 (Iwasawa cocycle).

The Iwasawa cocycle is the map σ:G×𝔞\sigma:G\times\mathcal{F}\rightarrow\mathfrak{a} which assigns to a pair (g,ξ)G×(g,\xi)\in G\times\mathcal{F} the unique element σ(g,ξ)𝔞\sigma(g,\xi)\in\mathfrak{a} such that gkKexp(σ(g,ξ))Ngk\in K\exp(\sigma(g,\xi))N^{-} where kKk\in K such that ξ=kP\xi=kP.

Definition 2.2 (Θ\Theta-Busemann function).

The Θ\Theta-Busemann function is the map βΘ:Θ×G×G𝔞Θ\beta^{\Theta}:\mathcal{F}_{\Theta}\times G\times G\to\mathfrak{a}_{\Theta} defined by

βξΘ(g,h):=pΘ(σ(g1,ξ~)σ(h1,ξ~))\displaystyle\beta^{\Theta}_{\xi}(g,h):=p_{\Theta}(\sigma(g^{-1},\tilde{\xi})-\sigma(h^{-1},\tilde{\xi}))

for all ξΘ\xi\in\mathcal{F}_{\Theta} and g,hGg,h\in G, where ξ~\tilde{\xi}\in\mathcal{F} is any lift of ξ\xi under the natural projection Θ\mathcal{F}\to\mathcal{F}_{\Theta}. This is well-defined, i.e., independent of the choice of ξ~\tilde{\xi} [52, Lemma 6.1], GG-equivariant, and satisfies the identity

βξΘ(g1,g2)=βξΘ(g1,h)+βξΘ(h,g2)\displaystyle\beta^{\Theta}_{\xi}(g_{1},g_{2})=\beta^{\Theta}_{\xi}(g_{1},h)+\beta^{\Theta}_{\xi}(h,g_{2})

for all ξΘ\xi\in\mathcal{F}_{\Theta} and g1,g2,hGg_{1},g_{2},h\in G.

Define a left GG-action on Θ(2)×𝔞Θ\mathcal{F}_{\Theta}^{(2)}\times\mathfrak{a}_{\Theta} by

g(ξ,η,v)=(gξ,gη,v+βgξΘ(e,g))=(gξ,gη,v+pΘ(σ(g,ξ)))\displaystyle g\cdot(\xi,\eta,v)=\bigl{(}g\xi,g\eta,v+\beta^{\Theta}_{g\xi}(e,g)\bigr{)}=(g\xi,g\eta,v+p_{\Theta}(\sigma(g,\xi)))

for all gGg\in G and (ξ,η,v)Θ(2)×𝔞Θ(\xi,\eta,v)\in\mathcal{F}_{\Theta}^{(2)}\times\mathfrak{a}_{\Theta}. Then the stabilizer of (e+,e,0)(e^{+},e^{-},0) is SΘS_{\Theta} and we have the following diffeomorphism.

Definition 2.3 (Θ\Theta-Hopf parameterization).

The Θ\Theta-Hopf parameterization is the left GG-equivariant diffeomorphism G/SΘΘ(2)×𝔞ΘG/S_{\Theta}\to\mathcal{F}_{\Theta}^{(2)}\times\mathfrak{a}_{\Theta} defined by

gSΘ(g+,g,βg+Θ(e,g)).\displaystyle gS_{\Theta}\mapsto\bigl{(}g^{+},g^{-},\beta^{\Theta}_{g^{+}}(e,g)\bigr{)}.

We note that the right AΘA_{\Theta}-action on G/SΘG/S_{\Theta} corresponds to the 𝔞Θ\mathfrak{a}_{\Theta}-translation on the 𝔞Θ\mathfrak{a}_{\Theta}-coordinate of Θ(2)×𝔞Θ\mathcal{F}_{\Theta}^{(2)}\times\mathfrak{a}_{\Theta}.

Definition 2.4 (Θ\Theta-Gromov product).

For (ξ,η)=(g+,g)Θ(2)(\xi,\eta)=(g^{+},g^{-})\in\mathcal{F}_{\Theta}^{(2)}, the Θ\Theta-Gromov product is defined as

[ξ,η]Θ:=βg+Θ(e,g)+𝗂βg𝗂Θ(e,g)\displaystyle[\xi,\eta]_{\Theta}:=\beta^{\Theta}_{g^{+}}(e,g)+\mathsf{i}\beta^{\mathsf{i}\Theta}_{g^{-}}(e,g)

which is independent of choice of gg.

2.2. Zariski dense discrete subgroups

Let Γ<G\Gamma<G be a Zariski dense discrete subgroup.

Definition 2.5 (Θ\Theta-limit set).

The Θ\Theta-limit set ΛΘ\Lambda_{\Theta} is the set of limit points of Γe+\Gamma e^{+} in Θ\mathcal{F}_{\Theta}. It is the unique minimal nonempty closed Γ\Gamma-invariant subset of Θ\mathcal{F}_{\Theta} [52, Corollary 5.2, Lemma 6.3], [40, Lemma 2.13].

The Jordan projection of an element gGg\in G is the unique element λ(g)𝔞+\lambda(g)\in\mathfrak{a}^{+} such that the hyperbolic component of the Jordan decomposition of gg is conjugate to exp(λ(g))\exp(\lambda(g)).

Definition 2.6 (Θ\Theta-limit cone).

The Θ\Theta-limit cone Θ𝔞Θ+\mathcal{L}_{\Theta}\subset\mathfrak{a}_{\Theta}^{+} of Γ\Gamma is the unique minimal closed cone in 𝔞Θ+\mathfrak{a}_{\Theta}^{+} containing pΘ(λ(Γ))p_{\Theta}(\lambda(\Gamma)). It is a convex cone with nonempty interior [3].

The Cartan projection of an element gGg\in G is the unique element μ(g)𝔞+\mu(g)\in\mathfrak{a}^{+} such that gKaμ(g)Kg\in Ka_{\mu(g)}K for all gGg\in G.

Definition 2.7 (Θ\Theta-growth indicator).

The growth indicator ψΘ:𝔞Θ+{}\psi_{\Theta}:\mathfrak{a}^{+}_{\Theta}\to\mathbb{R}\cup\{-\infty\} of Γ\Gamma is defined by

ψΘ(v)=vinfopen cones 𝒞𝔞Θ+v𝒞τ𝒞Θ for all v𝔞Θ\displaystyle\psi_{\Theta}(v)=\|v\|\inf_{\begin{subarray}{c}\text{open cones }\mathcal{C}\subset\mathfrak{a}_{\Theta}^{+}\\ v\in\mathcal{C}\end{subarray}}\tau_{\mathcal{C}}^{\Theta}\qquad\text{ for all }v\in\mathfrak{a}_{\Theta}

with the convention that 0()=00\cdot(-\infty)=0, where \|\cdot\| is any Euclidean norm on 𝔞Θ\mathfrak{a}_{\Theta} and τ𝒞Θ\tau_{\mathcal{C}}^{\Theta} is the abscissa of convergence of the Poincaré series

γΓ,pΘ(μ(γ))𝒞etpΘ(μ(γ)).\displaystyle\sum_{\gamma\in\Gamma,\,p_{\Theta}(\mu(\gamma))\in\mathcal{C}}e^{-t\|p_{\Theta}(\mu(\gamma))\|}.

This definition does not depend on the choice of norm \|\cdot\| (though we have fixed the one induced by the Killing form). Quint [53] showed that ψΘ\psi_{\Theta} is homogeneous of degree 11, concave, upper semicontinuous, and satisfies ψΘ|ext(Θ)=\psi_{\Theta}|_{\operatorname{ext}(\mathcal{L}_{\Theta})}=-\infty, ψΘ|Θ0\psi_{\Theta}|_{\mathcal{L}_{\Theta}}\geq 0, and ψΘ|int(Θ)>0\psi_{\Theta}|_{\operatorname{int}(\mathcal{L}_{\Theta})}>0. Denote by 𝗎ΘintΘ\mathsf{u}_{\Theta}\in\operatorname{int}\mathcal{L}_{\Theta} any maximal growth direction of ψΘ\psi_{\Theta}.

The following generalization of Patterson–Sullivan measures [45, 68] is also due to Quint [52] (see also [1]).

Definition 2.8 ((Γ,ψ)(\Gamma,\psi)-conformal measure).

Let ψ𝔞Θ\psi\in\mathfrak{a}_{\Theta}^{*}. A Borel probability measure ν\nu on Θ\mathcal{F}_{\Theta} is called a (Γ,ψ)(\Gamma,\psi)-conformal measure if

dγνdν(ξ)=eψ(βξΘ(e,γ)) for all γΓ,ξΘ.\displaystyle\frac{d\gamma_{*}\nu}{d\nu}(\xi)=e^{\psi(\beta^{\Theta}_{\xi}(e,\gamma))}\qquad\text{ for all }\gamma\in\Gamma,\,\xi\in\mathcal{F}_{\Theta}.

We say that ψ𝔞Θ\psi\in\mathfrak{a}_{\Theta}^{*} is tangent to ψΘ\psi_{\Theta} at vΘv\in\mathcal{L}_{\Theta} if ψψΘ\psi\geq\psi_{\Theta} and ψ(v)=ψΘ(v)\psi(v)=\psi_{\Theta}(v). We also denote the dual limit cone

Θ:={ψ𝔞Θ:ψ|Θ0}.\displaystyle\mathcal{L}_{\Theta}^{*}:=\{\psi\in\mathfrak{a}_{\Theta}^{*}:\psi|_{\mathcal{L}_{\Theta}}\geq 0\}. (1)
Theorem 2.9 (53, Theorem 8.4).

If ψ𝔞Θ\psi\in\mathfrak{a}_{\Theta}^{*} is tangent to ψΘ\psi_{\Theta} at vint𝔞Θ+Θv\in\operatorname{int}\mathfrak{a}_{\Theta}^{+}\cap\mathcal{L}_{\Theta}, then there exists a (Γ,ψ)(\Gamma,\psi)-conformal measure.

2.3. Anosov subgroups

The notation introduced in the remainder of this section will be fixed throughout the paper. The following definition of Anosov subgroup is due to Guichard–Wienhard [25, Corollary 4.16] (see also [36, 24, 27] for other equivalent characterizations).

Definition 2.10 (Anosov subgroup).

We call a Zariski dense discrete subgroup Γ<G\Gamma<G a (Θ\Theta-)Anosov subgroup if it is a finitely generated Gromov hyperbolic group with Gromov boundary Γ\partial\Gamma and it admits continuous Γ\Gamma-equivariant boundary maps ζΘ:ΓΘ\zeta_{\Theta}:\partial\Gamma\to\mathcal{F}_{\Theta} and ζ𝗂Θ:Γ𝗂Θ\zeta_{\mathsf{i}\Theta}:\partial\Gamma\to\mathcal{F}_{\mathsf{i}\Theta} such that (ζΘ(x),ζ𝗂Θ(y))Θ(2)(\zeta_{\Theta}(x),\zeta_{\mathsf{i}\Theta}(y))\in\mathcal{F}_{\Theta}^{(2)} for all (x,y)Γ(2):={(x,y)Γ2:xy}(x,y)\in\partial\Gamma^{(2)}:=\{(x,y)\in\partial\Gamma^{2}:x\neq y\}.

Let Γ\Gamma be a Zariski dense Θ\Theta-Anosov subgroup with boundary maps ζΘ\zeta_{\Theta} and ζ𝗂Θ\zeta_{\mathsf{i}\Theta}. We record some basic properties in the following theorem. The first 3 properties are [9, Theorem 6.1], [46, Proposition 4.6], [25, Lemma 3.1], respectively, and the last three properties follow from [61, Theorem A] and [58, Lemma 4.8] (cf. [58, Corollary 4.9] and [60, Theorem 4.20]).

Theorem 2.11.

We have the following properties:

  1. ((1))

    the boundary maps ζΘ\zeta_{\Theta} and ζ𝗂Θ\zeta_{\mathsf{i}\Theta} are Hölder homeomorphisms onto the limit sets ΛΘ\Lambda_{\Theta} and Λ𝗂Θ\Lambda_{\mathsf{i}\Theta}, respectively;

  2. ((2))

    the limit cone Θ\mathcal{L}_{\Theta} is contained in int𝔞Θ+{0}\operatorname{int}\mathfrak{a}_{\Theta}^{+}\cup\{0\};

  3. ((3))

    every nontorsion element γΓ\gamma\in\Gamma is loxodromic;

  4. ((4))

    the growth indicator ψΘ\psi_{\Theta} is analytic and strictly concave except along radial directions on intΘ\operatorname{int}\mathcal{L}_{\Theta} and vertically tangent on Θ\partial\mathcal{L}_{\Theta};

  5. ((5))

    if ψ𝔞Θ\psi\in\mathfrak{a}_{\Theta}^{*} is tangent to ψΘ\psi_{\Theta}, then ψ|Θ{0}>0\psi|_{\mathcal{L}_{\Theta}-\{0\}}>0;

  6. ((6))

    if ψintΘ\psi\in\operatorname{int}\mathcal{L}_{\Theta}^{*}, then there exists a unique

    δψ:=lim supt+1tlog#{[γ]:ψ(λ(γ))t}(0,+),\displaystyle\delta_{\psi}:=\limsup_{t\to+\infty}\frac{1}{t}\log\#\{[\gamma]:\psi(\lambda(\gamma))\leq t\}\in(0,+\infty), (2)

    where [γ][\gamma] denotes the conjugacy class of γΓ\gamma\in\Gamma, such that δψψ\delta_{\psi}\psi is tangent to ψΘ\psi_{\Theta}.

For ψ𝔞Θ\psi\in\mathfrak{a}_{\Theta}^{*}, let πψ:Θ(2)×𝔞ΘΘ(2)×\pi_{\psi}:\mathcal{F}_{\Theta}^{(2)}\times\mathfrak{a}_{\Theta}\to\mathcal{F}_{\Theta}^{(2)}\times\mathbb{R} be the projection map defined by

πψ(x,y,v)=(x,y,ψ(v)) for all (x,y,v)Θ(2)×𝔞Θ.\displaystyle\pi_{\psi}(x,y,v)=(x,y,\psi(v))\qquad\text{ for all }(x,y,v)\in\mathcal{F}_{\Theta}^{(2)}\times\mathfrak{a}_{\Theta}. (3)

The Γ\Gamma-action on G/SθΘ(2)×𝔞ΘG/S_{\theta}\cong\mathcal{F}_{\Theta}^{(2)}\times\mathfrak{a}_{\Theta} descends to an action on Θ(2)×\mathcal{F}_{\Theta}^{(2)}\times\mathbb{R} via the projection πψ\pi_{\psi} given explicitly by

γ(x,y,s)\displaystyle\gamma\cdot(x,y,s) =(γx,γy,s+ψ(βγxΘ(e,γ))) for all γΓ,(x,y,s)Θ(2)×.\displaystyle=\bigl{(}\gamma x,\gamma y,s+\psi(\beta^{\Theta}_{\gamma x}(e,\gamma))\bigr{)}\qquad\text{ for all }\gamma\in\Gamma,\,(x,y,s)\in\mathcal{F}_{\Theta}^{(2)}\times\mathbb{R}. (4)

The left Γ\Gamma-action on G/SθΘ(2)×𝔞ΘG/S_{\theta}\cong\mathcal{F}_{\Theta}^{(2)}\times\mathfrak{a}_{\Theta} is not necessarily properly discontinuous. For example, when G=SL3()G=\operatorname{SL}_{3}(\mathbb{R}) and #Θ=1\#\Theta=1, the Γ\Gamma-action is not properly discontinuous (see [4, 35, 19] for a general criterion) but on the other hand, when G=SLn()G=\operatorname{SL}_{n}(\mathbb{R}) for some n4n\geq 4, there are examples where the Γ\Gamma-action is properly discontinuous (see [24, Corollaries 1.10 and 1.11]). The Γ\Gamma-action on Θ(2)×\mathcal{F}_{\Theta}^{(2)}\times\mathbb{R} is also not necessarily properly discontinuous (see Theorem A.2 for a self-contained construction of numerous examples). However, restricting the Γ\Gamma-action to ΛΘ(2)×\Lambda_{\Theta}^{(2)}\times\mathbb{R} where ΛΘ(2):=(ΛΘ×Λ𝗂Θ)Θ(2)\Lambda_{\Theta}^{(2)}:=(\Lambda_{\Theta}\times\Lambda_{\mathsf{i}\Theta})\cap\mathcal{F}_{\Theta}^{(2)}, we have the following theorem.

Theorem 2.12.

If ψintΘ\psi\in\operatorname{int}\mathcal{L}_{\Theta}^{*}, then the Γ\Gamma-action on ΛΘ(2)×\Lambda_{\Theta}^{(2)}\times\mathbb{R} induced by πψ\pi_{\psi} is properly discontinuous and cocompact.

In light of the above theorem, for ψintΘ\psi\in\operatorname{int}\mathcal{L}_{\Theta}^{*}, let

𝒳ψ:=Γ\(ΛΘ(2)×)\displaystyle\mathcal{X}_{\psi}:=\Gamma\backslash\bigl{(}\Lambda_{\Theta}^{(2)}\times\mathbb{R}\bigr{)}

where the Γ\Gamma-action is the one induced by πψ\pi_{\psi}. The space 𝒳ψ\mathcal{X}_{\psi} is equipped with a Bowen–Margulis–Sullivan (BMS) measure m𝒳ψm_{\mathcal{X}}^{\psi}. Let δψ>0\delta_{\psi}>0 be as in Theorem 2.11(6). By [33, Theorem 1.11], there exists a unique (Γ,δψψ)(\Gamma,\delta_{\psi}\psi)-conformal measure νψ\nu_{\psi} (resp. (Γ,δψψ𝗂)(\Gamma,\delta_{\psi}\psi\circ\mathsf{i})-conformal measure νψ𝗂\nu_{\psi\circ\mathsf{i}}) on Θ\mathcal{F}_{\Theta} (resp. 𝗂Θ\mathcal{F}_{\mathsf{i}\Theta}) and moreover, νψ\nu_{\psi} (resp. νψ𝗂\nu_{\psi\circ\mathsf{i}}) is supported on ΛΘ\Lambda_{\Theta} (resp. Λ𝗂Θ\Lambda_{\mathsf{i}\Theta}). We define a locally finite Borel measure m~𝒳ψ\tilde{m}^{\psi}_{\mathcal{X}} on ΛΘ(2)×\Lambda_{\Theta}^{(2)}\times\mathbb{R} by

dm~𝒳ψ(x,y,s)=eδψψ([x,y]Θ)dνψ(x)dνψ𝗂(y)ds for all (x,y,s)ΛΘ(2)×d\tilde{m}^{\psi}_{\mathcal{X}}(x,y,s)=e^{\delta_{\psi}\psi([x,y]_{\Theta})}\,d\nu_{\psi}(x)\,d\nu_{\psi\circ\mathsf{i}}(y)\,ds\qquad\text{ for all }(x,y,s)\in\Lambda_{\Theta}^{(2)}\times\mathbb{R} (5)

where dsds denotes the Lebesgue measure on \mathbb{R}. Observe that m~𝒳ψ\tilde{m}^{\psi}_{\mathcal{X}} is left Γ\Gamma-invariant, so it descends to a probability measure m𝒳ψm_{\mathcal{X}}^{\psi} on 𝒳ψ\mathcal{X}_{\psi} (after renormalization). By [61, Proposition 3.3.2], m𝒳ψm_{\mathcal{X}}^{\psi} is the measure of maximal entropy for the translation flows on 𝒳ψ\mathcal{X}_{\psi}.

A consequence of Theorem 2.12 is that the restriction of the Γ\Gamma-action on Θ(2)×𝔞Θ\mathcal{F}_{\Theta}^{(2)}\times\mathfrak{a}_{\Theta} to ΛΘ(2)×𝔞Θ\Lambda_{\Theta}^{(2)}\times\mathfrak{a}_{\Theta} is properly discontinuous. Moreover, Ω:=Γ\(ΛΘ(2)×𝔞Θ)\Omega:=\Gamma\backslash(\Lambda_{\Theta}^{(2)}\times\mathfrak{a}_{\Theta}) is a trivial kerψ\ker\psi-vector bundle over 𝒳ψ\mathcal{X}_{\psi} by a standard argument.

2.4. Translation flows

Given a vector 𝗏𝔞Θ\mathsf{v}\in\mathfrak{a}_{\Theta}, we have a flow {at𝗏ψ}t\{a^{\psi}_{t\mathsf{v}}\}_{t\in\mathbb{R}} on Ω\Omega which is given by translation by t𝗏t\mathsf{v} in the 𝔞Θ\mathfrak{a}_{\Theta}-coordinate. The flow {at𝗏ψ}\{a^{\psi}_{t\mathsf{v}}\} descends via πψ\pi_{\psi} to what we call a translation flow {atψ,𝗏}t\{a^{\psi,\mathsf{v}}_{t}\}_{t\in\mathbb{R}} on 𝒳ψ\mathcal{X}_{\psi} given explicitly by

atψ,𝗏Γ(x,y,s):=Γ(x,y,s+ψ(𝗏)t) for all t,(x,y,s)ΛΘ(2)×.\displaystyle a^{\psi,\mathsf{v}}_{t}\cdot\Gamma(x,y,s):=\Gamma(x,y,s+\psi({\mathsf{v}})t)\qquad\text{ for all }t\in\mathbb{R},\,(x,y,s)\in\Lambda_{\Theta}^{(2)}\times\mathbb{R}.

By [61], the topological entropy of the flow {atψ,𝗏}t\bigl{\{}a^{\psi,\mathsf{v}}_{t}\bigr{\}}_{t\in\mathbb{R}} is δψψ(𝗏)\delta_{\psi}\psi(\mathsf{v}) (see Eq. 2). Note that only the value of ψ(𝗏)\psi(\mathsf{v}) is required to determine the translation flow on 𝒳ψ\mathcal{X}_{\psi} and furthermore, rescaling ψ\psi simply rescales the \mathbb{R}-coordinate of 𝒳ψ\mathcal{X}_{\psi}. We introduce the following notation to fix a family of translation flows without the redundacies from scaling ψ\psi or 𝗏\mathsf{v}. For 𝗏intΘ\mathsf{v}\in\operatorname{int}\mathcal{L}_{\Theta}, set

ψ𝗏=ψΘ(𝗏),ψΘ(𝗏);\displaystyle\psi_{\mathsf{v}}=\frac{\langle\nabla\psi_{\Theta}(\mathsf{v}),\cdot\rangle}{\psi_{\Theta}(\mathsf{v})}; 𝒳𝗏:=𝒳ψ𝗏;\displaystyle\mathcal{X}_{\mathsf{v}}:=\mathcal{X}_{\psi_{\mathsf{v}}}; at𝗏:=atψ𝗏,𝗏;\displaystyle a^{\mathsf{v}}_{t}:=a^{\psi_{\mathsf{v}},\mathsf{v}}_{t}; m𝒳𝗏:=m𝒳ψ𝗏.\displaystyle m_{\mathcal{X}}^{\mathsf{v}}:=m_{\mathcal{X}}^{\psi_{\mathsf{v}}}. (6)

Note that ψΘ(𝗏)ψ𝗏=ψΘ(𝗏),\psi_{\Theta}(\mathsf{v})\psi_{\mathsf{v}}=\langle\nabla\psi_{\Theta}(\mathsf{v}),\cdot\rangle is the unique linear form tangent to ψΘ\psi_{\Theta} at 𝗏\mathsf{v}. Then, the aforementioned redundancies are removed if 𝗏\mathsf{v} is chosen so that ψ𝗏(𝗏)=1\psi_{\mathsf{v}}(\mathsf{v})=1 and the topological entropy is then δ𝗏:=δψ𝗏=ψΘ(𝗏)\delta_{\mathsf{v}}:=\delta_{\psi_{\mathsf{v}}}=\psi_{\Theta}(\mathsf{v}) by Theorem 2.11(6).

We recall the relation between the translation flow and the (unique up to Hölder reparametrization (see Definition 3.9)) Gromov geodesic flow (𝒢,{at𝒢}t)\bigl{(}\mathcal{G},\bigl{\{}a^{\mathcal{G}}_{t}\bigr{\}}_{t\in\mathbb{R}}\bigr{)} associated to Γ\Gamma.

Theorem 2.13.

For all 𝗏intΘ\mathsf{v}\in\operatorname{int}\mathcal{L}_{\Theta}, there exists a Hölder homemorphism 𝒳𝗏𝒢\mathcal{X}_{\mathsf{v}}\to\mathcal{G} conjugating {at𝗏}t\{a^{\mathsf{v}}_{t}\}_{t\in\mathbb{R}} to a Hölder reparametrization of {at𝒢}t\{a^{\mathcal{G}}_{t}\}_{t\in\mathbb{R}}.

Henceforth, fix 𝗏0intΘ\mathsf{v}_{0}\in\operatorname{int}\mathcal{L}_{\Theta}, say 𝗏0:=𝗎Θ\mathsf{v}_{0}:=\mathsf{u}_{\Theta}. The above theorem in particular shows that for any 𝗏intΘ\mathsf{v}\in\operatorname{int}\mathcal{L}_{\Theta} the dynamical system (𝒳𝗏,{at𝗏}t)(\mathcal{X}_{\mathsf{v}},\{a^{\mathsf{v}}_{t}\}_{t\in\mathbb{R}}) is conjugate to a reparametrization of (𝒳𝗏0,{at𝗏0}t)(\mathcal{X}_{\mathsf{v}_{0}},\{a^{\mathsf{v}_{0}}_{t}\}_{t\in\mathbb{R}}). In fact, we will see that the conjugating homeomorphism and reparametrizations are bi-Lipschitz (see Theorem 3.13). In light of this fact, we take a unified viewpoint where we have a single compact metric space

𝒳:=𝒳𝗏0\displaystyle\mathcal{X}:=\mathcal{X}_{\mathsf{v}_{0}}

equipped with a family of pairs of BMS measures and translation flows

{(m𝒳𝗏,{at𝗏}t)}𝗏intΘ.\displaystyle\{(m_{\mathcal{X}}^{\mathsf{v}},\{a^{\mathsf{v}}_{t}\}_{t\in\mathbb{R}})\}_{\mathsf{v}\in\operatorname{int}\mathcal{L}_{\Theta}}.

To distinguish the family for 𝗏=𝗏𝟢\mathsf{v}=\sf{v_{0}}, we set

ψ:=ψ𝗏0,\displaystyle\psi:=\psi_{\mathsf{v}_{0}}, m𝒳:=m𝒳𝗏0\displaystyle m_{\mathcal{X}}:=m_{\mathcal{X}}^{\mathsf{v}_{0}} {at:=at𝗏0}t.\displaystyle\{a_{t}:=a^{\mathsf{v}_{0}}_{t}\}_{t\in\mathbb{R}}.

The translation flows are then homothety equivariant, i.e., we have

atc𝗏=act𝗏for all t and c>0.\displaystyle a^{c\mathsf{v}}_{t}=a^{\mathsf{v}}_{ct}\qquad\text{for all $t\in\mathbb{R}$ and $c>0$}. (7)

When Θ=Π\Theta=\Pi, Theorems 2.12 and 2.13 follow from the authors’ prequel work [18, Theorem 4.15]. Such theorems first appeared in the work of Sambarino [59, Theorem 3.2] for fundamental groups of compact negatively curved manifolds and in [9, Proposition 4.2] for projective Anosov subgroups whose generalization was outline in [13, Theorem A.2]. The proofs can be generalized for arbitrary Θ\Theta (cf. [61] and see [33, Theorem 9.1] for a complete alternative proof).

3. Anosov property of the translation flow

The purpose of this section is to prove that the translation flow {at}t\{a_{t}\}_{t\in\mathbb{R}} on 𝒳\mathcal{X} is metric Anosov (see Definition 3.7). Before stating the precise theorem, we first define the metric on 𝒳\mathcal{X} we will be using. Fix a bi-invariant Riemannian metric on the compact subgroup KK. Since Θ\mathcal{F}_{\Theta} and 𝗂Θ\mathcal{F}_{\mathsf{i}\Theta} are quotients of KK, the metric on KK along with Euclidean metric on \mathbb{R} induces a product metric dΘ×d𝗂Θ×dd_{\mathcal{F}_{\Theta}}\times d_{\mathcal{F}_{\mathsf{i}\Theta}}\times d_{\mathbb{R}} on Θ×𝗂Θ×\mathcal{F}_{\Theta}\times\mathcal{F}_{\mathsf{i}\Theta}\times\mathbb{R} which we restrict to ΛΘ(2)×\Lambda_{\Theta}^{(2)}\times\mathbb{R}. We call any metric on ΛΘ(2)×\Lambda_{\Theta}^{(2)}\times\mathbb{R} which is locally bi-Lipschitz equivalent to the product metric on ΛΘ(2)×\Lambda_{\Theta}^{(2)}\times\mathbb{R} a locally product-like metric. Construction of a Γ\Gamma-invariant locally product-like metric on 𝒳\mathcal{X} can be done as in [14, Lemma 4.11] (cf. [9, Lemma 5.2]). We equip ΛΘ(2)×\Lambda_{\Theta}^{(2)}\times\mathbb{R} (resp. 𝒳\mathcal{X}) with any Γ\Gamma-invariant locally product-like metric d~\tilde{d} (resp. dd).

Remark 3.1.

When Θ=Π\Theta=\Pi, we have the diffeomorphism G/M(2)×𝔞G/M\cong\mathcal{F}^{(2)}\times\mathfrak{a} where MM is a subgroup of KK. Then using a Γ\Gamma-equivariant continuous section Λ(2)×Λ(2)×𝔞\Lambda^{(2)}\times\mathbb{R}\to\Lambda^{(2)}\times\mathfrak{a}, any left GG-invariant and right KK-invariant Riemannian metric on GG descends to a metric on G/MG/M pulls back to a locally product-like metric on Λ(2)×\Lambda^{(2)}\times\mathbb{R}.

This section is devoted to proving the following theorem.

Theorem 3.2.

The translation flow {at}t\{a_{t}\}_{t\in\mathbb{R}} on (𝒳,d)(\mathcal{X},d) is a metric Anosov flow with respect to the pair of foliations (Wsu,Wss)(W^{\mathrm{su}},W^{\mathrm{ss}}) induced by the Θ\Theta-horospherical foliations (see Subsection 3.2).

We obtain the following as a consequence of Theorems 3.2 and 3.8.

Corollary 3.3.

The translation flow {at}t\{a_{t}\}_{t\in\mathbb{R}} on (𝒳,d)(\mathcal{X},d) has a Markov section with respect to (Wsu,Wss)(W^{\mathrm{su}},W^{\mathrm{ss}}) (see Definition 3.6).

3.1. Metric Anosov flows, Markov sections, and reparametrizations

In this subsection, we briefly recall the definition of metric Anosov flows and the fact that they admit Markov sections due to Pollicott [49]. The reader can also refer to [9, §3.2] and [18, §4.1] for details omitted in this subsection. We also discuss reparametrizations and give a Lipschitz criterion for when a reparametrization of a metric Anosov flow is also metric Anosov.

For this subsection, let {ϕt}t\{\phi_{t}\}_{t\in\mathbb{R}} be an arbitrary continuous flow on an arbitrary metric space (𝒴,d)(\mathcal{Y},d). Given a foliation WW of 𝒴\mathcal{Y}, a point x𝒴x\in\mathcal{Y} and ϵ>0\epsilon>0, let W(x)W(x) denote the leaf through xx and Wϵ(x)={yW(x):d(x,y)<ϵ}W_{\epsilon}(x)=\{y\in W(x):d(x,y)<\epsilon\}. Given a foliation WW transverse to {ϕt}t\{\phi_{t}\}_{t\in\mathbb{R}}, we define the corresponding central foliation WcW^{\mathrm{c}} such that for all x,y𝒴x,y\in\mathcal{Y}, we have yWc(x)y\in W^{\mathrm{c}}(x) if and only if ϕt(y)W(x)\phi_{t}(y)\in W(x) for some tt\in\mathbb{R}.

Definition 3.4 (Local product structure).

A pair of foliations (W,W)(W,W^{\prime}) of 𝒴\mathcal{Y} is said to have local product structure if there exists ϵ>0\epsilon>0 such that for all x𝒴x\in\mathcal{Y}, there exists a homeomorphism [,]:Wϵ(x)×Wϵ(x)Ox[\cdot,\cdot]:W_{\epsilon}(x)\times W^{\prime}_{\epsilon}(x)\to O_{x} where Ox𝒴O_{x}\subset\mathcal{Y} is a neighborhood of xx such that [,]1[\cdot,\cdot]^{-1} is a chart for both WW and WW^{\prime}.

Let (Wsu,Wss)(W^{\mathrm{su}},W^{\mathrm{ss}}) be a pair of foliations transverse to {ϕt}t\{\phi_{t}\}_{t\in\mathbb{R}}. Denote Wwu=(Wsu)cW^{\mathrm{wu}}=(W^{\mathrm{su}})^{\mathrm{c}} and Wws=(Wss)cW^{\mathrm{ws}}=(W^{\mathrm{ss}})^{\mathrm{c}} and suppose (Wwu,Wss)(W^{\mathrm{wu}},W^{\mathrm{ss}}) and (Wws,Wsu)(W^{\mathrm{ws}},W^{\mathrm{su}}) have local product structures given by [,][\cdot,\cdot] defined on Wϵ0ws×Wϵ0suW^{\mathrm{ws}}_{\epsilon_{0}}\times W^{\mathrm{su}}_{\epsilon_{0}} for some ϵ0(0,1)\epsilon_{0}\in(0,1) sufficiently small so that the image sets are always contained in a unit ball. For subsets UWϵ0su(x)U\subset W_{\epsilon_{0}}^{\mathrm{su}}(x) and SWϵ0ss(x)S\subset W_{\epsilon_{0}}^{\mathrm{ss}}(x) such that xU=int(U)¯x\in U=\overline{\operatorname{int}(U)} and xS=int(S)¯x\in S=\overline{\operatorname{int}(S)}, we call

R=[U,S]={[u,s]𝒴:uU,sS}𝒴\displaystyle R=[U,S]=\{[u,s]\in\mathcal{Y}:u\in U,s\in S\}\subset\mathcal{Y}

a rectangle of size δ^\hat{\delta} if diam(R)δ^\operatorname{diam}(R)\leq\hat{\delta} for some δ^>0\hat{\delta}>0, and xx the center of RR. For any rectangle R=[U,S]R=[U,S], we can extend the map [,][\cdot,\cdot] to [,]:RR[\cdot,\cdot]:R\to R defined by [v1,v2]=[u1,s2][v_{1},v_{2}]=[u_{1},s_{2}] for all v1=[u1,s1][U,S]v_{1}=[u_{1},s_{1}]\in[U,S] and v2=[u2,s2][U,S]v_{2}=[u_{2},s_{2}]\in[U,S].

Definition 3.5 (Complete set of rectangles).

A set ={R1,R2,,RN}={[U1,S1],[U2,S2],,[UN,SN]}\mathcal{R}=\{R_{1},R_{2},\dotsc,R_{N}\}=\{[U_{1},S_{1}],[U_{2},S_{2}],\dotsc,[U_{N},S_{N}]\} for some NN\in\mathbb{N} consisting of rectangles of size δ^\hat{\delta} with respect to (Wsu,Wss)(W^{\mathrm{su}},W^{\mathrm{ss}}) in 𝒴\mathcal{Y} is called a complete set of rectangles of size δ^\hat{\delta} if:

  1. ((1))

    RjRk=R_{j}\cap R_{k}=\varnothing for all 1j,kN1\leq j,k\leq N with jkj\neq k;

  2. ((2))

    𝒴=j=1Nt[0,δ^]ϕt(Rj)\mathcal{Y}=\bigcup_{j=1}^{N}\bigcup_{t\in[0,\hat{\delta}]}\phi_{t}(R_{j}).

Let ={R1,R2,,RN}={[U1,S1],[U2,S2],,[UN,SN]}\mathcal{R}=\{R_{1},R_{2},\dotsc,R_{N}\}=\{[U_{1},S_{1}],[U_{2},S_{2}],\dotsc,[U_{N},S_{N}]\} be a complete set of rectangles of size δ^(0,ϵ0)\hat{\delta}\in(0,\epsilon_{0}) in 𝒴\mathcal{Y}. We introduce some notation related to \mathcal{R}. Let

R\displaystyle R =j=1NRj,\displaystyle=\bigsqcup_{j=1}^{N}R_{j}, U\displaystyle U =j=1NUj.\displaystyle=\bigsqcup_{j=1}^{N}U_{j}.

Define the first return time map τ:R\tau:R\to\mathbb{R} by

τ(u)=inf{t>0:ϕt(u)R}for all uR.\displaystyle\tau(u)=\inf\{t\in\mathbb{R}_{>0}:\phi_{t}(u)\in R\}\qquad\text{for all $u\in R$}.

Define

τ¯\displaystyle\overline{\tau} =supuRτ(u),\displaystyle=\sup_{u\in R}\tau(u), τ¯=infuRτ(u).\displaystyle\underline{\tau}=\inf_{u\in R}\tau(u). (8)

Define the Poincaré first return map 𝒫:RR\mathcal{P}:R\to R by

𝒫(u)=ϕτ(u)(u)for all uR.\displaystyle\mathcal{P}(u)=\phi_{\tau(u)}(u)\qquad\text{for all $u\in R$}.

Let σ=(projU𝒫)|U:UU\sigma=(\operatorname{proj}_{U}\circ\mathcal{P})|_{U}:U\to U be its projection where projU:RU\operatorname{proj}_{U}:R\to U is the projection defined by projU([u,s])=u\operatorname{proj}_{U}([u,s])=u for all [u,s]R[u,s]\in R. We define the cores

R^\displaystyle\hat{R} ={uR:𝒫k(u)int(R) for all k},\displaystyle=\{u\in R:\mathcal{P}^{k}(u)\in\operatorname{int}(R)\text{ for all }k\in\mathbb{Z}\},
U^\displaystyle\hat{U} ={uU:σk(u)int(U) for all k0}.\displaystyle=\{u\in U:\sigma^{k}(u)\in\operatorname{int}(U)\text{ for all }k\in\mathbb{Z}_{\geq 0}\}.
Definition 3.6 (Markov section).

We say the complete set of rectangles \mathcal{R} is a Markov section (with respect to (Wsu,Wss)(W^{\mathrm{su}},W^{\mathrm{ss}})) if is satisfies the Markov property: [int(Uk),𝒫(u)]𝒫([int(Uj),u])[\operatorname{int}(U_{k}),\mathcal{P}(u)]\subset\mathcal{P}([\operatorname{int}(U_{j}),u]) and 𝒫([u,int(Sj)])[𝒫(u),int(Sk)]\mathcal{P}([u,\operatorname{int}(S_{j})])\subset[\mathcal{P}(u),\operatorname{int}(S_{k})] for all uRu\in R such that uint(Rj)𝒫1(int(Rk))u\in\operatorname{int}(R_{j})\cap\mathcal{P}^{-1}(\operatorname{int}(R_{k}))\neq\varnothing, for all 1j,kN1\leq j,k\leq N.

Observe that if \mathcal{R} is a Markov section, then τ\tau is constant on [u,Sj][u,S_{j}] for all uUju\in U_{j} and 1jN1\leq j\leq N.

Definition 3.7 (Metric Anosov flow).

The flow {ϕt}t\{\phi_{t}\}_{t\in\mathbb{R}} is said to be metric Anosov (with respect to (Wsu,Wss)(W^{\mathrm{su}},W^{\mathrm{ss}})) if there exist ϵ>0\epsilon>0, η>0\eta>0, and C>0C>0 such that

  1. ((1))

    (Wwu,Wss)(W^{\mathrm{wu}},W^{\mathrm{ss}}) and (Wws,Wsu)(W^{\mathrm{ws}},W^{\mathrm{su}}) have local product structures;

  2. ((2))

    it satisfies the Anosov property: for all x𝒴x\in\mathcal{Y}, yWϵsu(x)y\in W_{\epsilon}^{\mathrm{su}}(x), and zWϵss(x)z\in W_{\epsilon}^{\mathrm{ss}}(x), we have

    d(ϕt(x),ϕt(y))\displaystyle d(\phi_{-t}(x),\phi_{-t}(y)) Ceηtd(x,y),\displaystyle\leq Ce^{-\eta t}d(x,y), d(ϕt(x),ϕt(z))\displaystyle d(\phi_{t}(x),\phi_{t}(z)) Ceηtd(x,z)\displaystyle\leq Ce^{-\eta t}d(x,z)

    for all t0t\geq 0.

The existence of Markov sections for metric Anosov flows is due to Pollicott [49] and generalizes results of Bowen and Ratner [5, 54].

Theorem 3.8.

If {ϕt}t\{\phi_{t}\}_{t\in\mathbb{R}} is a metric Anosov flow with respect to (Wsu,Wss)(W^{\mathrm{su}},W^{\mathrm{ss}}), then there exists a Markov section with respect to (Wsu,Wss)(W^{\mathrm{su}},W^{\mathrm{ss}}) (see Definition 3.6).

We now recall the definition of reparametrization of a flow and give a Lipschitz criterion to verify if a given reparametrization of a metric Anosov flow is also metric Anosov with respect to a given pair of foliations.

Definition 3.9 (Reparametrization).

A flow {ϕ^t}t\{\hat{\phi}_{t}\}_{t\in\mathbb{R}} on 𝒴\mathcal{Y} is called a reparametrization of {ϕt}t\{\phi_{t}\}_{t\in\mathbb{R}} if it is of the form ϕ^t(x)=ϕκ(x,t)(x)\hat{\phi}_{t}(x)=\phi_{\kappa(x,t)}(x) for all x𝒴x\in\mathcal{Y} and tt\in\mathbb{R}, where κ:𝒴×\kappa:\mathcal{Y}\times\mathbb{R}\to\mathbb{R} is a continuous map satisfying

  1. ((1))

    positivity: κ(x,t)>0\kappa(x,t)>0 for all x𝒴x\in\mathcal{Y} and t>0t>0;

  2. ((2))

    the cocycle condition: κ(x,s+t)=κ(ϕ^s(x),t)+κ(x,s)\kappa(x,s+t)=\kappa(\hat{\phi}_{s}(x),t)+\kappa(x,s) for all x𝒴x\in\mathcal{Y} and s,ts,t\in\mathbb{R}.

In that case, ϕ\phi is itself a reparametrization of ϕ^\hat{\phi} for some continuous map κ:𝒴×\kappa^{*}:\mathcal{Y}\times\mathbb{R}\to\mathbb{R} which we call the inverse of κ\kappa and satisfies

  1. ((1))

    (κ)=κ(\kappa^{*})^{*}=\kappa,

  2. ((2))

    κ(x,κ(x,t))=κ(x,κ(x,t))=t\kappa(x,\kappa^{*}(x,t))=\kappa^{*}(x,\kappa(x,t))=t for all (x,t)𝒴×(x,t)\in\mathcal{Y}\times\mathbb{R}.

We say that a reparametrization is Lipschitz (resp. Hölder) if κ(,t)\kappa(\cdot,t) is Lipschitz (resp. Hölder) continuous for all tt\in\mathbb{R}.

Proposition 3.10.

Let {ϕt}t\{\phi_{t}\}_{t\in\mathbb{R}} be a metric Anosov flow with respect to (Wsu,Wss)(W^{\mathrm{su}},W^{\mathrm{ss}}) and with constants (η,C)(\eta,C). Suppose that {ϕ^t:=ϕκ(,t)}t\bigl{\{}\hat{\phi}_{t}:=\phi_{\kappa(\cdot,t)}\bigr{\}}_{t\in\mathbb{R}} is a reparametrization of {ϕt}t\{\phi_{t}\}_{t\in\mathbb{R}} and (W^su,W^ss)(\hat{W}^{\mathrm{su}},\hat{W}^{\mathrm{ss}}) is a pair of foliations transverse to ϕ^\hat{\phi} such that W^suWwu\hat{W}^{\mathrm{su}}\subset W^{\mathrm{wu}} and W^ssWws\hat{W}^{\mathrm{ss}}\subset W^{\mathrm{ws}} leafwise. Suppose there exists ϵ>0\epsilon>0, c1c\geq 1, and c>0c^{\prime}>0 such that

  1. ((1))

    for all x𝒴x\in\mathcal{Y}, yW^ϵsu(x)y\in\hat{W}_{\epsilon}^{\mathrm{su}}(x), and zW^ϵss(x)z\in\hat{W}_{\epsilon}^{\mathrm{ss}}(x), if yϕ(y)Wsu(x)y^{\prime}\in\phi_{\mathbb{R}}(y)\cap W^{\mathrm{su}}(x) and zϕ(z)Wss(x)z^{\prime}\in\phi_{\mathbb{R}}(z)\cap W^{\mathrm{ss}}(x), then

    c1d(x,y)d(x,y)cd(x,y),\displaystyle c^{-1}d(x,y^{\prime})\leq d(x,y)\leq cd(x,y^{\prime}), c1d(x,z)d(x,z)cd(x,z);\displaystyle c^{-1}d(x,z^{\prime})\leq d(x,z)\leq cd(x,z^{\prime});
  2. ((2))

    κ(x,t)ct\kappa(x,t)\geq c^{\prime}t for all t>0t>0 and x𝒴x\in\mathcal{Y}.

Then, ϕ^\hat{\phi} is also metric Anosov with respect to (W^su,W^ss)(\hat{W}^{\mathrm{su}},\hat{W}^{\mathrm{ss}}) and with constants (cη,c2C)(c^{\prime}\eta,c^{2}C).

Proof.

Let {ϕt}t\{\phi_{t}\}_{t\in\mathbb{R}}, {ϕ^t}t\{\hat{\phi}_{t}\}_{t\in\mathbb{R}}, ϵ\epsilon, cc, and cc^{\prime} be as in the proposition. It suffices to check the Anosov property for ϕ^\hat{\phi}. We check it for the stable foliation and the unstable foliation is done similarly. We may assume ϵ>0\epsilon>0 is small enough so that by the Anosov property of ϕ\phi, there exist η>0\eta>0 and C>0C>0 such that for all x𝒴x\in\mathcal{Y}, zWc1ϵss(x)z^{\prime}\in W_{c^{-1}\epsilon}^{\mathrm{ss}}(x), and t>0t>0 we have

d(ϕt(x),ϕt(z))Ceηtd(x,z).\displaystyle d(\phi_{t}(x),\phi_{t}(z^{\prime}))\leq Ce^{-\eta t}d(x,z^{\prime}).

Then, for all x𝒴x\in\mathcal{Y}, zW^ϵss(x)z\in\hat{W}_{\epsilon}^{\mathrm{ss}}(x), t>0t>0, and zϕ(z)Wss(x)z^{\prime}\in\phi_{\mathbb{R}}(z)\cap W^{\mathrm{ss}}(x) we have

d(ϕ^t(x),ϕ^t(z))\displaystyle d(\hat{\phi}_{t}(x),\hat{\phi}_{t}(z)) =d(ϕκ(x,t)(x),ϕκ(z,t)(z))\displaystyle=d(\phi_{\kappa(x,t)}(x),\phi_{\kappa(z,t)}(z))
cd(ϕκ(x,t)(x),ϕκ(x,t)(z))by Property (1)\displaystyle\leq cd(\phi_{\kappa(x,t)}(x),\phi_{\kappa(x,t)}(z^{\prime}))\qquad\text{by \lx@cref{creftype~refnum}{itm:FoliationProperty}}
cCeηκ(x,t)d(x,z)by the Anosov property\displaystyle\leq cCe^{-\eta\kappa(x,t)}d(x,z^{\prime})\qquad\text{by the Anosov property}
c2Cecηtd(x,z)by Properties (1) and (2).\displaystyle\leq c^{2}Ce^{-c^{\prime}\eta t}d(x,z)\qquad\text{by \lx@cref{creftypeplural~refnum}{itm:FoliationProperty} and~\lx@cref{refnum}{itm:KappaProperty}}.

3.2. Stable and unstable foliations for the translation flow

In this subsection, we introduce a pair of natural foliations on 𝒳\mathcal{X} transverse to the translation flow. The right NΘ±N_{\Theta}^{\pm}-orbits in GG give the Θ\Theta-horospherical foliations. Recall that PΘP_{\Theta} is the normalizer of NΘN_{\Theta}^{-} in GG. In particular, SΘLΘ=PΘw0P𝗂Θw01S_{\Theta}\subset L_{\Theta}=P_{\Theta}\cap w_{0}P_{\mathsf{i}\Theta}w_{0}^{-1} normalizes both NΘN_{\Theta}^{-} and NΘ+=w0N𝗂Θw01N_{\Theta}^{+}=w_{0}N_{\mathsf{i}\Theta}^{-}w_{0}^{-1}. Then the Θ\Theta-horospherical foliations descend to foliations of G/SΘΘ(2)×𝔞ΘG/S_{\Theta}\cong\mathcal{F}_{\Theta}^{(2)}\times\mathfrak{a}_{\Theta} which we denote by HΘ,±H^{\mathcal{F}_{\Theta},\pm}, and its restriction to ΛΘ(2)×𝔞Θ\Lambda_{\Theta}^{(2)}\times\mathfrak{a}_{\Theta} by H±H^{\pm}. More explicitly,

HΘ,+(gSΘ):=\displaystyle H^{\mathcal{F}_{\Theta},+}(gS_{\Theta}):= {((gh)+,(gh),β(gh)+Θ(e,gh)):hNΘ+}\displaystyle\,\bigl{\{}\bigl{(}(gh)^{+},(gh)^{-},\beta^{\Theta}_{(gh)^{+}}(e,gh)\bigr{)}\,:\,h\in N_{\Theta}^{+}\bigr{\}}
=\displaystyle= {((gh)+,g,βg+Θ(e,g)+[(gh)+,g]Θ[g+,g]Θ):hNΘ+},\displaystyle\,\bigl{\{}\bigl{(}(gh)^{+},g^{-},\beta^{\Theta}_{g^{+}}(e,g)+[(gh)^{+},g^{-}]_{\Theta}-[g^{+},g^{-}]_{\Theta}\bigr{)}:h\in N_{\Theta}^{+}\bigr{\}},
HΘ,(gSΘ):=\displaystyle H^{\mathcal{F}_{\Theta},-}(gS_{\Theta}):= {((gn)+,(gn),β(gn)+Θ(e,gn)):nNΘ}\displaystyle\,\bigl{\{}\bigl{(}(gn)^{+},(gn)^{-},\beta^{\Theta}_{(gn)^{+}}(e,gn)\bigr{)}\,:\,n\in N_{\Theta}^{-}\bigr{\}}
=\displaystyle= {(g+,(gn),βg+Θ(e,g)):nNΘ},\displaystyle\,\bigl{\{}\bigl{(}g^{+},(gn)^{-},\beta^{\Theta}_{g^{+}}(e,g)\bigr{)}\,:\,n\in N_{\Theta}^{-}\bigr{\}},

(see [32, Lemma 7.4] for the equalities) for all gSΘG/SΘgS_{\Theta}\in G/S_{\Theta}. We denote the image foliations induced by HΘ,±H^{\mathcal{F}_{\Theta},\pm} and H±H^{\pm} under the projection πψ\pi_{\psi} (see Eq. 3) by WΘ,su/ssW^{\mathcal{F}_{\Theta},\mathrm{su}/\mathrm{ss}} and Wsu/ssW^{\mathrm{su}/\mathrm{ss}} and call them the strong unstable and strong stable foliations, respectively. Finally, we use the same notation Wsu/ssW^{\mathrm{su}/\mathrm{ss}} for the image foliation under the vector bundle projection ΛΘ(2)×𝒳\Lambda_{\Theta}^{(2)}\times\mathbb{R}\to\mathcal{X}. The following lemma ensures that the above procedure produces well-defined foliations on Θ(2)×\mathcal{F}_{\Theta}^{(2)}\times\mathbb{R}, ΛΘ(2)×\Lambda_{\Theta}^{(2)}\times\mathbb{R}, and 𝒳\mathcal{X}.

Lemma 3.11.

Let g1,g2Gg_{1},g_{2}\in G such that

(g1+,g1,ψ(βg1+Θ(e,g1)))=(g2+,g2,ψ(βg2+Θ(e,g2))).\displaystyle\bigl{(}g_{1}^{+},g_{1}^{-},\psi\bigl{(}\beta^{\Theta}_{g_{1}^{+}}(e,g_{1})\bigr{)}\bigr{)}=\bigl{(}g_{2}^{+},g_{2}^{-},\psi\bigl{(}\beta^{\Theta}_{g_{2}^{+}}(e,g_{2})\bigr{)}\bigr{)}.

Then, for all h1,h2NΘ+h_{1},h_{2}\in N_{\Theta}^{+} with (g1h1)+=(g2h2)+(g_{1}h_{1})^{+}=(g_{2}h_{2})^{+}, we have

ψ(β(g1h1)+Θ(e,g1h1))=ψ(β(g2h2)+Θ(e,g2h2)).\displaystyle\psi\bigl{(}\beta^{\Theta}_{(g_{1}h_{1})^{+}}(e,g_{1}h_{1})\bigr{)}=\psi\bigl{(}\beta^{\Theta}_{(g_{2}h_{2})^{+}}(e,g_{2}h_{2})\bigr{)}.
Proof.

Let g1,g2Gg_{1},g_{2}\in G as in the lemma and suppose h1,h2NΘ+h_{1},h_{2}\in N_{\Theta}^{+} with (g1h1)+=(g2h2)+(g_{1}h_{1})^{+}=(g_{2}h_{2})^{+}. Since (g1+,g1)=(g2+,g2)(g_{1}^{+},g_{1}^{-})=(g_{2}^{+},g_{2}^{-}), there exists arAΘSΘ=LΘar\in A_{\Theta}S_{\Theta}=L_{\Theta} such that g1=g2arg_{1}=g_{2}ar. Then, by an elementary computation, we have

βg1+Θ(e,g1)=βg2+Θ(e,g2)+a.\displaystyle\beta^{\Theta}_{g_{1}^{+}}(e,g_{1})=\beta^{\Theta}_{g_{2}^{+}}(e,g_{2})+a.

Since ψ(βg1+Θ(e,g1))=ψ(βg2+Θ(e,g2))\psi\bigl{(}\beta^{\Theta}_{g_{1}^{+}}(e,g_{1})\bigr{)}=\psi\bigl{(}\beta^{\Theta}_{g_{2}^{+}}(e,g_{2})\bigr{)} by hypothesis, we have

ψ(a)=0.\displaystyle\psi(a)=0.

Since ((g1h1)+,(g1h1))=((g1h1)+,g1)=((g2h2)+,g2)=((g2h2)+,(g2h2))((g_{1}h_{1})^{+},(g_{1}h_{1})^{-})=((g_{1}h_{1})^{+},g_{1}^{-})=((g_{2}h_{2})^{+},g_{2}^{-})=((g_{2}h_{2})^{+},(g_{2}h_{2})^{-}), we also have g1h1=g2h2arg_{1}h_{1}=g_{2}h_{2}a^{\prime}r^{\prime} for some arAΘSΘa^{\prime}r^{\prime}\in A_{\Theta}S_{\Theta}. Combining with g1=g2arg_{1}=g_{2}ar, we obtain arh1=h2ar=ar((ar)1h2(ar))arh_{1}=h_{2}a^{\prime}r^{\prime}=a^{\prime}r^{\prime}((a^{\prime}r^{\prime})^{-1}h_{2}(a^{\prime}r^{\prime})). Since LΘL_{\Theta} normalizes NΘ+N_{\Theta}^{+} and LΘNΘ+L_{\Theta}N_{\Theta}^{+} is a direct product, it follows that

a=a.\displaystyle a^{\prime}=a.

Similar to before, we have

β(g1h1)+Θ(e,g1h1)=β(g2h2)+Θ(e,g2h2)+a=β(g2h2)+Θ(e,g2h2)+a\displaystyle\beta^{\Theta}_{(g_{1}h_{1})^{+}}(e,g_{1}h_{1})=\beta^{\Theta}_{(g_{2}h_{2})^{+}}(e,g_{2}h_{2})+a^{\prime}=\beta^{\Theta}_{(g_{2}h_{2})^{+}}(e,g_{2}h_{2})+a

and using the hypothesis ψ(β(g1h1)+Θ(e,g1))=ψ(β(g2h2)+Θ(e,g2))\psi\bigl{(}\beta^{\Theta}_{(g_{1}h_{1})^{+}}(e,g_{1})\bigr{)}=\psi\bigl{(}\beta^{\Theta}_{(g_{2}h_{2})^{+}}(e,g_{2})\bigr{)}, we conclude that ψ(β(g1h1)+Θ(e,g1h1))=ψ(β(g2h2)+Θ(e,g2h2))\psi\bigl{(}\beta^{\Theta}_{(g_{1}h_{1})^{+}}(e,g_{1}h_{1})\bigr{)}=\psi\bigl{(}\beta^{\Theta}_{(g_{2}h_{2})^{+}}(e,g_{2}h_{2})\bigr{)}. ∎

Hence, we obtain (weak/strong) (stable/unstable) foliations on 𝒳\mathcal{X} as follows: for all z=Γ(x,y,s)𝒳z=\Gamma(x,y,s)\in\mathcal{X}, we have

Wsu(z)={Γ((gh)+,g,s+ψ([(gh)+,g]Θ[g+,g]Θ)):hNΘ+,(gh)+ΛΘ},Wss(z)={Γ(g+,(gn),s):nNΘ,(gn)Λ𝗂Θ},Wwu(z)=tatWss,Wws(z)=tatWss\displaystyle\begin{aligned} W^{\mathrm{su}}(z)&=\bigl{\{}\Gamma\bigl{(}(gh)^{+},g^{-},s+\psi\bigl{(}[(gh)^{+},g^{-}]_{\Theta}-[g^{+},g^{-}]_{\Theta}\bigr{)}\bigr{)}:h\in N_{\Theta}^{+},(gh)^{+}\in\Lambda_{\Theta}\bigr{\}},\\ W^{\mathrm{ss}}(z)&=\bigl{\{}\Gamma\bigl{(}g^{+},(gn)^{-},s\bigr{)}:n\in N_{\Theta}^{-},(gn)^{-}\in\Lambda_{\mathsf{i}\Theta}\bigr{\}},\\ W^{\mathrm{wu}}(z)&=\bigcup_{t\in\mathbb{R}}a_{t}W^{\mathrm{ss}},\\ W^{\mathrm{ws}}(z)&=\bigcup_{t\in\mathbb{R}}a_{t}W^{\mathrm{ss}}\end{aligned} (9)

for any choice of gGg\in G such that z=(g+,g,ψ(βg+Θ(e,g)))z=\bigl{(}g^{+},g^{-},\psi\bigl{(}\beta^{\Theta}_{g^{+}}(e,g)\bigr{)}\bigr{)}. It is easy to see that WssW^{\mathrm{ss}} and WsuW^{\mathrm{su}} are transverse to the translation flow, and (Wss,Wwu)(W^{\mathrm{ss}},W^{\mathrm{wu}}) and (Wsu,Wws)(W^{\mathrm{su}},W^{\mathrm{ws}}) have local product structures.

3.3. Projective Anosov representations

In this subsection, we recall from [9] facts about projective Anosov representations that we will need for the proof of Theorem 3.2. Let 𝔭Θ\mathfrak{p}_{\Theta} denote the Lie algebra of PΘP_{\Theta}. Then the adjoint representation Ad:GAut(𝔤)\operatorname{Ad}:G\to\operatorname{Aut}(\mathfrak{g}) induces the representation ρ0:GPSL(dim𝔭Θ𝔤)\rho_{0}:G\to\operatorname{PSL}\bigl{(}\bigwedge^{\dim\mathfrak{p}_{\Theta}}\mathfrak{g}\bigr{)}. Define the vector space V:=ρ0(G)(dim𝔭Θ𝔭Θ)V:=\rho_{0}(G)\bigl{(}\bigwedge^{\dim\mathfrak{p}_{\Theta}}\mathfrak{p}_{\Theta}\bigr{)} (one can see that VV is a vector space using the Iwasawa decomposition G=KANG=KAN and the fact that ,\langle\cdot,\cdot\rangle is AdK\operatorname{Ad}_{K}-invariant). Then, we have the induced irreducible representation ρ:GPSL(V)\rho:G\to\operatorname{PSL}(V), called the Plücker representation, and it induces diffeomorphisms Θ(V)\mathcal{F}_{\Theta}\to\mathbb{P}(V) and 𝗂Θ(V)\mathcal{F}_{\mathsf{i}\Theta}\to\mathbb{P}(V^{*}) [38, Theorem 7.25]. By [25, Proposition 4.3]), the restriction ρ|Γ:ΓPSL(V)\rho|_{\Gamma}:\Gamma\to\operatorname{PSL}(V) is a projective Anosov representation in the sense of [9, Definition 2.2] and in particular, the previous diffeomorphisms restrict to Γ\Gamma-equivariant bi-Lipschitz maps ζρ:ΛΘ(V)\zeta_{\rho}:\Lambda_{\Theta}\to\mathbb{P}(V) and ζρ:Λ𝗂Θ(V)\zeta_{\rho}^{*}:\Lambda_{\mathsf{i}\Theta}\to\mathbb{P}(V^{*}).

The \mathbb{R}-bundle

𝒢~ρ:={(x,y,(v,Ψ)):(x,y)ΛΘ(2),(v,Ψ)ζρ(x)×ζρ(y),Ψ(v)=1}/\displaystyle\tilde{\mathcal{G}}_{\rho}:=\bigl{\{}(x,y,(v,\Psi)):(x,y)\in\Lambda_{\Theta}^{(2)},(v,\Psi)\in\zeta_{\rho}(x)\times\zeta_{\rho}^{*}(y),\Psi(v)=1\bigr{\}}/{\sim}

over ΛΘ(2)\Lambda_{\Theta}^{(2)}, where (v,Ψ)(v,Ψ)(v,\Psi)\sim(-v,-\Psi) is equipped with a Γ\Gamma-action and a flow {a~ρ,t}t\{\tilde{a}_{\rho,t}\}_{t\in\mathbb{R}} that commute with each other which are given as follows: for all (x,y,(v,Ψ))𝒢~ρ(x,y,(v,\Psi))\in\tilde{\mathcal{G}}_{\rho}, γΓ\gamma\in\Gamma, and tt\in\mathbb{R}, let

γ(x,y,(v,Ψ))\displaystyle\gamma(x,y,(v,\Psi)) :=(γx,γy,(ρ(γ)v,Ψρ(γ)1));\displaystyle:=(\gamma x,\gamma y,(\rho(\gamma)v,\Psi\circ\rho(\gamma)^{-1}));
a~ρ,t(x,y,(v,Ψ))\displaystyle\tilde{a}_{\rho,t}(x,y,(v,\Psi)) :=(x,y,(etv,etΨ)).\displaystyle:=(x,y,(e^{t}v,e^{-t}\Psi)).

Let 𝒢ρ:=Γ\𝒢~ρ\mathcal{G}_{\rho}:=\Gamma\backslash\tilde{\mathcal{G}}_{\rho}. Then {a~ρ,t}t\{\tilde{a}_{\rho,t}\}_{t\in\mathbb{R}} descends to a flow {aρ,t}t\{a_{\rho,t}\}_{t\in\mathbb{R}} on 𝒢ρ\mathcal{G}_{\rho} called the ρ\rho-geodesic flow.

Any Euclidean metric on VV induces a metric on (V)×(V)×(V×V)\mathbb{P}(V)\times\mathbb{P}(V^{*})\times(V\times V^{*}) and this further induces a metric on 𝒢~ρ\tilde{\mathcal{G}}_{\rho} in a natural way. Any metric on 𝒢~ρ\tilde{\mathcal{G}}_{\rho} obtained in this fashion is called a linear metric.

Theorem 3.12 (9, Proposition 5.7).

There exists a Γ\Gamma-invariant metric d~ρ\tilde{d}_{\rho} on 𝒢~ρ\tilde{\mathcal{G}}_{\rho}, which is bi-Lipschitz equivalent to any linear metric, such that it descends to a metric dρd_{\rho} on 𝒢ρ\mathcal{G}_{\rho} for which the flow {aρ,t}t\{a_{\rho,t}\}_{t\in\mathbb{R}} is metric Anosov with respect to the pair of foliations (Wρsu,Wρss)\bigl{(}W^{\mathrm{su}}_{\rho},W^{\mathrm{ss}}_{\rho}\bigr{)} which are defined as follows: for all z=Γ(x,y,(v,Ψ))𝒢ρz=\Gamma(x,y,(v,\Psi))\in\mathcal{G}_{\rho},

Wρsu(z)\displaystyle W^{\mathrm{su}}_{\rho}(z) :={Γ(x,y,(v,Ψ)):(x,y)ΛΘ(2),(v,Ψ)ζρ(x)×ζρ(y),Ψ(v)=1};\displaystyle:=\bigl{\{}\Gamma(x^{\prime},y,(v^{\prime},\Psi)):(x^{\prime},y)\in\Lambda_{\Theta}^{(2)},(v^{\prime},\Psi)\in\zeta_{\rho}(x^{\prime})\times\zeta_{\rho}^{*}(y),\Psi(v^{\prime})=1\bigr{\}};
Wρss(z)\displaystyle W^{\mathrm{ss}}_{\rho}(z) :={Γ(x,y,(v,Ψ)):(x,y)ΛΘ(2),(v,Ψ)ζρ(x)×ζρ(y),Ψ(v)=1}.\displaystyle:=\bigl{\{}\Gamma(x,y^{\prime},(v,\Psi^{\prime})):(x,y^{\prime})\in\Lambda_{\Theta}^{(2)},(v,\Psi^{\prime})\in\zeta_{\rho}(x)\times\zeta_{\rho}^{*}(y^{\prime}),\Psi^{\prime}(v)=1\bigr{\}}.

3.4. Proof of Theorem 3.2

Theorem 3.13.

There exists a bi-Lipschitz homeomorphism :(𝒢ρ,dρ)(𝒳,d)\mathcal{H}:(\mathcal{G}_{\rho},d_{\rho})\to(\mathcal{X},d) which conjugates the translation flow {at}t\{a_{t}\}_{t\in\mathbb{R}} to a Lipschitz reparametrization{aρ,κ(,t)}t\{a_{\rho,\kappa(\cdot,t)}\}_{t\in\mathbb{R}} of the ρ\rho-geodesic flow {aρ,t}t\{a_{\rho,t}\}_{t\in\mathbb{R}}.

Proof.

We proceed in the following two steps.

Step 1: Construction of \mathcal{H}. We only give an overview (omitting details) of the construction of \mathcal{H} as it is similar to the full construction in [18, Theorem 4.15] which deals with the case Θ=Π\Theta=\Pi. Consider the trivial \mathbb{R}-bundle

~:=𝒢~ρ×>0\displaystyle\tilde{\mathcal{I}}:=\tilde{\mathcal{G}}_{\rho}\times\mathbb{R}_{>0}

equipped with a left Γ\Gamma-action defined by

γ(z,r):=(γz,reψ(βxΘ(γ1,e))) for all γΓ and (z,r)~\displaystyle\gamma\cdot(z,r):=\bigl{(}\gamma z,re^{\psi(\beta^{\Theta}_{x}(\gamma^{-1},e))}\bigr{)}\qquad\text{ for all }\gamma\in\Gamma\text{ and }(z,r)\in\tilde{\mathcal{I}}

and a flow {ϕ~t}t\{\tilde{\phi}^{\mathcal{I}}_{t}\}_{t\in\mathbb{R}} defined by

ϕ~t(z,r):=(a~ρ,tz,r) for all γΓ,(z,r)~, and t\displaystyle\tilde{\phi}^{\mathcal{I}}_{t}(z,r):=(\tilde{a}_{\rho,t}z,r)\qquad\text{ for all }\gamma\in\Gamma,(z,r)\in\tilde{\mathcal{I}},\text{ and }t\in\mathbb{R}

which commutes with the Γ\Gamma-action.

Let 𝒰={γ𝒰j}γΓ, 1jj0\mathcal{U}=\{\gamma\mathcal{U}_{j}\}_{\gamma\in\Gamma,\,1\leq j\leq j_{0}} for some j0j_{0}\in\mathbb{N} be a locally finite open cover of 𝒢~ρ\tilde{\mathcal{G}}_{\rho} such that γ𝒰jγ𝒰j=\gamma\mathcal{U}_{j}\cap\gamma^{\prime}\mathcal{U}_{j}=\varnothing for all distinct γ,γΓ\gamma,\gamma^{\prime}\in\Gamma and 1jj01\leq j\leq j_{0}. Using a partition of unity {φi}iI\{\varphi_{i}\}_{i\in I}, for some index set II, subordinate to 𝒰\mathcal{U} such that φi:ΛΘ(2)×[0,+)\varphi_{i}:\Lambda_{\Theta}^{(2)}\times\mathbb{R}\to[0,+\infty) is smooth along the ρ\rho-geodesic flow for all iIi\in I, we can construct a Γ\Gamma-equivariant section u0:𝒢~ρ~u_{0}:\tilde{\mathcal{G}}_{\rho}\to\tilde{\mathcal{I}} of the form u0(z):=(z,u^0(z))u_{0}(z):=(z,\hat{u}_{0}(z)). Explicitly, we can take

logu^0(z)=iI,suppφiγ𝒰jψ(βxΘ(e,γ))φi(z) for all z=(x,y,(v,Ψ))𝒢~ρ.\displaystyle\log\hat{u}_{0}(z)=\sum_{i\in I,\,\operatorname{supp}\varphi_{i}\subset\gamma\mathcal{U}_{j}}\psi\bigl{(}\beta^{\Theta}_{x}(e,\gamma)\bigr{)}\cdot\varphi_{i}(z)\qquad\text{ for all }z=(x,y,(v,\Psi))\in\tilde{\mathcal{G}}_{\rho}.

Let ~:=𝒢~ρ×>0\tilde{\mathcal{L}}:=\tilde{\mathcal{G}}_{\rho}\times\mathbb{R}_{>0} denote the image of ~\tilde{\mathcal{I}} under the exponential map on the \mathbb{R}-coordinate. Equip ~\tilde{\mathcal{L}} with the Γ\Gamma-action and flow {ϕ~t}t\{\tilde{\phi}^{\mathcal{L}}_{t}\}_{t\in\mathbb{R}} it inherits from ~\tilde{\mathcal{I}} and the unique Γ\Gamma-invariant bundle norm 0~\|\cdot\|^{\tilde{\mathcal{L}}}_{0} satisfying u0(z)0~=1\|u_{0}(z)\|_{0}^{\tilde{\mathcal{L}}}=1 for all z𝒢~ρz\in\tilde{\mathcal{G}}_{\rho}. Then, :=Γ\~\mathcal{L}:=\Gamma\backslash\tilde{\mathcal{I}} is an >0\mathbb{R}_{>0}-bundle over 𝒢ρ\mathcal{G}_{\rho}, {ϕ~t}t\{\tilde{\phi}^{\mathcal{L}}_{t}\}_{t\in\mathbb{R}} descends to a flow {ϕt}t\{\phi^{\mathcal{L}}_{t}\}_{t\in\mathbb{R}} on \mathcal{L}, and 0~\|\cdot\|_{0}^{\tilde{\mathcal{L}}} descends to a norm 0\|\cdot\|_{0}^{\mathcal{L}} on \mathcal{L}. Using the Morse property [21, Theorem 4.13] of Kapovich–Leeb–Porti [27, Proposition 5.16] and an analogue of Sullivan’s Shadow Lemma [32, Lemma 3.1], one can show that {ϕt}t\{\phi^{\mathcal{L}}_{t}\}_{t\in\mathbb{R}} is discretely contracting with respect to 0\|\cdot\|_{0}^{\mathcal{L}}, i.e., there exists T>0T>0 and η>0\eta>0 such that ϕT()0eη0\|\phi_{T}^{\mathcal{L}}(\ell)\|_{0}^{\mathcal{L}}\leq e^{-\eta}\|\ell\|_{0}^{\mathcal{L}} for all \ell\in\mathcal{L}. By [9, Lemma 4.3], the norm ~=0Tes/Tϕs()0~ds\|\cdot\|^{\tilde{\mathcal{L}}}=\int_{0}^{T}e^{s/T}\|\phi_{s}^{\mathcal{L}}(\cdot)\|_{0}^{\tilde{\mathcal{L}}}\,ds on ~\tilde{\mathcal{L}} descends to a norm \|\cdot\|^{\mathcal{L}} on \mathcal{L} such that {ϕt}t\{\phi_{t}^{\mathcal{L}}\}_{t\in\mathbb{R}} is uniformly contracting with respect to \|\cdot\|^{\mathcal{L}}. Note that (z,u^(z))~=1\|(z,\hat{u}(z))\|^{\tilde{\mathcal{L}}}=1 where

1u^(z)=0Tes/Tu^0(a~ρ,s(z))𝑑s for all z𝒢~ρ.\displaystyle\frac{1}{\hat{u}(z)}=\int_{0}^{T}\frac{e^{s/T}}{\hat{u}_{0}(\tilde{a}_{\rho,s}(z))}\,ds\qquad\text{ for all }z\in\tilde{\mathcal{G}}_{\rho}.

Let ~:𝒢~ρΛΘ(2)×\tilde{\mathcal{H}}:\tilde{\mathcal{G}}_{\rho}\to\Lambda_{\Theta}^{(2)}\times\mathbb{R} be the Γ\Gamma-equivariant map defined by

~(z)=(x,y,logu^(z)) for all z=(x,y,(v,Ψ))𝒢~ρ.\displaystyle\tilde{\mathcal{H}}(z)=(x,y,\log\hat{u}(z))\qquad\text{ for all }z=(x,y,(v,\Psi))\in\tilde{\mathcal{G}}_{\rho}.

Then, it can be shown as in [18, Theorem 4.15] that u^\hat{u} is locally Lipschitz and ~\tilde{\mathcal{H}} descends to a Lipschitz homeomorphism :𝒢ρ𝒳\mathcal{H}:\mathcal{G}_{\rho}\to\mathcal{X} and the reparametrization is Lipschitz. (Here, \mathcal{H} is Lipschitz instead of Hölder as in [18] since the induced maps Θ(V)\mathcal{F}_{\Theta}\to\mathbb{P}(V) and 𝗂Θ(V)\mathcal{F}_{\mathsf{i}\Theta}\to\mathbb{P}(V^{*}) are Lipschitz.)

Step 2: \mathcal{H} is bi-Lipschitz. By the above, it suffices to show that ~1\tilde{\mathcal{H}}^{-1} is locally Lipschitz with respect to d~\tilde{d} and d~ρ\tilde{d}_{\rho}. Recall that d~\tilde{d} is a locally bi-Lipschitz equivalent to the product metric on ΛΘ(2)×\Lambda_{\Theta}^{(2)}\times\mathbb{R} and d~ρ\tilde{d}_{\rho} is locally bi-Lipschitz equivalent to any linear metric on 𝒢~ρ\tilde{\mathcal{G}}_{\rho}. Thus, it suffices to show that ~1\tilde{\mathcal{H}}^{-1} is locally Lipschitz with respect to the product metric on ΛΘ(2)×\Lambda_{\Theta}^{(2)}\times\mathbb{R} and any linear metric on 𝒢~ρ\tilde{\mathcal{G}}_{\rho}. Fix a Euclidean metric dEd_{\mathrm{E}} on VV. We also use dEd_{\mathrm{E}} to denote the induced metrics on VV^{*} and V×VV\times V^{*}. Fix a compact subset 𝒦𝒢~ρ\mathcal{K}\subset\tilde{\mathcal{G}}_{\rho} that is “radially symmetric” in the sense that if (xi,yi,(vi,Ψi))𝒦(x_{i},y_{i},(v_{i},\Psi_{i}))\in\mathcal{K} for i{1,2}i\in\{1,2\}, then p(z1,z2):=(x1,y1,(v2v1v1,v1v2Ψ1))𝒦p(z_{1},z_{2}):=\Bigl{(}x_{1},y_{1},\Bigl{(}\tfrac{\|v_{2}\|}{\|v_{1}\|}v_{1},\tfrac{\|v_{1}\|}{\|v_{2}\|}\Psi_{1}\Bigr{)}\Bigr{)}\in\mathcal{K} as well. We need to show that there exists c1>0c_{1}>0 such that for any zi=(xi,yi,(vi,Ψi))𝒦z_{i}=(x_{i},y_{i},(v_{i},\Psi_{i}))\in\mathcal{K} for i{1,2}i\in\{1,2\}, we have

dE((v1,Ψ1),(v2,Ψ2))c1((dΘ×d𝗂Θ)((x1,y1),(x2,y2))+|logu^(z1)logu^(z2)|).\displaystyle d_{\mathrm{E}}((v_{1},\Psi_{1}),(v_{2},\Psi_{2}))\leq c_{1}((d_{\mathcal{F}_{\Theta}}\times d_{\mathcal{F}_{\mathsf{i}\Theta}})((x_{1},y_{1}),(x_{2},y_{2}))+|\log\hat{u}(z_{1})-\log\hat{u}(z_{2})|).

For convenience, let t(z1,z2):=log(v2/v1)t(z_{1},z_{2}):=\log(\|v_{2}\|/\|v_{1}\|) so that

p(z1,z2)=(x1,y1,(et(z1,z2)v1,et(z1,z2)Ψ1))=a~ρ,t(z1,z2)(z1).\displaystyle p(z_{1},z_{2})=\bigl{(}x_{1},y_{1},\bigl{(}e^{t(z_{1},z_{2})}v_{1},e^{-t(z_{1},z_{2})}\Psi_{1}\bigr{)}\bigr{)}=\tilde{a}_{\rho,t(z_{1},z_{2})}(z_{1}).

Observe that since 𝒦\mathcal{K} is compact, there exists constants c3>c2>0c_{3}>c_{2}>0 such that

dE((v1,Ψ1),(et(z1,z2)v1,et(z1,z2)Ψ1))c2t(z1,z2)\displaystyle d_{\mathrm{E}}\bigl{(}(v_{1},\Psi_{1}),\bigl{(}e^{t(z_{1},z_{2})}v_{1},e^{-t(z_{1},z_{2})}\Psi_{1}\bigr{)}\bigr{)}\leq c_{2}t(z_{1},z_{2})

and

dE((v2v1v1,v1v2Ψ1),(v2,Ψ2))\displaystyle d_{\mathrm{E}}\Bigl{(}\Bigl{(}\tfrac{\|v_{2}\|}{\|v_{1}\|}v_{1},\tfrac{\|v_{1}\|}{\|v_{2}\|}\Psi_{1}\Bigr{)},(v_{2},\Psi_{2})\Bigr{)} c2dΘ(x1,x2)+1v2dE(v1Ψ1,v2Ψ2)\displaystyle\leq c_{2}d_{\mathcal{F}_{\Theta}}(x_{1},x_{2})+\tfrac{1}{\|v_{2}\|}d_{\mathrm{E}}\bigr{(}\|v_{1}\|\Psi_{1},\|v_{2}\|\Psi_{2}\bigl{)}
c3(dΘ(x1,x2)+d𝗂Θ(y1,y2)),\displaystyle\leq c_{3}(d_{\mathcal{F}_{\Theta}}(x_{1},x_{2})+d_{\mathcal{F}_{\mathsf{i}\Theta}}(y_{1},y_{2})),

where the last inequality uses the observation that for i{1,2}i\in\{1,2\}, the linear form viΨiζρ(yi)\|v_{i}\|\Psi_{i}\in\zeta_{\rho}^{*}(y_{i}) which satisfies viΨi(vivi)=1\|v_{i}\|\Psi_{i}\bigl{(}\tfrac{v_{i}}{\|v_{i}\|}\bigr{)}=1 is smoothly determined by yiy_{i} and the unit vector vivi\tfrac{v_{i}}{\|v_{i}\|} (or equivalently, xix_{i}). The triangle inequality gives

dE((v1,Ψ1),(v2,Ψ2))c3((dΘ×d𝗂Θ)((x1,y1),(x2,y2))+|t(z1,z2)|).\displaystyle d_{\mathrm{E}}((v_{1},\Psi_{1}),(v_{2},\Psi_{2}))\leq c_{3}((d_{\mathcal{F}_{\Theta}}\times d_{\mathcal{F}_{\mathsf{i}\Theta}})((x_{1},y_{1}),(x_{2},y_{2}))+|t(z_{1},z_{2})|).

Hence, it suffices to show that there exists a constant c4>0c_{4}>0 such that for all z1,z2𝒦z_{1},z_{2}\in\mathcal{K}, we have

|t(z1,z2)|c4((dΘ×d𝗂Θ)((x1,y1),(x2,y2))+|logu^(z1)logu^(z2)|).\displaystyle|t(z_{1},z_{2})|\leq c_{4}((d_{\mathcal{F}_{\Theta}}\times d_{\mathcal{F}_{\mathsf{i}\Theta}})((x_{1},y_{1}),(x_{2},y_{2}))+|\log\hat{u}(z_{1})-\log\hat{u}(z_{2})|).

Since {ϕt}t\{\phi_{t}^{\mathcal{L}}\}_{t\in\mathbb{R}} is uniformly contracting with respect to \|\cdot\|^{\mathcal{L}}, there exists η>0\eta>0 such that

1u^(a~ρ,t(z))eηtu^(z) for all t>0 and z𝒢~ρ,\displaystyle\frac{1}{\hat{u}(\tilde{a}_{\rho,t}(z))}\leq\frac{e^{-\eta t}}{\hat{u}(z)}\qquad\text{ for all }t>0\text{ and }z\in\tilde{\mathcal{G}}_{\rho},

i.e., ηtlogu^(a~ρ,t(z1))logu^(z1)\eta t\leq\log\hat{u}(\tilde{a}_{\rho,t}(z_{1}))-\log\hat{u}(z_{1}). Then, for all z1,z2𝒦z_{1},z_{2}\in\mathcal{K}, we have

|logu^(z1)logu^(z2)|\displaystyle|\log\hat{u}(z_{1})-\log\hat{u}(z_{2})|
\displaystyle\geq |logu^(z1)logu^(a~ρ,t(z1,z2)(z1))||logu^(p1(z1,z2))logu^(z2)|\displaystyle|\log\hat{u}(z_{1})-\log\hat{u}(\tilde{a}_{\rho,t(z_{1},z_{2})}(z_{1}))|-|\log\hat{u}(p_{1}(z_{1},z_{2}))-\log\hat{u}(z_{2})|
\displaystyle\geq η|t(z1,z2)||logu^(p1(z1,z2))logu^(z2)|.\displaystyle\eta|t(z_{1},z_{2})|-|\log\hat{u}(p_{1}(z_{1},z_{2}))-\log\hat{u}(z_{2})|.

Lastly, since u^\hat{u} is locally Lipschitz with respect to any linear metric, there exists c5>0c_{5}>0 such that |logu^(p1(z1,z2))logu^(z2)|c5(dΘ×d𝗂Θ)((x1,y1),(x2,y2))|\log\hat{u}(p_{1}(z_{1},z_{2}))-\log\hat{u}(z_{2})|\leq c_{5}(d_{\mathcal{F}_{\Theta}}\times d_{\mathcal{F}_{\mathsf{i}\Theta}})((x_{1},y_{1}),(x_{2},y_{2})) for all z1,z2𝒦z_{1},z_{2}\in\mathcal{K} and this completes the proof. ∎

The following is an immediate consequence of Theorems 3.12 and 3.13.

Proposition 3.14.

The reparametrization {aκ(1(),t)}t\{a_{\kappa^{*}(\mathcal{H}^{-1}(\cdot),t)}\}_{t\in\mathbb{R}} of {at}t\{a_{t}\}_{t\in\mathbb{R}} is a metric Anosov flow with respect to ((Wρsu),(Wρss))\bigl{(}\mathcal{H}(W^{\mathrm{su}}_{\rho}),\mathcal{H}(W^{\mathrm{ss}}_{\rho})\bigr{)}.

Proof of Theorem 3.2.

In view of Proposition 3.14, by considering {at}t\{a_{t}\}_{t\in\mathbb{R}} as a reparametrization of {aκ(1(),t)}t\{a_{\kappa^{*}(\mathcal{H}^{-1}(\cdot),t)}\}_{t\in\mathbb{R}}, it suffices to check Properties (1) and (2) in Proposition 3.10.

Proof of Property (1). Fix a compact fundamental domain DΛΘ(2)×D\subset\Lambda_{\Theta}^{(2)}\times\mathbb{R} for the Γ\Gamma-action. Fix r>0r>0 sufficiently small so that the closed rr-balls in ΛΘ(2)×\Lambda_{\Theta}^{(2)}\times\mathbb{R} injectively project into 𝒳\mathcal{X} and that the projection DΛΘ(2)D^{\prime}\subset\Lambda_{\Theta}^{(2)} of the closed rr-neighborhood of DD is compact. Since DD is compact and d~\tilde{d} is locally bi-Lipschitz equivalent to the product metric, there exists c1>1c_{1}>1 such that for all zDz\in D and zΛΘ(2)×z^{\prime}\in\Lambda_{\Theta}^{(2)}\times\mathbb{R} with d~(z,z)<r\tilde{d}(z,z^{\prime})<r, we have

c11(dΘ×d𝗂Θ×d)(z,z)d(Γz,Γz)=d~(z,z)c1(dΘ×d𝗂Θ×d)(z,z).c_{1}^{-1}(d_{\mathcal{F}_{\Theta}}\times d_{\mathcal{F}_{\mathsf{i}\Theta}}\times d_{\mathbb{R}})(z,z^{\prime})\leq d(\Gamma z,\Gamma z^{\prime})=\tilde{d}(z,z^{\prime})\leq c_{1}(d_{\mathcal{F}_{\Theta}}\times d_{\mathcal{F}_{\mathsf{i}\Theta}}\times d_{\mathbb{R}})(z,z^{\prime}). (10)

Let ϵ(0,r)\epsilon\in(0,r) which will be specified later. Fix z=(x,y,t)Dz=(x,y,t)\in D, z1=(x,y,t1)W~ϵsu(z)z_{1}=(x^{\prime},y,t_{1})\in\tilde{W}_{\epsilon}^{\mathrm{su}}(z), and z2=(x,y,t2)ΛΘ(2)×z_{2}=(x^{\prime},y,t_{2})\in\Lambda_{\Theta}^{(2)}\times\mathbb{R} such that Γz2a(Γz1)(Wρsu(1(Γz))\Gamma z_{2}\in a_{\mathbb{R}}(\Gamma z_{1})\cap\mathcal{H}(W^{\mathrm{su}}_{\rho}(\mathcal{H}^{-1}(\Gamma z)). By compactness of DD, we can uniformly choose ϵ\epsilon sufficiently small so that d~(z,z2)<r\tilde{d}(z,z_{2})<r. We want to show that there exists a uniform constant c>1c>1 such that

c1d(Γz,Γz2)d(Γz,Γz1)cd(Γz,Γz2).c^{-1}d(\Gamma z,\Gamma z_{2})\leq d(\Gamma z,\Gamma z_{1})\leq cd(\Gamma z,\Gamma z_{2}). (11)

By (10), for each j{1,2}j\in\{1,2\}, we have

c11(dΘ(x,x)+|tti|)d~(z,zi)c1(dΘ(x,x)+|tti|).\displaystyle c_{1}^{-1}(d_{\mathcal{F}_{\Theta}}(x,x^{\prime})+|t-t_{i}|)\leq\tilde{d}(z,z_{i})\leq c_{1}(d_{\mathcal{F}_{\Theta}}(x,x^{\prime})+|t-t_{i}|).

By (9), we have |tt1|=|ψ([x,y]Θ[x,y]Θ)|c2dΘ(x,x)|t-t_{1}|=|\psi([x,y]_{\Theta}-[x^{\prime},y]_{\Theta})|\leq c_{2}d_{\mathcal{F}_{\Theta}}(x,x^{\prime}) for a uniform constant c2>0c_{2}>0 by smoothness of the Θ\Theta-Gromov product restricted to DD^{\prime}. Then

d(Γz,Γz1)c1(dΘ(x,x)+|tt1|)c1(1+c2)dΘ(x,x)c12(1+c2)d(Γz,Γz2).\displaystyle d(\Gamma z,\Gamma z_{1})\leq c_{1}(d_{\mathcal{F}_{\Theta}}(x,x^{\prime})+|t-t_{1}|)\leq c_{1}(1+c_{2})d_{\mathcal{F}_{\Theta}}(x,x^{\prime})\leq c_{1}^{2}(1+c_{2})d(\Gamma z,\Gamma z_{2}).

This establishes the right hand inequality of (11).

For the left hand inequality of (11), we observe that compactness of DD and the Lipschitz properties of the maps u^\hat{u} and \mathcal{H} in the proof of Theorem 3.13 implies that there exists a uniform constant c3>0c_{3}>0 such that

|tt2|=|logu^(1)(z)logu^(1)(z2)|c3d~(z,z2).\displaystyle|t-t_{2}|=|\log\hat{u}(\mathcal{H}^{-1})(z)-\log\hat{u}(\mathcal{H}^{-1})(z_{2})|\leq c_{3}\tilde{d}(z,z_{2}).

This establishes Property (1) for the strong unstable foliations. The argument for the strong stable foliations is similar.

Proof of Property (2). Property (2) is immediate from the fact that κ(,t)=0tf(aρ,s())𝑑s\kappa(\cdot,t)=\int_{0}^{t}f(a_{\rho,s}(\cdot))\,ds for all tt\in\mathbb{R} for some positive continuous function f:𝒢ρf:\mathcal{G}_{\rho}\to\mathbb{R} (see [18, Remark 4.11]). ∎

Remark 3.15.

Although we have only shown that the distinguished translation flow {at}t\{a_{t}\}_{t\in\mathbb{R}} on 𝒳\mathcal{X} is metric Anosov, it is not difficult to see that the above work can be adapted to show that any translation flow {at𝗏}t\{a^{\mathsf{v}}_{t}\}_{t\in\mathbb{R}} with 𝗏intΘ\mathsf{v}\in\operatorname{int}\mathcal{L}_{\Theta} on 𝒳\mathcal{X} is a Lipschitz reparametrization of {at}t\{a_{t}\}_{t\in\mathbb{R}} and also metric Anosov. Indeed, the proof of Theorems 3.13 and 3.14 is easily modified to show that for all 𝗏intΘ\mathsf{v}\in\operatorname{int}\mathcal{L}_{\Theta} there exists a bi-Lipschitz homeomorphism 𝗏0,𝗏:𝒳𝒳𝗏\mathcal{H}^{\mathsf{v}_{0},\mathsf{v}}:\mathcal{X}\to\mathcal{X}_{\mathsf{v}} conjugating {at𝗏}t\{a^{\mathsf{v}}_{t}\}_{t\in\mathbb{R}} to a Lipschitz reparametrization of {at}t\{a_{t}\}_{t\in\mathbb{R}} and that {at𝗏}t\{a^{\mathsf{v}}_{t}\}_{t\in\mathbb{R}} is metric Anosov. In fact, such a homeomorphism 𝗏0,𝗏\mathcal{H}^{\mathsf{v}_{0},\mathsf{v}} can be explicitly described as follows. Let πkerψ:𝔞Θkerψ\pi_{\ker\psi}:\mathfrak{a}_{\Theta}\to\ker\psi be the projection map determined by the decomposition 𝔞Θ=𝗏0kerψ\mathfrak{a}_{\Theta}=\mathbb{R}\mathsf{v}_{0}\oplus\ker\psi, 𝒰={γ𝒰j}γΓ, 1jj0\mathcal{U}=\{\gamma\mathcal{U}_{j}\}_{\gamma\in\Gamma,\,1\leq j\leq j_{0}} be a locally finite open cover of ΛΘ(2)×\Lambda_{\Theta}^{(2)}\times\mathbb{R} such that γ𝒰jγ𝒰j=\gamma\mathcal{U}_{j}\cap\gamma^{\prime}\mathcal{U}_{j}=\varnothing for all distinct γ,γΓ\gamma,\gamma^{\prime}\in\Gamma and 1jj01\leq j\leq j_{0} and {φi}iI\{\varphi_{i}\}_{i\in I} for some index set II be a partition of unity subordinate to 𝒰\mathcal{U} such that φi:ΛΘ(2)×[0,+)\varphi_{i}:\Lambda_{\Theta}^{(2)}\times\mathbb{R}\to[0,+\infty) is smooth along the ρ\rho-geodesic flow for all iIi\in I and {φi}iI\{\varphi_{i}\}_{i\in I} be a partition of unity subordinate to the cover {γUj}γΓ,1ij0\{\gamma U_{j}\}_{\gamma\in\Gamma,1\leq i\leq j_{0}}. Define the map ~𝗏0,𝗏:ΛΘ(2)×ΛΘ(2)×\tilde{\mathcal{H}}^{\mathsf{v}_{0},\mathsf{v}}:\Lambda_{\Theta}^{(2)}\times\mathbb{R}\to\Lambda_{\Theta}^{(2)}\times\mathbb{R} by ~𝗏0,𝗏(x,y,r)=(x,y,rψ𝗏(𝗏0)+logu^(x,y,r))\tilde{\mathcal{H}}^{\mathsf{v}_{0},\mathsf{v}}(x,y,r)=(x,y,r\psi_{\mathsf{v}}({\mathsf{v}}_{0})+\log\hat{u}(x,y,r)) where u^\hat{u} is given by

1u^(z)=0Tes/Texp(iI,suppφiγUj(ψ𝗏πkerψ)(βxΘ(e,γ))φi(x,y,r+t))𝑑s\displaystyle\frac{1}{\hat{u}(z)}=\int_{0}^{T}\frac{e^{s/T}}{\exp\left(\sum_{i\in I,\,\operatorname{supp}\varphi_{i}\subset\gamma U_{j}}(\psi_{\mathsf{v}}\circ\pi_{\ker\psi})(\beta^{\Theta}_{x}(e,\gamma))\cdot\varphi_{i}(x,y,r+t)\right)}\,ds

for all z=(x,y,r)ΛΘ(2)×z=(x,y,r)\in\Lambda_{\Theta}^{(2)}\times\mathbb{R}, for some T>0T>0. Then, it descends to the desired map 𝗏0,𝗏:𝒳𝒳𝗏\mathcal{H}^{\mathsf{v}_{0},\mathsf{v}}:\mathcal{X}\to\mathcal{X}_{\sf{v}}.

4. Transfer operators, Dolgopyat’s method, and the proof of Theorem 1.5 and applications

In this section, we cover the necessary background for transfer operators in order to state Theorem 4.7 which is the main technical theorem regarding spectral bounds. We also outline how to derive the main theorem stated in Theorem 1.5 from Theorem 4.7. A key point in this section is that all the theorems include the precise dependence on a parameter associated to 𝗏intΘ\mathsf{v}\in\operatorname{int}\mathcal{L}_{\Theta}.

4.1. Reduction of Theorem 1.5 by rescaling

Due to the homothety equivariance property for the family of translation flows (see Eq. 7), it is convenient to fix a scaling for each direction in intΘ\operatorname{int}\mathcal{L}_{\Theta} for the purpose of proving Theorem 1.5. It turns out that there is a particular scaling so that we have additional properties. Using the isomorphism 𝔞Θ𝔞Θ\mathfrak{a}_{\Theta}\cong\mathfrak{a}_{\Theta}^{*} induced by the inner product ,\langle\cdot,\cdot\rangle on 𝔞Θ\mathfrak{a}_{\Theta}, we abuse notation and identify the dual limit cone Θ𝔞Θ\mathcal{L}_{\Theta}^{*}\subset\mathfrak{a}_{\Theta}^{*} with a dual limit cone Θ𝔞Θ\mathcal{L}_{\Theta}^{*}\subset\mathfrak{a}_{\Theta} (see Eq. 1). Note that Θint𝔞Θ+{0}\mathcal{L}_{\Theta}\subset\operatorname{int}\mathfrak{a}_{\Theta}^{+}\cup\{0\} (see Theorem 2.11(2)) and Proposition B.1 implies that 𝔞Θ+intΘ{0}\partial\mathfrak{a}_{\Theta}^{+}\subset\operatorname{int}\mathcal{L}_{\Theta}^{*}\cup\{0\}. For all 𝗐intΘ\mathsf{w}\in\operatorname{int}\mathcal{L}_{\Theta}^{*} with 𝗐=1\|\mathsf{w}\|=1, we define 𝗏(𝗐)intΘ\mathsf{v}(\mathsf{w})\in\operatorname{int}\mathcal{L}_{\Theta} to be the unique vector such that ψ𝗏(𝗐)=𝗐\nabla\psi_{\mathsf{v}(\mathsf{w})}=\mathsf{w}. Then, Eq. 6 gives

ψΘ(𝗏(𝗐))ψΘ(𝗏(𝗐))=1.\displaystyle\frac{\|\nabla\psi_{\Theta}(\mathsf{v}(\mathsf{w}))\|}{\psi_{\Theta}(\mathsf{v}(\mathsf{w}))}=1. (12)

Therefore, Theorem 1.5 follows from the following exponential mixing theorem. Theorem 1.4 also follows either directly from the following or via Theorem 1.5. Note that one can repeat a standard convolution argument as in [34, Appendix] using Lipschitz continuous bump functions to handle general α\alpha-Hölder functions (see also [42, Corollary 5.2] and [62, Theorem 3.1.4]).

Theorem 4.1.

There exists a bounded elementary function EΘ:>0>0E_{\Theta}:\mathbb{R}_{>0}\to\mathbb{R}_{>0} with vanishing limit at ++\infty such that for all open neighborhoods 𝒩𝔞Θ+\mathcal{N}\supset\partial\mathfrak{a}_{\Theta}^{+} and 𝗏intΘ\mathsf{v}\in\operatorname{int}\mathcal{L}_{\Theta} with ψ𝗏=1\|\nabla\psi_{\mathsf{v}}\|=1 and ψ𝗏𝒩\nabla\psi_{\mathsf{v}}\notin\mathcal{N}, there exist C𝗏>0C_{\mathsf{v}}>0, k𝗏k_{\mathsf{v}}\in\mathbb{N}, a finite set {λk(𝗏)}k=1k𝗏(1,0)×i(1,1)\{\lambda_{k}(\mathsf{v})\}_{k=1}^{k_{\mathsf{v}}}\subset(-1,0)\times i(-1,1) which come in conjugate pairs, and a finite set of finite-rank positive bilinear forms {𝗏,k}k=1k𝗏\{\mathcal{B}_{\mathsf{v},k}\}_{k=1}^{k_{\mathsf{v}}}, such that for all ϕ1,ϕ2L(𝒳)\phi_{1},\phi_{2}\in L(\mathcal{X}) and t>0t>0, we have

|𝒳ϕ1(at𝗏x)ϕ2(x)𝑑m𝒳𝗏(x)(m𝒳𝗏(ϕ1)m𝒳𝗏(ϕ2)+k=1k𝗏eλk(𝗏)t𝗏,k(ϕ1,ϕ2))|CeEΘ(ψΘ(𝗏))tϕ1Lipϕ2Lip.\left|\int_{\mathcal{X}}\phi_{1}(a^{\mathsf{v}}_{t}x)\phi_{2}(x)\,dm_{\mathcal{X}}^{\mathsf{v}}(x)-\left(m_{\mathcal{X}}^{\mathsf{v}}(\phi_{1})m_{\mathcal{X}}^{\mathsf{v}}(\phi_{2})+\sum_{k=1}^{k_{\mathsf{v}}}e^{\lambda_{k}(\mathsf{v})t}\mathcal{B}_{\mathsf{v},k}(\phi_{1},\phi_{2})\right)\right|\\ \leq Ce^{-E_{\Theta}(\|\nabla\psi_{\Theta}(\mathsf{v})\|)t}\|\phi_{1}\|_{\operatorname{Lip}}\|\phi_{2}\|_{\operatorname{Lip}}.

The rest of the paper is devoted to the proof of the above theorem. To this end, we fix an open neighborhood 𝒩𝔞Θ+\mathcal{N}\supset\partial\mathfrak{a}_{\Theta}^{+} henceforth and define the subset of unit vectors

𝓌={𝗐intΘ:𝗐=1,𝗐𝒩}.\displaystyle\mathscr{w}=\{\mathsf{w}\in\operatorname{int}\mathcal{L}_{\Theta}^{*}:\|\mathsf{w}\|=1,\mathsf{w}\notin\mathcal{N}\}.

Note that we have a corresponding compact subset of unit vectors

𝓌¯={𝗐Θ:𝗐=1,𝗐𝒩}.\displaystyle\overline{\mathscr{w}}=\{\mathsf{w}\in\mathcal{L}_{\Theta}:\|\mathsf{w}\|=1,\mathsf{w}\notin\mathcal{N}\}.

4.2. Compatible family of Markov sections

Corollary 3.3 gives a Markov section for any fixed translation flow. Recalling that we have a family of translation flows {{at𝗏(𝗐)}t}𝗐𝓌\bigl{\{}\bigl{\{}a^{\mathsf{v}(\mathsf{w})}_{t}\bigr{\}}_{t\in\mathbb{R}}\bigr{\}}_{\mathsf{w}\in\mathscr{w}}, we fix a corresponding family of Markov sections {𝗐}𝗐𝓌\{\mathcal{R}^{\mathsf{w}}\}_{\mathsf{w}\in\mathscr{w}} on 𝒳\mathcal{X}. They have several other corresponding objects (see Subsection 3.1) all of which we denote using a superscript ‘𝗐\mathsf{w}’ on the same notations. By the same observation as in [18, §4.2] regarding Markov sections of reparametrized flows, we deduce the remarkable property that the family of Markov sections can be chosen such that for different unit vectors 𝗐,𝗐𝓌\mathsf{w},\mathsf{w}^{\prime}\in\mathscr{w}, the Markov sections 𝗐\mathcal{R}^{\mathsf{w}} and 𝗐\mathcal{R}^{\mathsf{w}^{\prime}} consist of corresponding rectangles which project to each other along the translation flow foliation while fixing their centers. Consequently, for some fixed NN\in\mathbb{N}, we denote

𝗐={R1𝗐,R2𝗐,,RN𝗐}={[U1𝗐,S1𝗐],[U2𝗐,S2𝗐],,[UN𝗐,SN𝗐]}for all 𝗐𝓌\displaystyle\mathcal{R}^{\mathsf{w}}=\{R^{\mathsf{w}}_{1},R^{\mathsf{w}}_{2},\dotsc,R^{\mathsf{w}}_{N}\}=\{[U^{\mathsf{w}}_{1},S^{\mathsf{w}}_{1}],[U^{\mathsf{w}}_{2},S^{\mathsf{w}}_{2}],\dotsc,[U^{\mathsf{w}}_{N},S^{\mathsf{w}}_{N}]\}\qquad\text{for all $\mathsf{w}\in\mathscr{w}$}

and the family {𝗐}𝗐𝓌\{\mathcal{R}^{\mathsf{w}}\}_{\mathsf{w}\in\mathscr{w}} comes equipped with a family

{Ψ𝗐,𝗐:R𝗐R𝗐}𝗐,𝗐𝓌\displaystyle\{\Psi^{\mathsf{w},\mathsf{w}^{\prime}}:R^{\mathsf{w}}\to R^{\mathsf{w}^{\prime}}\}_{\mathsf{w},\mathsf{w}^{\prime}\in\mathscr{w}}

of compatibility maps which are Lipschitz homeomorphisms and vary smoothly in 𝗐,𝗐𝓌\mathsf{w},\mathsf{w}^{\prime}\in\mathscr{w}. More precisely, there exists a family {t𝗐,𝗐:R𝗐}𝗐,𝗐𝓌\{t^{\mathsf{w},\mathsf{w}^{\prime}}:R^{\mathsf{w}}\to\mathbb{R}\}_{\mathsf{w},\mathsf{w}^{\prime}\in\mathscr{w}} of Lipschitz continuous functions which vary smoothly in 𝗐,𝗐𝓌\mathsf{w},\mathsf{w}^{\prime}\in\mathscr{w} such that t𝗐,𝗐(wj)=0t^{\mathsf{w},\mathsf{w}^{\prime}}(w_{j})=0 at the centers wjRj𝗐w_{j}\in R^{\mathsf{w}}_{j} for all 1jN1\leq j\leq N and

Φ𝗐,𝗐(u)=at𝗐,𝗐(u)𝗏(𝗐)uR𝗐for all uR𝗐 and 𝗐,𝗐𝓌.\displaystyle\Phi^{\mathsf{w},\mathsf{w}^{\prime}}(u)=a^{\mathsf{v}(\mathsf{w})}_{t^{\mathsf{w},\mathsf{w}^{\prime}}(u)}u\in R^{\mathsf{w}^{\prime}}\qquad\text{for all $u\in R^{\mathsf{w}}$ and $\mathsf{w},\mathsf{w}^{\prime}\in\mathscr{w}$}.

All the properties are clear save injectivity which we now justify. For all 𝗐,𝗐𝓌\mathsf{w},\mathsf{w}^{\prime}\in\mathscr{w}, the map Ψ𝗐,𝗐\Psi^{\mathsf{w},\mathsf{w}^{\prime}} moves the points in R𝗐R^{\mathsf{w}} along the translation flow foliation to which the strong stable and unstable foliations corresponding to both 𝗐,𝗐𝓌\mathsf{w},\mathsf{w}^{\prime}\in\mathscr{w} are transverse and hence preserves the time ordering along the translation flow foliation. Now, if Ψ𝗐,𝗐(u1)=Ψ𝗐,𝗐(u2)\Psi^{\mathsf{w},\mathsf{w}^{\prime}}(u_{1})=\Psi^{\mathsf{w},\mathsf{w}^{\prime}}(u_{2}) for some distinct u1,u2R𝗐u_{1},u_{2}\in R^{\mathsf{w}}, then the time ordering would be violated since u1u_{1} and u2u_{2} must lie on the same translation flow orbit; implying injectivity.

We now use the compatibility maps Ψ𝗐0,𝗐\Psi^{\mathsf{w}_{0},\mathsf{w}} to identify the Markov sections 𝗐\mathcal{R}^{\mathsf{w}} for all 𝗐𝓌\mathsf{w}\in\mathscr{w} with a fixed Markov section :=𝗐0\mathcal{R}:=\mathcal{R}^{\mathsf{w}_{0}} of some size δ^(0,ϵ0)\hat{\delta}\in(0,\epsilon_{0}), where 𝗐0\mathsf{w}_{0} corresponds to 𝗏0\mathsf{v}_{0}, i.e., 𝗏(𝗐0)>0𝗏0\mathsf{v}(\mathsf{w}_{0})\in\mathbb{R}_{>0}\mathsf{v}_{0}. We need to address a subtle point. For any 𝗐𝓌\mathsf{w}\in\mathscr{w} outside of some open neighborhood of 𝗐0𝓌\mathsf{w}_{0}\in\mathscr{w}, the corresponding Markov sections need not be of size δ^\hat{\delta}. However, apart from that, they are legitimate Markov sections and come from the local product structures on the image Ψ𝗐0,𝗐(j𝒜[Wϵ0su(wj),Wϵ0ss(wj)])\Psi^{\mathsf{w}_{0},\mathsf{w}}\bigl{(}\bigsqcup_{j\in\mathcal{A}}[W^{\mathrm{su}}_{\epsilon_{0}}(w_{j}),W^{\mathrm{ss}}_{\epsilon_{0}}(w_{j})]\bigr{)}, where wjRj𝗐0w_{j}\in R_{j}^{\mathsf{w}_{0}} are the centers for all j𝒜j\in\mathcal{A}. Thus, we can drop all superscripts henceforth, except for {τ𝗐}𝗐𝓌\{\tau^{\mathsf{w}}\}_{\mathsf{w}\in\mathscr{w}} which we view as a family of first return time maps on the same set UU; since they still vary in 𝗐\mathsf{w}, they need to be distinguished. A useful property is that they remain positive, again due to preservation of time ordering. We record this and other useful properties in the following lemma (recall Eq. 8).

Lemma 4.2.

The family {τ𝗐}𝗐𝓌\{\tau^{\mathsf{w}}\}_{\mathsf{w}\in\mathscr{w}} is smooth in 𝗐𝓌\mathsf{w}\in\mathscr{w} and there exists Cτ>0C_{\tau}>0 such that

  1. ((1))

    0<τ𝗐(u)Cτ2τ𝗐0¯τ𝗐0(u)Cτ20<\tau^{\mathsf{w}}(u)\leq\frac{C_{\tau}}{2\overline{\tau^{\mathsf{w}_{0}}}}\tau^{\mathsf{w}_{0}}(u)\leq\frac{C_{\tau}}{2} for all uUu\in U and 𝗐𝓌\mathsf{w}\in\mathscr{w},

  2. ((2))

    max(j,k)Lip(τ𝗐|𝙲[j,k])Cτ2\max_{(j,k)}\operatorname{Lip}\bigl{(}\tau^{\mathsf{w}}|_{\mathtt{C}[j,k]}\bigr{)}\leq\frac{C_{\tau}}{2} for all 𝗐𝓌\mathsf{w}\in\mathscr{w}.

Consequently, {max(j,k)τ𝗐|𝙲[j,k]Lip}𝗐𝓌\bigl{\{}\max_{(j,k)}\|\tau^{\mathsf{w}}|_{\mathtt{C}[j,k]}\|_{\operatorname{Lip}}\bigr{\}}_{\mathsf{w}\in\mathscr{w}} is bounded above by CτC_{\tau}.

Proof.

Positivity of τ𝗐\tau^{\mathsf{w}} for all 𝗐𝓌\mathsf{w}\in\mathscr{w}, as mentioned above, is clear and so we prove the other properties. We use notations from Subsection 5.1 and terminology from Subsection 4.3 to shorten the proof. Namely, we use the first return vector map 𝖪:U𝔞Θ\mathsf{K}:U\to\mathfrak{a}_{\Theta} descended from the one in Subsection 5.1 which is Lipschitz continuous on cylinders of length 11. Then, we have the identity τ𝗐=ψ𝗏(𝗐)𝖪\tau^{\mathsf{w}}=\psi_{\mathsf{v}(\mathsf{w})}\circ\mathsf{K} for all 𝗐𝓌\mathsf{w}\in\mathscr{w}. Thus, the lemma follows from this and the property that ψ𝗏(𝗐)=1\|\nabla\psi_{\mathsf{v}(\mathsf{w})}\|=1. ∎

Due to the above lemma, we conclude that the smooth family {τ𝗐}𝗐𝓌\{\tau^{\mathsf{w}}\}_{\mathsf{w}\in\mathscr{w}} can be extended to a smooth family of uniformly essentially Lipschitz (i.e., with a uniform Lipschitz constant on cylinders of length 11) nonnegative first return time maps {τ𝗐}𝗐𝓌¯\{\tau^{\mathsf{w}}\}_{\mathsf{w}\in\overline{\mathscr{w}}}.

Using a compatible family of Markov sections as constructed in this subsection is necessary to execute Dolgopyat’s method in a uniform fashion for the family of translation flows. Although we cannot make sense of a translation flow on 𝒳\mathcal{X} or even transfer operators associated to any 𝗏Θ\mathsf{v}\in\partial\mathcal{L}_{\Theta}, the above uniform bounds and the extended family of first return time maps is used to obtain other Lipschitz bounds and a uniform version of LNIC to derive the precise form of the exponential decay rate in Theorem 4.7 for the family of transfer operators in 𝗐𝓌\mathsf{w}\in\mathscr{w}.

4.3. Symbolic dynamics

Let 𝒜={1,2,,N}\mathcal{A}=\{1,2,\dotsc,N\} be the alphabet of \mathcal{R}. Define the N×NN\times N transition matrix TT by

Tj,k={1,int(Rj)𝒫1(int(Rk))0,otherwisefor all 1j,kN.\displaystyle T_{j,k}=\begin{cases}1,&\operatorname{int}(R_{j})\cap\mathcal{P}^{-1}(\operatorname{int}(R_{k}))\neq\varnothing\\ 0,&\text{otherwise}\end{cases}\qquad\text{for all $1\leq j,k\leq N$}.

The transition matrix TT is topologically mixing as a consequence of [18, Theorem 8.1], i.e., there exists NTN_{T}\in\mathbb{N} such that TNTT^{N_{T}} consists only of positive entries. This was assumed in [18, §4.3] as well which is a minor inaccuracy; see Appendix C for the correction. Define the spaces of bi-infinite and infinite admissible sequences

Σ\displaystyle\Sigma ={(,x1,x0,x1,)𝒜:Txj,xj+1=1 for all j};\displaystyle=\{(\dotsc,x_{-1},x_{0},x_{1},\dotsc)\in\mathcal{A}^{\mathbb{Z}}:T_{x_{j},x_{j+1}}=1\text{ for all }j\in\mathbb{Z}\};
Σ+\displaystyle\Sigma^{+} ={(x0,x1,)𝒜0:Txj,xj+1=1 for all j0}\displaystyle=\{(x_{0},x_{1},\dotsc)\in\mathcal{A}^{\mathbb{Z}_{\geq 0}}:T_{x_{j},x_{j+1}}=1\text{ for all }j\in\mathbb{Z}_{\geq 0}\}

respectively. We will use the term admissible sequences for finite sequences as well in the natural way. For any fixed β0(0,1)\beta_{0}\in(0,1), we endow Σ\Sigma with the metric dd defined by d(x,y)=β0inf{|j|0:xjyj}d(x,y)=\beta_{0}^{\inf\{|j|\in\mathbb{Z}_{\geq 0}:x_{j}\neq y_{j}\}} for all x,yΣx,y\in\Sigma. We similarly endow Σ+\Sigma^{+} with a metric which we also denote by dd.

Definition 4.3 (Cylinder).

For all admissible sequences x=(x0,x1,,xk)x=(x_{0},x_{1},\dotsc,x_{k}) of length len(x)=k0\operatorname{len}(x)=k\in\mathbb{Z}_{\geq 0}, we define the cylinder of length len(𝙲[x])=k\operatorname{len}(\mathtt{C}[x])=k to be

𝙲[x]\displaystyle\mathtt{C}[x] ={uU:σj(u)int(Uxj) for all 0jk}.\displaystyle=\{u\in U:\sigma^{j}(u)\in\operatorname{int}(U_{x_{j}})\text{ for all }0\leq j\leq k\}.

We denote cylinders simply by 𝙲\mathtt{C} (or other typewriter style letters) when we do not need to specify the admissible sequence.

By a slight abuse of notation, let σ\sigma also denote the shift map on Σ\Sigma or Σ+\Sigma^{+}. There exist natural continuous surjections η:ΣR\eta:\Sigma\to R and η+:Σ+U\eta^{+}:\Sigma^{+}\to U defined by η(x)=j=𝒫j(int(Rxj))¯\eta(x)=\bigcap_{j=-\infty}^{\infty}\overline{\mathcal{P}^{-j}(\operatorname{int}(R_{x_{j}}))} for all xΣx\in\Sigma and η+(x)=j=0σj(int(Uxj))¯\eta^{+}(x)=\bigcap_{j=0}^{\infty}\overline{\sigma^{-j}(\operatorname{int}(U_{x_{j}}))} for all xΣ+x\in\Sigma^{+}. Define Σ^=η1(R^)\hat{\Sigma}=\eta^{-1}(\hat{R}) and Σ^+=(η+)1(U^)\hat{\Sigma}^{+}=(\eta^{+})^{-1}(\hat{U}). Then the restrictions η|Σ^:Σ^R^\eta|_{\hat{\Sigma}}:\hat{\Sigma}\to\hat{R} and η+|Σ^+:Σ^+U^\eta^{+}|_{\hat{\Sigma}^{+}}:\hat{\Sigma}^{+}\to\hat{U} are bijective and satisfy η|Σ^σ|Σ^=𝒫|R^η|Σ^\eta|_{\hat{\Sigma}}\circ\sigma|_{\hat{\Sigma}}=\mathcal{P}|_{\hat{R}}\circ\eta|_{\hat{\Sigma}} and η+|Σ^+σ|Σ^+=σ|U^η+|Σ^+\eta^{+}|_{\hat{\Sigma}^{+}}\circ\sigma|_{\hat{\Sigma}^{+}}=\sigma|_{\hat{U}}\circ\eta^{+}|_{\hat{\Sigma}^{+}}.

We can take β0(0,1)\beta_{0}\in(0,1) sufficiently close to 11 so that η\eta and η+\eta^{+} are Lipschitz continuous [6, Lemma 2.2]. Let L(Σ,)L(\Sigma,\mathbb{R}) denote the space of Lipschitz continuous functions f:Σf:\Sigma\to\mathbb{R}. We use similar notations for Lipschitz function spaces with domain Σ+\Sigma^{+} and other codomains. For all function spaces, we suppress the codomain when it is \mathbb{R}.

Let 𝗐𝓌¯\mathsf{w}\in\overline{\mathscr{w}}. Since the horospherical foliations on GG are smooth, we conclude that τ𝗐\tau^{\mathsf{w}} is Lipschitz continuous on cylinders of length 11. Then the maps (τ𝗐η)|Σ^(\tau^{\mathsf{w}}\circ\eta)|_{\hat{\Sigma}} and (τ𝗐η+)|Σ^+(\tau^{\mathsf{w}}\circ\eta^{+})|_{\hat{\Sigma}^{+}} are Lipschitz continuous and hence there exist unique Lipschitz extensions which, by abuse of notation, we also denote by τ𝗐:Σ\tau^{\mathsf{w}}:\Sigma\to\mathbb{R} and τ𝗐:Σ+\tau^{\mathsf{w}}:\Sigma^{+}\to\mathbb{R}, respectively.

4.4. Thermodynamics

Definition 4.4 (Pressure).

For all fL(Σ)f\in L(\Sigma), called the potential, the pressure is defined by

Prσ(f)=supνσ1(Σ){Σf𝑑ν+hν(σ)}\displaystyle\operatorname{Pr}_{\sigma}(f)=\sup_{\nu\in\mathcal{M}^{1}_{\sigma}(\Sigma)}\left\{\int_{\Sigma}f\,d\nu+h_{\nu}(\sigma)\right\}

where σ1(Σ)\mathcal{M}^{1}_{\sigma}(\Sigma) is the set of σ\sigma-invariant Borel probability measures on Σ\Sigma and hν(σ)h_{\nu}(\sigma) is the measure theoretic entropy of σ\sigma with respect to ν\nu.

For all fL(Σ)f\in L(\Sigma), there exists a unique σ\sigma-invariant Borel probability measure νf\nu_{f} on Σ\Sigma which attains the supremum in Definition 4.4, called the ff-equilibrium state [7, Theorems 2.17 and 2.20]. It satisfies νf(Σ^)=1\nu_{f}(\hat{\Sigma})=1 [15, Corollary 3.2].

Let 𝗐𝓌\mathsf{w}\in\mathscr{w}. Associated to ψ𝗏(𝗐)\psi_{\mathsf{v}(\mathsf{w})}, we relabel δ𝗐:=δ𝗏(𝗐)=δψ𝗏(𝗐)\delta_{\mathsf{w}}:=\delta_{\mathsf{v}(\mathsf{w})}=\delta_{\psi_{\mathsf{v}(\mathsf{w})}} (see Eq. 2) and recall from Subsection 2.4 that δ𝗐=ψΘ(𝗏(𝗐))=ψΘ(𝗏(𝗐))\delta_{\mathsf{w}}=\psi_{\Theta}(\mathsf{v}(\mathsf{w}))=\|\nabla\psi_{\Theta}(\mathsf{v}(\mathsf{w}))\| where we have used Eq. 12. Thanks to Corollary 3.3, we can use [13, Theorem A.2] which states that m𝒳𝗏(𝗐)m_{\mathcal{X}}^{\mathsf{v}(\mathsf{w})} is a measure of maximal entropy for the translation flow which attains the maximal entropy of δ𝗐\delta_{\mathsf{w}}. We will consider in particular the probability measure νδ𝗐τ𝗐\nu_{-\delta_{\mathsf{w}}\tau^{\mathsf{w}}} on Σ\Sigma with corresponding pressure Prσ(δ𝗐τ𝗐)=0\operatorname{Pr}_{\sigma}(-\delta_{\mathsf{w}}\tau^{\mathsf{w}})=0 [8, Proposition 3.1] (cf. [15, Theorem 4.4]), which we will denote simply by νΣ𝗐\nu_{\Sigma}^{\mathsf{w}}. Define νR𝗐=η(νΣ𝗐)\nu_{R}^{\mathsf{w}}=\eta_{*}(\nu_{\Sigma}^{\mathsf{w}}) and νU𝗐=(projU)(νR𝗐)\nu_{U}^{\mathsf{w}}=(\operatorname{proj}_{U})_{*}(\nu_{R}^{\mathsf{w}}). Note that νU𝗐(τ𝗐)=νR𝗐(τ𝗐)\nu_{U}^{\mathsf{w}}(\tau^{\mathsf{w}})=\nu_{R}^{\mathsf{w}}(\tau^{\mathsf{w}}). We refer the reader to [18, §4.4] for the relation between m𝒳𝗏(𝗐)m_{\mathcal{X}}^{\mathsf{v}(\mathsf{w})} and νR𝗐\nu_{R}^{\mathsf{w}}, which we do not require directly in this paper.

4.5. Transfer operators

Throughout the paper, we will use the notation ξ=a+ib\xi=a+ib\in\mathbb{C} for the complex parameter for the transfer operators.

Definition 4.5 (Transfer operators).

For all ξ\xi\in\mathbb{C} and 𝗐𝓌¯\mathsf{w}\in\overline{\mathscr{w}}, the transfer operator ξτ𝗐:C(U,)C(U,)\mathcal{L}_{\xi\tau^{\mathsf{w}}}:C\bigl{(}U,\mathbb{C}\bigr{)}\to C\bigl{(}U,\mathbb{C}\bigr{)} is defined by

ξτ𝗐(H)(u)=uσ1(u)eξτ𝗐(u)H(u)\displaystyle\mathcal{L}_{\xi\tau^{\mathsf{w}}}(H)(u)=\sum_{u^{\prime}\in\sigma^{-1}(u)}e^{\xi\tau^{\mathsf{w}}(u^{\prime})}H(u^{\prime})

for all uUu\in U and HC(U,)H\in C\bigl{(}U,\mathbb{C}\bigr{)}.

We recall the Ruelle–Perron–Frobenius (RPF) theorem along with the theory of Gibbs measures in this setting [7, 44]. For all aa\in\mathbb{R} and 𝗐𝓌\mathsf{w}\in\mathscr{w}, there exist a unique positive function ha,𝗐L(U)h_{a,\mathsf{w}}\in L(U) and a unique Borel probability measure νa,𝗐\nu_{a,\mathsf{w}} on UU such that νa,𝗐(ha,𝗐)=1\nu_{a,\mathsf{w}}(h_{a,\mathsf{w}})=1 and

(δ𝗐+a)τ𝗐(ha,𝗐)\displaystyle\mathcal{L}_{-(\delta_{\mathsf{w}}+a)\tau^{\mathsf{w}}}(h_{a,\mathsf{w}}) =λa,𝗐ha,𝗐,\displaystyle=\lambda_{a,\mathsf{w}}h_{a,\mathsf{w}}, (δ𝗐+a)τ𝗐(νa,𝗐)\displaystyle\mathcal{L}_{-(\delta_{\mathsf{w}}+a)\tau^{\mathsf{w}}}^{*}(\nu_{a,\mathsf{w}}) =λa,𝗐νa,𝗐\displaystyle=\lambda_{a,\mathsf{w}}\nu_{a,\mathsf{w}}

where λa,𝗐:=ePrσ((δ𝗐+a)τ𝗐)\lambda_{a,\mathsf{w}}:=e^{\operatorname{Pr}_{\sigma}(-(\delta_{\mathsf{w}}+a)\tau^{\mathsf{w}})} is the maximal simple eigenvalue of (δ𝗐+a)τ𝗐\mathcal{L}_{-(\delta_{\mathsf{w}}+a)\tau^{\mathsf{w}}} and the rest of the spectrum of (δ𝗐+a)τ𝗐|L(U,)\mathcal{L}_{-(\delta_{\mathsf{w}}+a)\tau^{\mathsf{w}}}|_{L(U,\mathbb{C})} is contained in a disk of radius strictly less than λa,𝗐\lambda_{a,\mathsf{w}}. Moreover, dνU𝗐=h0,𝗐dν0,𝗐d\nu_{U}^{\mathsf{w}}=h_{0,\mathsf{w}}\,d\nu_{0,\mathsf{w}} and λ0,𝗐=1\lambda_{0,\mathsf{w}}=1 (see Subsection 4.4).

Let 𝗐𝓌\mathsf{w}\in\mathscr{w}. As usual, it is convenient to normalize the transfer operators. For all aa\in\mathbb{R} define the map

τ𝗐,(a)=(δ𝗐+a)τ𝗐+logh0,𝗐logh0,𝗐σL(U).\displaystyle\tau^{\mathsf{w},(a)}=-(\delta_{\mathsf{w}}+a)\tau^{\mathsf{w}}+\log\circ h_{0,\mathsf{w}}-\log\circ h_{0,\mathsf{w}}\circ\sigma\in L(U).

For all kk\in\mathbb{N}, we define the Lipschitz continuous maps τk𝗐:U\tau_{k}^{\mathsf{w}}:U\to\mathbb{R} and τk𝗐,(a):U\tau_{k}^{\mathsf{w},(a)}:U\to\mathbb{R} by

τk𝗐(u)\displaystyle\tau_{k}^{\mathsf{w}}(u) =j=0k1τ𝗐(σj(u)),\displaystyle=\sum_{j=0}^{k-1}\tau^{\mathsf{w}}(\sigma^{j}(u)), τk𝗐,(a)(u)\displaystyle\tau_{k}^{\mathsf{w},(a)}(u) =j=0k1τ𝗐,(a)(σj(u))\displaystyle=\sum_{j=0}^{k-1}\tau^{\mathsf{w},(a)}(\sigma^{j}(u))

and τ0𝗐(u)=τ0(a),𝗐(u)=0\tau_{0}^{\mathsf{w}}(u)=\tau_{0}^{(a),\mathsf{w}}(u)=0 for all uUu\in U. For all ξ\xi\in\mathbb{C}, we define ξ,𝗐:C(U,)C(U,)\mathcal{L}_{\xi,\mathsf{w}}:C(U,\mathbb{C})\to C(U,\mathbb{C}) and its kkth iterates for all k0k\in\mathbb{Z}_{\geq 0} by

ξ,𝗐k(H)(u)=uσk(u)e(τk𝗐,(a)+ibτk𝗐)(u)H(u)\displaystyle\mathcal{L}_{\xi,\mathsf{w}}^{k}(H)(u)=\sum_{u^{\prime}\in\sigma^{-k}(u)}e^{(\tau_{k}^{\mathsf{w},(a)}+ib\tau_{k}^{\mathsf{w}})(u^{\prime})}H(u^{\prime})

for all uUu\in U and HC(U,)H\in C(U,\mathbb{C}). With this normalization, for all aa\in\mathbb{R}, the maximal simple eigenvalue of a,𝗐\mathcal{L}_{a,\mathsf{w}} is λa,𝗐\lambda_{a,\mathsf{w}} with eigenvector ha,𝗐h0,𝗐\frac{h_{a,\mathsf{w}}}{h_{0,\mathsf{w}}}. Moreover, we have 0,𝗐(νU𝗐)=νU𝗐\mathcal{L}_{0,\mathsf{w}}^{*}(\nu_{U}^{\mathsf{w}})=\nu_{U}^{\mathsf{w}}.

We fix some related constants. Let 𝗐𝓌\mathsf{w}\in\mathscr{w}. By perturbation theory of operators as in [28, Chapter 7] and [44, Proposition 4.6], we can fix a0>0a_{0}^{\prime}>0 such that the map [a0,a0][-a_{0}^{\prime},a_{0}^{\prime}]\to\mathbb{R} defined by aλa,𝗐a\mapsto\lambda_{a,\mathsf{w}} and the map [a0,a0]C(U,)[-a_{0}^{\prime},a_{0}^{\prime}]\to C(U,\mathbb{R}) defined by aha,𝗐a\mapsto h_{a,\mathsf{w}} are Lipschitz uniformly in 𝗐𝓌\mathsf{w}\in\mathscr{w}. To see uniformity in 𝗐𝓌\mathsf{w}\in\mathscr{w}, a standard calculation using the eigenvalue equation and Lemma 4.2 gives |dda|a=aλa,𝗐|=|νa,𝗐(τ𝗐ha,𝗐)|Cτ\bigl{|}\left.\frac{d}{da^{\prime}}\right|_{a^{\prime}=a}\lambda_{a^{\prime},\mathsf{w}}\bigr{|}=|\nu_{a,\mathsf{w}}(\tau^{\mathsf{w}}h_{a,\mathsf{w}})|\leq C_{\tau}. Similar estimates in the construction of the family of eigenvectors {ha,𝗐}a[a0,a0]\{h_{a,\mathsf{w}}\}_{a\in[-a_{0}^{\prime},a_{0}^{\prime}]} (see [44, Theorem 2.2]) gives the latter. Fix Aτ>0A_{\tau}>0 such that it is greater that the aforementioned Lipschitz constants and so that |τ𝗐,(a)(u)τ𝗐,(0)(u)|Aτ|a||\tau^{\mathsf{w},(a)}(u)-\tau^{\mathsf{w},(0)}(u)|\leq A_{\tau}|a| for all uUu\in U and |a|a0|a|\leq a_{0}^{\prime}. Again by similar estimates, we also have ha,𝗐Lipδ𝗐=ψΘ(𝗏(𝗐))\|h_{a,\mathsf{w}}\|_{\operatorname{Lip}}\ll\delta_{\mathsf{w}}=\|\nabla\psi_{\Theta}(\mathsf{v}(\mathsf{w}))\| which tends to ++\infty as 𝗐\mathsf{w} tends to the boundary of intΘ\operatorname{int}\mathcal{L}_{\Theta}^{*}. Fix

T𝗐>\displaystyle T_{\mathsf{w}}>{} max(max𝗐𝓌¯max(j,k)τ𝗐|𝙲[j,k]Lip,sup|a|a0max(j,k)τ𝗐,(a)|𝙲[j,k]Lip)\displaystyle\max\bigg{(}\max_{\mathsf{w}\in\overline{\mathscr{w}}}\max_{(j,k)}\|\tau^{\mathsf{w}}|_{\mathtt{C}[j,k]}\|_{\operatorname{Lip}},\sup_{|a|\leq a_{0}^{\prime}}\max_{(j,k)}\bigl{\|}\tau^{\mathsf{w},(a)}|_{\mathtt{C}[j,k]}\bigr{\|}_{\operatorname{Lip}}\bigg{)}

which can be chosen as some elementary function evaluated at δ𝗐=ψΘ(𝗏(𝗐))\delta_{\mathsf{w}}=\|\nabla\psi_{\Theta}(\mathsf{v}(\mathsf{w}))\|, for all 𝗐𝓌\mathsf{w}\in\mathscr{w} (see also [46, Lemma 4.1] for taking the supremum over [a0,a0][-a_{0}^{\prime},a_{0}^{\prime}]).

4.6. Dolgopyat’s Method and the proof of Theorem 4.1 and applications

Recall that L(U,)L(U,\mathbb{C}) denotes the space of complex Lipschitz continuous functions on UU. It is a Banach space with the Lipschitz seminorm and norm

Lip(H)\displaystyle\operatorname{Lip}(H) =supu,uU,uu|H(u)H(u)|d(u,u),\displaystyle=\sup_{u,u^{\prime}\in U,u\neq u^{\prime}}\frac{|H(u)-H(u^{\prime})|}{d(u,u^{\prime})}, HLip\displaystyle\|H\|_{\operatorname{Lip}} =H+Lip(H)\displaystyle=\|H\|_{\infty}+\operatorname{Lip}(H)

for all HL(U,)H\in L(U,\mathbb{C}), where \|\cdot\|_{\infty} denotes the usual LL^{\infty}-norm. For all bb\in\mathbb{R}, we also generalize the above norm to

H1,b=H+1max(1,|b|)Lip(H)\displaystyle\|H\|_{1,b}=\|H\|_{\infty}+\frac{1}{\max(1,|b|)}\operatorname{Lip}(H)

for all HL(U,)H\in L(U,\mathbb{C}), which will be required later.

Following Stoyanov [67, §5], we define the convenient new metric DD on UU by

D(u,u)={0,u=uminu,u𝙲cylinder 𝙲diam(𝙲¯),uu and u,uUj for some j𝒜1,otherwise.\displaystyle D(u,u^{\prime})=\begin{cases}0,&u=u^{\prime}\\ \displaystyle\min_{\begin{subarray}{c}u,u^{\prime}\in\mathtt{C}\\ \text{cylinder }\mathtt{C}\end{subarray}}\operatorname{diam}(\overline{\mathtt{C}}),&u\neq u^{\prime}\text{ and }u,u^{\prime}\in U_{j}\text{ for some }j\in\mathcal{A}\\ 1,&\text{otherwise}.\end{cases}

We denote LD(U,)L_{D}(U,\mathbb{R}) to be the space of DD-Lipschitz continuous functions and LipD(h)\operatorname{Lip}_{D}(h) to be the DD-Lipschitz constant for all hLD(U,)h\in L_{D}(U,\mathbb{R}). We define the cones

𝒞B(U)\displaystyle\mathcal{C}_{B}(U) ={hLD(U,):h>0,|h(u)h(u)|Bh(u)D(u,u) for all u,uU}\displaystyle=\{h\in L_{D}(U,\mathbb{R}):h>0,|h(u)-h(u^{\prime})|\leq Bh(u)D(u,u^{\prime})\text{ for all }u,u^{\prime}\in U\}
𝒞~B(U)\displaystyle\tilde{\mathcal{C}}_{B}(U) ={hLD(U,):h>0,eBD(u,u)h(u)h(u)eBD(u,u) for all u,uU}.\displaystyle=\Big{\{}h\in L_{D}(U,\mathbb{R}):h>0,e^{-BD(u,u^{\prime})}\leq\frac{h(u)}{h(u^{\prime})}\leq e^{BD(u,u^{\prime})}\text{ for all }u,u^{\prime}\in U\Big{\}}.
Remark 4.6.

If h𝒞B(U)h\in\mathcal{C}_{B}(U), then using the convexity of log-\log, we can derive that hh is log\log-DD-Lipschitz, i.e., |(logh)(u)(logh)(u)|BD(u,u)|(\log\circ h)(u)-(\log\circ h)(u^{\prime})|\leq BD(u,u^{\prime}) for all u,uUu,u^{\prime}\in U. It follows that 𝒞B(U)𝒞~B(U)\mathcal{C}_{B}(U)\subset\tilde{\mathcal{C}}_{B}(U), however the reverse containment is not true.

We now state Theorem 4.7 regarding spectral bounds of transfer operators. Such bounds are the key for obtaining exponential mixing results.

Theorem 4.7.

There exist an elementary function η:>0>0\eta:\mathbb{R}_{>0}\to\mathbb{R}_{>0} with vanishing limit at ++\infty, a0>0a_{0}>0, and b0>0b_{0}>0 such that for all 𝗐𝓌\mathsf{w}\in\mathscr{w}, there exists C𝗐>0C_{\mathsf{w}}>0 such that for all ξ\xi\in\mathbb{C} with |a|<a0|a|<a_{0} and |b|>b0|b|>b_{0}, kk\in\mathbb{N}, and HL(U,)H\in L(U,\mathbb{C}), we have

ξ,𝗐k(H)2C𝗐eη(δ𝗐)kH1,b.\displaystyle\big{\|}\mathcal{L}_{\xi,\mathsf{w}}^{k}(H)\big{\|}_{2}\leq C_{\mathsf{w}}e^{-\eta(\delta_{\mathsf{w}})k}\|H\|_{1,b}.

By a standard inductive argument (see the proof after [64, Theorem 5.4]) which goes back to Dolgopyat [22], Theorem 4.7 is a consequence of the following theorem which records the mechanism of Dolgopyat’s method.

Theorem 4.8.

There exist mm\in\mathbb{N}, an elementary function η:>0(0,1)\eta:\mathbb{R}_{>0}\to(0,1) which tends to 11 at ++\infty, E>max(1,1b0)E>\max\left(1,\frac{1}{b_{0}}\right), a0>0a_{0}>0, b0>0b_{0}>0, and a set of operators {𝒩a,J𝗐:LD(U,)LD(U,):𝗐𝓌,|a|<a0,J𝒥(b) for some |b|>b0}\{\mathcal{N}_{a,J}^{\mathsf{w}}:L_{D}(U,\mathbb{R})\to L_{D}(U,\mathbb{R}):\mathsf{w}\in\mathscr{w},|a|<a_{0},J\in\mathcal{J}(b)\text{ for some }|b|>b_{0}\}, where 𝒥(b)\mathcal{J}(b) is some finite set for all |b|>b0|b|>b_{0}, such that

  1. ((1))

    𝒩a,J𝗐(𝒞E|b|(U))𝒞E|b|(U)\mathcal{N}_{a,J}^{\mathsf{w}}(\mathcal{C}_{E|b|}(U))\subset\mathcal{C}_{E|b|}(U) for all 𝗐𝓌\mathsf{w}\in\mathscr{w}, |a|<a0|a|<a_{0}, J𝒥(b)J\in\mathcal{J}(b), and |b|>b0|b|>b_{0};

  2. ((2))

    𝒩a,J𝗐(h)2η(δ𝗐)h2\left\|\mathcal{N}_{a,J}^{\mathsf{w}}(h)\right\|_{2}\leq\eta(\delta_{\mathsf{w}})\|h\|_{2} for all h𝒞E|b|(U)h\in\mathcal{C}_{E|b|}(U), 𝗐𝓌\mathsf{w}\in\mathscr{w}, |a|<a0|a|<a_{0}, J𝒥(b)J\in\mathcal{J}(b), and |b|>b0|b|>b_{0};

  3. ((3))

    for all |a|<a0|a|<a_{0}, |b|>b0|b|>b_{0}, and 𝗐𝓌\mathsf{w}\in\mathscr{w}, if HL(U,)H\in L(U,\mathbb{C}) and h𝒞E|b|(U)h\in\mathcal{C}_{E|b|}(U) satisfy

    1. (1a)

      |H(u)|h(u)|H(u)|\leq h(u) for all uUu\in U;

    2. (1b)

      |H(u)H(u)|E|b|h(u)D(u,u)|H(u)-H(u^{\prime})|\leq E|b|h(u)D(u,u^{\prime}) for all u,uUu,u^{\prime}\in U;

    then there exists J𝒥(b)J\in\mathcal{J}(b) such that

    1. (2a)

      |ξ,𝗐m(H)(u)|𝒩a,J𝗐(h)(u)\bigl{|}\mathcal{L}_{\xi,\mathsf{w}}^{m}(H)(u)\bigr{|}\leq\mathcal{N}_{a,J}^{\mathsf{w}}(h)(u) for all uUu\in U;

    2. (2b)

      |ξ,𝗐m(H)(u)ξ,𝗐m(H)(u)|E|b|𝒩a,J𝗐(h)(u)D(u,u)\bigl{|}\mathcal{L}_{\xi,\mathsf{w}}^{m}(H)(u)-\mathcal{L}_{\xi,\mathsf{w}}^{m}(H)(u^{\prime})\bigr{|}\leq E|b|\mathcal{N}_{a,J}^{\mathsf{w}}(h)(u)D(u,u^{\prime}) for all u,uUu,u^{\prime}\in U.

Theorem 4.1 for Lipschitz continuous functions is derived from the spectral bounds of transfer operators in Theorem 4.7 using arguments of Pollicott and Paley–Wiener. The derivation is similar to that of [64, §10] and so we omit it and describe the differences. One minor difference is that in our case the derivation would be a Lipschitz version of loc. cit. (cf. [41, §9]). Another difference is that we include the Pollicott–Ruelle resonances since we cannot control simple eigenvalue of the transfer operators ξ,𝗐\mathcal{L}_{\xi,\mathsf{w}} for the complex parameter ξ\xi in a fixed neighborhood of 00\in\mathbb{C} uniformly in 𝗐𝓌\mathsf{w}\in\mathscr{w}. For this, we also refer the reader to [47]; more specifically, [47, Proposition 4], the proof of [47, Theorem 1], and the remark after [47, Corollary 1].

Let us give some more details for integrating out the stable direction in the derivation mentioned above. The constants η>0\eta>0 and C>0C>0 which come from the metric Anosov property vary smoothly in 𝗐\mathsf{w}. However, the decay rate η\eta could, a priori, get worse as 𝗐\mathsf{w} tends to Θ\partial\mathcal{L}_{\Theta}^{*}. The exact behavior comes from the lower bound for the derivative of the reparametrization κ(x,t)\kappa(x,t) in tt. But this comes exactly from the upper bound for supvΘ,v=1ψ𝗏(𝗐)\sup_{v\in\mathcal{L}_{\Theta},\|v\|=1}\psi_{\mathsf{v}(\mathsf{w})} which, by construction, is bounded above by 11 uniformly in 𝗐𝓌\mathsf{w}\in\mathscr{w}. Thus, one can choose a decay rate η\eta uniformly in 𝗐𝓌\mathsf{w}\in\mathscr{w}.

Remark 4.9.

We warn the reader of a subtle point. The derivation in [64, §10] crucially uses the Anosov property of the frame flow; the spectral bounds of transfer operators on their own are not enough. Analogously, although the proof of Theorem 4.7 does not require the metric Anosov property, the proof of Theorem 4.1 would be incomplete without it. Though the metric Anosov property of the translation flow is claimed in the literature on Anosov subgroups, it is far from a triviality. In our paper, the metric Anosov property was stated in Theorem 3.2 and a complete proof was provided in Subsection 3.4, thus, justifying the described derivation above of Theorem 4.1 for Lipschitz continuous functions.

Let us now describe the derivations of Theorems 1.6 and 1.8. The former is proved using the relationship between the zeros of the Selberg zeta function Z𝗏(𝗐)Z_{\mathsf{v}(\mathsf{w})} and the transfer operators ξ,𝗐\mathcal{L}_{\xi,\mathsf{w}} stated in [47, Proposition 4], and of course the spectral bounds in Theorem 4.7 (cf. proof of [47, Theorem 1]). The latter is proved using the number theoretic techniques in [50] where the weaker spectral bounds in [50, Eq. (2.1)] (cf. [50, Proposition 2]) is replaced again by the stronger spectral bounds in Theorem 4.7.

As a result of our discussion, it suffices to focus only on the proof of Theorem 4.8 for the rest of the paper.

5. Local non-integrability condition

This section is devoted to a crucial ingredient for Dolgopyat’s method stated in Proposition 5.11 called the local non-integrability condition (LNIC) as in [67], which is then upgraded to Proposition 5.12. Although Proposition 5.12 is stated in reverse Lipschitz form on int(U)\operatorname{int}(U), we obtain it via Proposition 5.11 which requires Lie theory and hence the smooth structure on GG. Thus, we first resolve technical issues related to the incompatibility of ΛΘ\Lambda_{\Theta} with the smooth structure on GG.

5.1. Smooth extensions of coding maps

Fix 𝗐𝓌\mathsf{w}\in\mathscr{w} throughout this subsection. As alluded to above, due to the fractal nature of ΛΘ\Lambda_{\Theta}, the Markov section does not inherit the smooth structure on GG. In order to circumvent this issue, we need to enlarge UU to an open set and smoothly extend the maps associated to the coding, τ𝗐\tau^{\mathsf{w}}, 𝒫\mathcal{P}, and σ\sigma. This type of construction was done in [64, §5.1] based on [57, Lemma 1.2]. However, even this procedure has many more technical issues in our setting. One problem is that 𝒳=Γ\(ΛΘ(2)×)\mathcal{X}=\Gamma\backslash\bigl{(}\Lambda_{\Theta}^{(2)}\times\mathbb{R}\bigr{)} is not naturally situated in a larger smooth manifold with a natural extension of the translation flow. Note that in general the action ΓΘ(2)×\Gamma\curvearrowright\mathcal{F}_{\Theta}^{(2)}\times\mathbb{R} is not properly discontinuous (see Theorem A.2). To avoid this problem, we work on the cover ΛΘ(2)×\Lambda_{\Theta}^{(2)}\times\mathbb{R} instead and use the inclusion

ΛΘ(2)×Θ(2)×\displaystyle\Lambda_{\Theta}^{(2)}\times\mathbb{R}\hookrightarrow\mathcal{F}_{\Theta}^{(2)}\times\mathbb{R}

which is a Lipschitz embedding into a smooth manifold where the translation flow is still defined. Another problem is that it is difficult to gain control on the Γ\Gamma-orbit of the open neighborhood of U=j𝒜UjU=\bigsqcup_{j\in\mathcal{A}}U_{j} in Θ(2)×\mathcal{F}_{\Theta}^{(2)}\times\mathbb{R} (e.g., diameter estimates, mutual disjointness, seperation distance) since we do not have a Γ\Gamma-invariant metric on Θ(2)×\mathcal{F}_{\Theta}^{(2)}\times\mathbb{R} and ΓΘ(2)×\Gamma\curvearrowright\mathcal{F}_{\Theta}^{(2)}\times\mathbb{R} may not be properly discontinuous. To avoid these topological obstructions, we allow dependence of the open neighborhoods on the arbitrarily large length mm\in\mathbb{N} of sections of σm\sigma^{m} which will appear in Proposition 5.11. We do not follow [57, Lemma 1.2] because, although we believe that it is true, it would take significantly more work to properly formulate and establish the eventually contracting property of the smooth extension of σ1\sigma^{-1} on some fixed open neighborhood of UU. See the recent related work of Delarue–Montclair–Sanders [20] for the case that Γ\Gamma is a projective Anosov subgroup of PSLn()\operatorname{PSL}_{n}(\mathbb{R}) which gives a construction of a flow-invariant smooth manifold 𝒳\mathcal{M}\supset\mathcal{X} on which the translation flow is a contact Axiom A flow.

In light of the above discussion, let us first fix a connected fundamental domain 𝖣ΛΘ(2)×\mathsf{D}\subset\Lambda_{\Theta}^{(2)}\times\mathbb{R} for 𝒳\mathcal{X}, unique isometric lifts 𝖱j=[𝖴j,𝖲j]𝖣\mathsf{R}_{j}=[\mathsf{U}_{j},\mathsf{S}_{j}]\subset\mathsf{D} of Rj=[Uj,Sj]R_{j}=[U_{j},S_{j}] for all j𝒜j\in\mathcal{A}, and write 𝖱:=j𝒜𝖱j\mathsf{R}:=\bigsqcup_{j\in\mathcal{A}}\mathsf{R}_{j} and 𝖴:=j𝒜𝖴j\mathsf{U}:=\bigsqcup_{j\in\mathcal{A}}\mathsf{U}_{j}. Denote by u~𝖱\tilde{u}\in\mathsf{R} the unique lift of uRu\in R. Abusing notation, we denote the lift of cylinders by the same symbol. Since the translation flow is defined on ΛΘ(2)×\Lambda_{\Theta}^{(2)}\times\mathbb{R}, the maps τ𝗐\tau^{\mathsf{w}}, 𝒫\mathcal{P}, and σ\sigma have natural lifts τ𝗐:Γ𝖱\tau^{\mathsf{w}}:\Gamma\mathsf{R}\to\mathbb{R}, 𝒫:Γ𝖱Γ𝖱\mathcal{P}:\Gamma\mathsf{R}\to\Gamma\mathsf{R}, and σ:Γ𝖴Γ𝖴\sigma:\Gamma\mathsf{U}\to\Gamma\mathsf{U}, abusing notation. For all u𝖱u\in\mathsf{R} we define 𝗀(u)Γ\mathsf{g}(u)\in\Gamma to be the unique element such that 𝒫(u)𝗀(u)𝖱\mathcal{P}(u)\in\mathsf{g}(u)\mathsf{R}. Similarly, for all admissible pairs (j,k)(j,k), we write 𝗀(j,k)Γ\mathsf{g}^{(j,k)}\in\Gamma for the unique element such that

σ(int(𝖴j))𝗀(j,k)int(𝖴k).\displaystyle\sigma(\operatorname{int}(\mathsf{U}_{j}))\cap\mathsf{g}^{(j,k)}\operatorname{int}(\mathsf{U}_{k})\neq\varnothing.

Then, 𝗀^={𝗀(j,k):(j,k) is an admissible pair}Γ\hat{\mathsf{g}}=\{\mathsf{g}^{(j,k)}:(j,k)\text{ is an admissible pair}\}\subset\Gamma is a generating subset. For all admissible sequences α=(α0,α1,,αk)\alpha=(\alpha_{0},\alpha_{1},\dotsc,\alpha_{k}) for some k0k\in\mathbb{Z}_{\geq 0}, we extend the notation and write

𝗀α:=j=0k1𝗀(αj,αj+1)\displaystyle\mathsf{g}^{\alpha}:=\prod_{j=0}^{k-1}\mathsf{g}^{(\alpha_{j},\alpha_{j+1})} (13)

in ascending order if k>0k>0, and 𝗀α:=e\mathsf{g}^{\alpha}:=e if k=0k=0, and 𝗀α:=(𝗀α)1\mathsf{g}^{-\alpha}:=(\mathsf{g}^{\alpha})^{-1}.

Let j𝒜j\in\mathcal{A}, wj𝖴jw_{j}\in\mathsf{U}_{j} be the center, and γΓ\gamma\in\Gamma. We have a metric on WΘ,ss(γwj)W^{\mathcal{F}_{\Theta},\mathrm{ss}}(\gamma w_{j}) (resp. WΘ,wu(γwj)W^{\mathcal{F}_{\Theta},\mathrm{wu}}(\gamma w_{j})) which is induced by d𝗂Θd_{\mathcal{F}_{\mathsf{i}\Theta}} (resp. dΘ×dd_{\mathcal{F}_{\Theta}}\times d_{\mathbb{R}}) using coordinate pullbacks and denoted by the same notation. Then, (WϵγΘ,wu(γwj),WϵγΘ,ss(γwj))\bigl{(}W_{\epsilon_{\gamma}}^{\mathcal{F}_{\Theta},\mathrm{wu}}(\gamma w_{j}),W_{\epsilon_{\gamma}}^{\mathcal{F}_{\Theta},\mathrm{ss}}(\gamma w_{j})\bigr{)} has a local product structure such that [WϵγΘ,su(γwj),WϵγΘ,ss(γwj)]\bigl{[}W_{\epsilon_{\gamma}}^{\mathcal{F}_{\Theta},\mathrm{su}}(\gamma w_{j}),W_{\epsilon_{\gamma}}^{\mathcal{F}_{\Theta},\mathrm{ss}}(\gamma w_{j})\bigr{]} contains γ𝖱j\gamma\mathsf{R}_{j}, for some ϵγ>0\epsilon_{\gamma}>0. Now, fix open neighborhoods 𝖴~jγWϵγΘ,su(γwj){}^{\gamma}\tilde{\mathsf{U}}_{j}\subset W_{\epsilon_{\gamma}}^{\mathcal{F}_{\Theta},\mathrm{su}}(\gamma w_{j}) of γ𝖴j\gamma\mathsf{U}_{j} and define 𝖱~jγ=[𝖴~jγ,γ𝖲j]{}^{\gamma}\tilde{\mathsf{R}}_{j}=[{}^{\gamma}\tilde{\mathsf{U}}_{j},\gamma\mathsf{S}_{j}] such that {𝖱~jγ}j𝒜,γΓ\{{}^{\gamma}\tilde{\mathsf{R}}_{j}\}_{j\in\mathcal{A},\gamma\in\Gamma} consists of mutually disjoint rectangles. Define 𝖱~j:=𝖱~je\tilde{\mathsf{R}}_{j}:={}^{e}\tilde{\mathsf{R}}_{j}, 𝖱~:=j𝒜𝖱~j\tilde{\mathsf{R}}:=\bigsqcup_{j\in\mathcal{A}}\tilde{\mathsf{R}}_{j}, 𝖱~jΓ:=γΓ𝖱~jγ{}^{\Gamma}\tilde{\mathsf{R}}_{j}:=\bigsqcup_{\gamma\in\Gamma}{}^{\gamma}\tilde{\mathsf{R}}_{j}, and 𝖱~Γ:=j𝒜,γΓ𝖱~jγ{}^{\Gamma}\tilde{\mathsf{R}}:=\bigsqcup_{j\in\mathcal{A},\gamma\in\Gamma}{}^{\gamma}\tilde{\mathsf{R}}_{j}, and similarly define 𝖴~j\tilde{\mathsf{U}}_{j}, 𝖴~\tilde{\mathsf{U}}, 𝖴~jΓ{}^{\Gamma}\tilde{\mathsf{U}}_{j}, and 𝖴~Γ{}^{\Gamma}\tilde{\mathsf{U}}. We similarly omit the superscript ‘ee’ and use the superscript ‘Γ\Gamma’ for other sets.

Let m0m\in\mathbb{Z}_{\geq 0}. Recall that the translation flow is defined on Θ(2)×\mathcal{F}_{\Theta}^{(2)}\times\mathbb{R} and smooth. Thus, for all j𝒜j\in\mathcal{A} and γΓ\gamma\in\Gamma, we obtain compactly contained open neighborhoods 𝖴~j(m)γ𝖴~jγ{}^{\gamma}\tilde{\mathsf{U}}_{j}^{(m)}\subset{}^{\gamma}\tilde{\mathsf{U}}_{j} of γ𝖴j\gamma\mathsf{U}_{j} which are decreasing in mm and the natural smooth injective extensions σα:𝖴~αm(m)Γ𝗀α𝖴~α0Γ\sigma^{-\alpha}:{}^{\Gamma\mathsf{g}^{\alpha}}\tilde{\mathsf{U}}_{\alpha_{m}}^{(m)}\to{}^{\Gamma}\tilde{\mathsf{U}}_{\alpha_{0}} of (σ|𝙲[α])1\bigl{(}\sigma|_{\mathtt{C}[\alpha]}\bigr{)}^{-1} with cylinders 𝙲~γ[α](m):=σα(𝖴~αm(m)γ𝗀α)𝖴~α0γ{}^{\gamma}\tilde{\mathtt{C}}[\alpha]^{(m)}:=\sigma^{-\alpha}\bigl{(}{}^{\gamma\mathsf{g}^{\alpha}}\tilde{\mathsf{U}}_{\alpha_{m}}^{(m)}\bigr{)}\subset{}^{\gamma}\tilde{\mathsf{U}}_{\alpha_{0}}; and we define 𝖱~j(m)γ:=[𝖴~j(m)γ,γ𝖲j]{}^{\gamma}\tilde{\mathsf{R}}_{j}^{(m)}:=[{}^{\gamma}\tilde{\mathsf{U}}_{j}^{(m)},\gamma\mathsf{S}_{j}] and σα=(σα)1:𝙲~Γ[α](m)𝖴~αm(m)Γ𝗀α\sigma^{\alpha}=(\sigma^{-\alpha})^{-1}:{}^{\Gamma}\tilde{\mathtt{C}}[\alpha]^{(m)}\to{}^{\Gamma\mathsf{g}^{\alpha}}\tilde{\mathsf{U}}_{\alpha_{m}}^{(m)}, for all admissible sequences α=(α0,α1,,αm)\alpha=(\alpha_{0},\alpha_{1},\dotsc,\alpha_{m}).

There are also natural smooth extensions τ(j,k)𝗐:𝙲~Γ[j,k](1)\tau^{\mathsf{w}}_{(j,k)}:{}^{\Gamma}\tilde{\mathtt{C}}[j,k]^{(1)}\to\mathbb{R} for all admissible pairs (j,k)(j,k). We define the smooth maps τα𝗐:𝙲~Γ[α](m)\tau^{\mathsf{w}}_{\alpha}:{}^{\Gamma}\tilde{\mathtt{C}}[\alpha]^{(m)}\to\mathbb{R} by

τα𝗐(u)\displaystyle\tau^{\mathsf{w}}_{\alpha}(u) =j=0m1τ(αj,αj+1)𝗐(σ(α0,α1,,αj)(u))\displaystyle=\sum_{j=0}^{m-1}\tau^{\mathsf{w}}_{(\alpha_{j},\alpha_{j+1})}(\sigma^{(\alpha_{0},\alpha_{1},\dotsc,\alpha_{j})}(u)) (14)

for all admissible sequences α=(α0,α1,,αm)\alpha=(\alpha_{0},\alpha_{1},\dotsc,\alpha_{m}).

Following [18, §5] we can construct a smooth section

𝖥:𝖱~ΓG\displaystyle\mathsf{F}:{}^{\Gamma}\tilde{\mathsf{R}}\to G

in a similar fashion such that:

  1. ((1))

    for all j𝒜j\in\mathcal{A} and γΓ\gamma\in\Gamma, and u,u𝖴~jγu,u^{\prime}\in{}^{\gamma}\tilde{\mathsf{U}}_{j}, there exists unique n+NΘ+n^{+}\in N^{+}_{\Theta} such that 𝖥(u)=𝖥(u)n+\mathsf{F}(u^{\prime})=\mathsf{F}(u)n^{+};

  2. ((2))

    for all j𝒜j\in\mathcal{A} and γΓ\gamma\in\Gamma, u𝖴~jγu\in{}^{\gamma}\tilde{\mathsf{U}}_{j}, and s,sγ𝖲js,s^{\prime}\in\gamma\mathsf{S}_{j}, there exists unique nNΘn^{-}\in N^{-}_{\Theta} such that 𝖥([u,s])=𝖥([u,s])n\mathsf{F}([u,s^{\prime}])=\mathsf{F}([u,s])n^{-}.

Observe that 𝖥\mathsf{F} can indeed be constructed such that it does not depend on 𝗐\mathsf{w} due to the properties of the compatible Markov sections.

We introduce the first return vector map and holonomy. Although we only need the first return time map, since it does not require excess work, we provide the general definitions and the subsequent fundamental lemmas in anticipation that it will turn out useful elsewhere.

Definition 5.1 (First return vector map, Holonomy).

The first return Θ\Theta-vector map and Θ\Theta-holonomy are Γ\Gamma-invariant maps 𝖪:Γ𝖱𝔞Θ\mathsf{K}:\Gamma\mathsf{R}\to\mathfrak{a}_{\Theta} and ϑ:Γ𝖱SΘ\vartheta:\Gamma\mathsf{R}\to S_{\Theta} respectively that associate to each uγ𝖱u\in\gamma\mathsf{R} for some γΓ\gamma\in\Gamma the unique elements 𝖪(u)𝔞Θ\mathsf{K}(u)\in\mathfrak{a}_{\Theta} and ϑ(u)SΘ\vartheta(u)\in S_{\Theta} which satisfy

𝖥(𝒫(u))=𝗀(u)𝖥(u)a𝖪(u)ϑ(u).\displaystyle\mathsf{F}(\mathcal{P}(u))=\mathsf{g}(u)\mathsf{F}(u)a_{\mathsf{K}(u)}\vartheta(u).

We often drop the prefix “Θ\Theta-”.

Again, 𝖪\mathsf{K} and ϑ\vartheta are independent of 𝗐\mathsf{w}. Similar to previous constructions, we have smooth extensions 𝖪(j,k):𝙲~Γ[j,k](1)𝔞Θ\mathsf{K}_{(j,k)}:{}^{\Gamma}\tilde{\mathtt{C}}[j,k]^{(1)}\to\mathfrak{a}_{\Theta} and ϑ(j,k):𝙲~Γ[j,k](1)SΘ\vartheta^{(j,k)}:{}^{\Gamma}\tilde{\mathtt{C}}[j,k]^{(1)}\to S_{\Theta} for all admissible pairs (j,k)(j,k). Again, for all m0m\in\mathbb{Z}_{\geq 0} and admissible sequences α=(α0,α1,,αm)\alpha=(\alpha_{0},\alpha_{1},\dotsc,\alpha_{m}), we define the smooth maps 𝖪α:𝙲~Γ[α](m)𝔞Θ\mathsf{K}_{\alpha}:{}^{\Gamma}\tilde{\mathtt{C}}[\alpha]^{(m)}\to\mathfrak{a}_{\Theta} and ϑα:𝙲~Γ[α](m)SΘ\vartheta^{\alpha}:{}^{\Gamma}\tilde{\mathtt{C}}[\alpha]^{(m)}\to S_{\Theta} as in Eq. 14 and in the style of Eq. 13 for holonomy.

5.2. Local non-integrability condition

We are now ready to prove Propositions 5.11 and 5.12. The proof follows the techniques developed in [64, 17]. The version here is actually more simplified in some parts and we obtain a stronger LNIC because we are dealing with the “geodesic flow” (see Remark 5.10).

We start with a slight generalization of [64, Definition 6.1]. Recall the identifications from Subsection 4.2.

Definition 5.2 (Associated sequence in GG).

Let 𝗐intΘ\mathsf{w}\in\operatorname{int}\mathcal{L}_{\Theta}^{*} with 𝗐=1\|\mathsf{w}\|=1. Let z1𝖱~1z_{1}\in\tilde{\mathsf{R}}_{1} be the center. Consider a sequence (z1,z2,z3,z4,z1)(𝖱~1)5(z_{1},z_{2},z_{3},z_{4},z_{1})\in(\tilde{\mathsf{R}}_{1})^{5} such that z2𝖲1z_{2}\in\mathsf{S}_{1}, z4𝖴~1z_{4}\in\tilde{\mathsf{U}}_{1}, and z3=[z4,z2]z_{3}=[z_{4},z_{2}]. We define an associated sequence in GG to be the unique sequence (g1,g2,,g5)G5(g_{1},g_{2},\dotsc,g_{5})\in G^{5} where

g1\displaystyle g_{1} =𝖥(z1),\displaystyle=\mathsf{F}(z_{1}),
g2\displaystyle g_{2} =𝖥(z2)g1NΘ such that πψ𝗏(𝗐)(g2SΘ)=z2,\displaystyle=\mathsf{F}(z_{2})\in g_{1}N^{-}_{\Theta}\text{ such that }\pi_{\psi_{\mathsf{v}(\mathsf{w})}}(g_{2}S_{\Theta})=z_{2},
g3\displaystyle g_{3} g2NΘ+ such that at𝗏(𝗐)(πψ𝗏(𝗐)(g3SΘ))=z3 for some t(τ¯,τ¯),\displaystyle\in g_{2}N^{+}_{\Theta}\text{ such that }a^{\mathsf{v}(\mathsf{w})}_{t}\bigl{(}\pi_{\psi_{\mathsf{v}(\mathsf{w})}}(g_{3}S_{\Theta})\bigr{)}=z_{3}\text{ for some }t\in(-\underline{\tau},\underline{\tau}),
g4\displaystyle g_{4} g3NΘ such that at𝗏(𝗐)(πψ𝗏(𝗐)(g4SΘ))=z4 for some t(τ¯,τ¯),\displaystyle\in g_{3}N^{-}_{\Theta}\text{ such that }a^{\mathsf{v}(\mathsf{w})}_{t}\bigl{(}\pi_{\psi_{\mathsf{v}(\mathsf{w})}}(g_{4}S_{\Theta})\bigr{)}=z_{4}\ \text{ for some }t\in(-\underline{\tau},\underline{\tau}),
g5\displaystyle g_{5} g4NΘ+ such that at𝗏(𝗐)(πψ𝗏(𝗐)(g5SΘ))=z1 for some t(τ¯,τ¯).\displaystyle\in g_{4}N^{+}_{\Theta}\text{ such that }a^{\mathsf{v}(\mathsf{w})}_{t}\bigl{(}\pi_{\psi_{\mathsf{v}(\mathsf{w})}}(g_{5}S_{\Theta})\bigr{)}=z_{1}\text{ for some }t\in(-\underline{\tau},\underline{\tau}).
Remark 5.3.

In the above definition, the associated sequence (g1,g2,g3,g4,g5)(g_{1},g_{2},g_{3},g_{4},g_{5}) corresponding to each (z1,z2,z3,z4,z1)(z_{1},z_{2},z_{3},z_{4},z_{1}) is independent of 𝗐\mathsf{w}.

We continue using the notation in the above definition. Define the subsets

(NΘ+)1\displaystyle(N^{+}_{\Theta})_{1} ={n+NΘ+:𝖥(z1)n+𝖥(𝖴~1)}NΘ+,\displaystyle=\{n^{+}\in N^{+}_{\Theta}:\mathsf{F}(z_{1})n^{+}\in\mathsf{F}(\tilde{\mathsf{U}}_{1})\}\subset N^{+}_{\Theta},
(NΘ)1\displaystyle(N^{-}_{\Theta})_{1} ={nNΘ:𝖥(z1)n𝖥(𝖲1)}NΘ,\displaystyle=\{n^{-}\in N^{-}_{\Theta}:\mathsf{F}(z_{1})n^{-}\in\mathsf{F}(\mathsf{S}_{1})\}\subset N^{-}_{\Theta},

where the first is open and the second is compact. Now, if the above sequence (z1,z2,z3,z4,z1)(z_{1},z_{2},z_{3},z_{4},z_{1}) corresponds to some n+(NΘ+)1n^{+}\in(N^{+}_{\Theta})_{1} and some n(NΘ)1n^{-}\in(N^{-}_{\Theta})_{1} such that 𝖥(z4)=𝖥(z1)n+\mathsf{F}(z_{4})=\mathsf{F}(z_{1})n^{+} and 𝖥(z2)=𝖥(z1)n\mathsf{F}(z_{2})=\mathsf{F}(z_{1})n^{-} respectively, then we can define the map Ξ:(NΘ+)1×(NΘ)1AΘSΘ\Xi:(N^{+}_{\Theta})_{1}\times(N^{-}_{\Theta})_{1}\to A_{\Theta}S_{\Theta} by

Ξ(n+,n)=g51g1AΘSΘ.\displaystyle\Xi(n^{+},n^{-})=g_{5}^{-1}g_{1}\in A_{\Theta}S_{\Theta}.

To view it as a function of the first coordinate for a fixed n(NΘ)1n^{-}\in(N^{-}_{\Theta})_{1}, we write Ξn:(NΘ+)1AΘSΘ\Xi_{n^{-}}:(N^{+}_{\Theta})_{1}\to A_{\Theta}S_{\Theta}.

Let z1𝖱~1z_{1}\in\tilde{\mathsf{R}}_{1} be the center. Let jj\in\mathbb{N} and α=(α0,α2,,αj1,1)\alpha=(\alpha_{0},\alpha_{2},\dotsc,\alpha_{j-1},1) be an admissible sequence. Then, there exists an element which we denote by nα(NΘ)1n_{\alpha}\in(N^{-}_{\Theta})_{1} such that

𝖥(𝒫j(σα(z1)))=𝖥(z1)nα.\displaystyle\mathsf{F}(\mathcal{P}^{j}(\sigma^{-\alpha}(z_{1})))=\mathsf{F}(z_{1})n_{\alpha}.

This is well-defined because σα(z1)𝗀α𝙲[α]𝗀α𝖴\sigma^{-\alpha}(z_{1})\in\mathsf{g}^{-\alpha}\mathtt{C}[\alpha]\subset\mathsf{g}^{-\alpha}\mathsf{U}.

Let 𝗐intΘ\mathsf{w}\in\operatorname{int}\mathcal{L}_{\Theta}^{*} with 𝗐=1\|\mathsf{w}\|=1. Let π𝔞Θ𝔰Θ:𝔤𝔞Θ𝔰Θ\pi_{\mathfrak{a}_{\Theta}\oplus\mathfrak{s}_{\Theta}}:\mathfrak{g}\to\mathfrak{a}_{\Theta}\oplus\mathfrak{s}_{\Theta} and π𝔞Θ:𝔤𝔞Θ\pi_{\mathfrak{a}_{\Theta}}:\mathfrak{g}\to\mathfrak{a}_{\Theta} be orthogonal projection maps. Let

π𝗐:𝔤𝗐\displaystyle\pi_{\mathsf{w}}:\mathfrak{g}\to\mathbb{R}\mathsf{w}

denote the orthogonal projection map, i.e., the projection with respect to the decomposition 𝔤𝗐kerψ𝗏(𝗐)𝔞Θ\mathfrak{g}\cong\mathbb{R}\mathsf{w}\oplus\ker\psi_{\mathsf{v}(\mathsf{w})}\oplus\mathfrak{a}_{\Theta}^{\perp}. Similarly, let

π~𝗐:AΘSΘexp(𝗐)\displaystyle\tilde{\pi}_{\mathsf{w}}:A_{\Theta}S_{\Theta}\to\exp(\mathbb{R}\mathsf{w})

be the Cartesian projection map of AΘSΘexp(𝗐)×exp(kerψ𝗏(𝗐))×SΘA_{\Theta}S_{\Theta}\cong\exp(\mathbb{R}\mathsf{w})\times\exp(\ker\psi_{\mathsf{v}(\mathsf{w})})\times S_{\Theta} onto the exp(𝗐)\exp(\mathbb{R}\mathsf{w}) factor. Note that (dπ~𝗐)e=π𝗐|𝔞Θ𝔰Θ(d\tilde{\pi}_{\mathsf{w}})_{e}=\pi_{\mathsf{w}}|_{\mathfrak{a}_{\Theta}\oplus\mathfrak{s}_{\Theta}}.

Also define (NΘ)1,ϵe={nNΘ:𝖥(z1)n𝖥(WϵeΘ,ss(z1))}(N^{-}_{\Theta})_{1,\epsilon_{e}}=\left\{n^{-}\in N^{-}_{\Theta}:\mathsf{F}(z_{1})n^{-}\in\mathsf{F}\left(W_{\epsilon_{e}}^{\mathcal{F}_{\Theta},\mathrm{ss}}(z_{1})\right)\right\} where ϵe\epsilon_{e} is as in Subsection 5.1 and z1𝖱~1z_{1}\in\tilde{\mathsf{R}}_{1} is the center.

In order to derive the LNIC in Proposition 5.11, we need the following two lemmas regarding Ξ\Xi. We omit the proofs since they are almost a verbatim repetition of that of [64, Lemmas 6.2 and 6.3].

Lemma 5.4.

Let mm\in\mathbb{N}, α=(α0,α1,,αm1,1)\alpha=(\alpha_{0},\alpha_{1},\dotsc,\alpha_{m-1},1) be an admissible sequence, and n=nα(NΘ)1n^{-}=n_{\alpha}\in(N^{-}_{\Theta})_{1}. Let u𝖴~1(m)u\in\tilde{\mathsf{U}}_{1}^{(m)} and n+(NΘ+)1n^{+}\in(N^{+}_{\Theta})_{1} such that 𝖥(u)=𝖥(z1)n+\mathsf{F}(u)=\mathsf{F}(z_{1})n^{+} where z1𝖱~1z_{1}\in\tilde{\mathsf{R}}_{1} is the center. Then, we have

Ξ(n+,n)=a𝖪α(σα(z1))ϑα(σα(z1))1a𝖪α(σα(u))ϑα(σα(u)).\displaystyle\Xi(n^{+},n^{-})=a_{-\mathsf{K}_{\alpha}(\sigma^{-\alpha}(z_{1}))}\vartheta^{\alpha}(\sigma^{-\alpha}(z_{1}))^{-1}a_{\mathsf{K}_{\alpha}(\sigma^{-\alpha}(u))}\vartheta^{\alpha}(\sigma^{-\alpha}(u)).

In particular, for all 𝗐intΘ\mathsf{w}\in\operatorname{int}\mathcal{L}_{\Theta}^{*} with 𝗐=1\|\mathsf{w}\|=1, we have

logπ~𝗐(Ξ(n+,n))=(τα𝗐(σα(u))τα𝗐(σα(z1)))𝗐.\displaystyle\log\tilde{\pi}_{\mathsf{w}}(\Xi(n^{+},n^{-}))=\bigl{(}\tau^{\mathsf{w}}_{\alpha}(\sigma^{-\alpha}(u))-\tau^{\mathsf{w}}_{\alpha}(\sigma^{-\alpha}(z_{1}))\bigr{)}\mathsf{w}.
Lemma 5.5.

For all n(NΘ)1n^{-}\in(N^{-}_{\Theta})_{1}, we have

(dΞn)e=π𝔞Θ𝔰ΘAdn|𝔫Θ+(dhn)e,\displaystyle(d\Xi_{n^{-}})_{e}=\pi_{\mathfrak{a}_{\Theta}\oplus\mathfrak{s}_{\Theta}}\circ\operatorname{Ad}_{n^{-}}|_{\mathfrak{n}^{+}_{\Theta}}\circ(dh_{n^{-}})_{e},

and in particular, for all 𝗐intΘ\mathsf{w}\in\operatorname{int}\mathcal{L}_{\Theta}^{*} with 𝗐=1\|\mathsf{w}\|=1, we have

d(logπ~𝗐Ξn)e=π𝗐Adn|𝔫Θ+(dhn)e,\displaystyle d(\log\circ\tilde{\pi}_{\mathsf{w}}\circ\Xi_{n^{-}})_{e}=\pi_{\mathsf{w}}\circ\operatorname{Ad}_{n^{-}}|_{\mathfrak{n}^{+}_{\Theta}}\circ(dh_{n^{-}})_{e},

where hn:(NΘ+)1NΘ+h_{n^{-}}:(N^{+}_{\Theta})_{1}\to N^{+}_{\Theta} is a diffeomorphism onto its image which is also smooth in n(NΘ)1,ϵen^{-}\in(N^{-}_{\Theta})_{1,\epsilon_{e}} and satisfies hn(e)=eh_{n^{-}}(e)=e and he=Id(NΘ+)1h_{e}=\operatorname{Id}_{(N^{+}_{\Theta})_{1}}.

For all αΠ\alpha\in\Pi, let Hα𝔞H_{\alpha}\in\mathfrak{a} be the corresponding root direction, i.e., α=,Hα\alpha=\langle\cdot,H_{\alpha}\rangle. The following lemma is more or less a standard Lie algebra fact. We give a proof for the sake of completeness and for comparison with the analogous statement in [17, Proposition 4.4]. Using loc. cit. more directions can be produced (including those in 𝔪\mathfrak{m}) whereas our lemma below gives the stronger quantifier “for all αΠ\alpha\in\Pi and xα𝔤αx_{\alpha}\in\mathfrak{g}_{\alpha}” (cf. Remark 5.10).

Lemma 5.6.

Let αΠ\alpha\in\Pi. For all nonzero xα𝔤αx_{\alpha}\in\mathfrak{g}_{\alpha}, there exists a nonzero xα𝔤αx_{-\alpha}\in\mathfrak{g}_{-\alpha} such that [xα,xα]>0Hα[x_{-\alpha},x_{\alpha}]\in\mathbb{R}_{>0}H_{\alpha}.

Proof.

Let αΠ\alpha\in\Pi and xα𝔤αx_{\alpha}\in\mathfrak{g}_{\alpha} be nonzero. We will show that xα:=θ(xα)x_{-\alpha}:=\theta(x_{\alpha}) satisfies the lemma. Indeed xα𝔤αx_{-\alpha}\in\mathfrak{g}_{-\alpha} because for all H𝔞H\in\mathfrak{a}, we have

[H,xα]=θ[θ(H),θ(xα)]=θ[H,xα]=α(H)θ(xα)=α(H)xα.\displaystyle[H,x_{-\alpha}]=\theta[\theta(H),\theta(x_{-\alpha})]=\theta[-H,x_{\alpha}]=-\alpha(H)\theta(x_{\alpha})=-\alpha(H)x_{-\alpha}.

Thus, [xα,xα][𝔤α,𝔤α]𝔤0=𝔞𝔪[x_{-\alpha},x_{\alpha}]\in[\mathfrak{g}_{-\alpha},\mathfrak{g}_{\alpha}]\subset\mathfrak{g}_{0}=\mathfrak{a}\oplus\mathfrak{m}. Moreover, we have

θ[xα,xα]=[θ(xα),θ(xα)]=[xα,xα]=[xα,xα]\displaystyle\theta[x_{-\alpha},x_{\alpha}]=[\theta(x_{-\alpha}),\theta(x_{\alpha})]=[x_{\alpha},x_{-\alpha}]=-[x_{-\alpha},x_{\alpha}]

which implies [xα,xα]𝔭[x_{-\alpha},x_{\alpha}]\in\mathfrak{p}. Hence, [xα,xα](𝔞𝔪)𝔭=𝔞[x_{-\alpha},x_{\alpha}]\in(\mathfrak{a}\oplus\mathfrak{m})\cap\mathfrak{p}=\mathfrak{a}. We use the symbol \propto to mean proportionality with a positive constant. Recalling that ,|𝔞×𝔞B|𝔞×𝔞\langle\cdot,\cdot\rangle|_{\mathfrak{a}\times\mathfrak{a}}\propto B|_{\mathfrak{a}\times\mathfrak{a}}, for all H𝔞H\in\mathfrak{a}, we have

H,[xα,xα]B(H,[xα,xα])=B([H,xα],xα)=α(H)B(xα,xα)\displaystyle\langle H,[x_{-\alpha},x_{\alpha}]\rangle\propto B(H,[x_{-\alpha},x_{\alpha}])=B([H,x_{-\alpha}],x_{\alpha})=-\alpha(H)B(x_{-\alpha},x_{\alpha})

which implies [xα,xα]B(xα,xα)Hα=Bθ(xα,xα)HαHα{0}[x_{-\alpha},x_{\alpha}]\propto-B(x_{-\alpha},x_{\alpha})H_{\alpha}=B_{\theta}(x_{\alpha},x_{\alpha})H_{\alpha}\in\mathbb{R}H_{\alpha}-\{0\} since BθB_{\theta} is positive definite. ∎

Repeating an analogous proof for arbitrary Θ\Theta, we have the following generalization of [23, Lemma 2.11] which is itself a generalization of [70, Proposition 3.12]. Alternatively, [29, Theorem 1.1] (which is a stronger measure theoretic statement but for subvarieties) also suffices for our purposes in the proof of Lemma 5.8.

Lemma 5.7.

For any open subset 𝒪Θ\mathcal{O}\subset\mathcal{F}_{\Theta} with ΛΘ𝒪\Lambda_{\Theta}\cap\mathcal{O}\neq\varnothing, the intersection ΛΘ𝒪\Lambda_{\Theta}\cap\mathcal{O} is not contained in any smooth submanifold of Θ\mathcal{F}_{\Theta} of lower dimension.

We write the next lemma and its corollary in its most general form, though we apply it for 𝗐𝓌¯\mathsf{w}\in\overline{\mathscr{w}}. Let 𝗐intΘ\mathsf{w}\in\operatorname{int}\mathcal{L}_{\Theta}^{*} with 𝗐=1\|\mathsf{w}\|=1. We define

Θ𝗐\displaystyle\Theta_{\mathsf{w}} :={αΘ:𝗐kerα}Θ,\displaystyle:=\{\alpha\in\Theta:\mathsf{w}\notin\ker\alpha\}\subset\Theta,
𝔖𝗐±\displaystyle\mathfrak{S}^{\pm}_{\mathsf{w}} :=αΘ𝗐𝔤α𝔫Θ±.\displaystyle:=\bigoplus_{\alpha\in\Theta_{\mathsf{w}}}\mathfrak{g}_{\mp\alpha}\subset\mathfrak{n}^{\pm}_{\Theta}.

Since Θ𝗐\Theta_{\mathsf{w}} is always nonempty, if #Θ=1\#\Theta=1, then we trivially have Θ𝗐=Θ\Theta_{\mathsf{w}}=\Theta. Morevoer, 𝗐𝓌¯\mathsf{w}\in\overline{\mathscr{w}} implies Θ𝗐=Θ\Theta_{\mathsf{w}}=\Theta and 𝔖𝗐±=𝔫Θ±\mathfrak{S}^{\pm}_{\mathsf{w}}=\mathfrak{n}^{\pm}_{\Theta}.

Lemma 5.8.

There exists C𝓌>0C_{\mathscr{w}}>0 such that the following holds. Let 𝗐𝓌¯\mathsf{w}\in\overline{\mathscr{w}}. For all nonzero n+𝔖𝗐+n^{+}\in\mathfrak{S}^{+}_{\mathsf{w}}, there exists n(NΘ)1n^{-}\in(N^{-}_{\Theta})_{1} such that

|π𝗐(Adn(n+))|C𝓌.\displaystyle|\pi_{\mathsf{w}}(\operatorname{Ad}_{n^{-}}(n^{+}))|\geq C_{\mathscr{w}}. (15)
Proof.

By compactness of 𝓌¯\overline{\mathscr{w}} and continuity of π𝗐(v)\pi_{\mathsf{w}}(v) in 𝗐𝓌¯\mathsf{w}\in\overline{\mathscr{w}} for all v𝔤v\in\mathfrak{g}, it suffices to show that for all 𝗐𝓌¯\mathsf{w}\in\overline{\mathscr{w}} and nonzero n+𝔖𝗐+n^{+}\in\mathfrak{S}^{+}_{\mathsf{w}}, there exists n(NΘ)1n^{-}\in(N^{-}_{\Theta})_{1} such that π𝗐(Adn(n+))0\pi_{\mathsf{w}}(\operatorname{Ad}_{n^{-}}(n^{+}))\neq 0. Let n+𝔖𝗐+n^{+}\in\mathfrak{S}^{+}_{\mathsf{w}} be nonzero. Let 𝗐𝓌¯\mathsf{w}\in\overline{\mathscr{w}} and suppose for the sake of contradiction that π𝗐(Ad(NΘ)1(n+))=0\pi_{\mathsf{w}}\bigl{(}\operatorname{Ad}_{(N^{-}_{\Theta})_{1}}(n^{+})\bigr{)}=0. Without loss of generality, we may assume that 𝖥(z1)=e\mathsf{F}(z_{1})=e at the center z1𝖱~1z_{1}\in\tilde{\mathsf{R}}_{1}. Consider the smooth map

L:NΘ\displaystyle L:N^{-}_{\Theta} \displaystyle\to\mathbb{R}
n\displaystyle n^{-} π𝗐(Adn(n+)).\displaystyle\mapsto\pi_{\mathsf{w}}(\operatorname{Ad}_{n^{-}}(n^{+})).

Then, we have dLe(n^)=π𝗐([n^,n+])dL_{e}(\hat{n}^{-})=\pi_{\mathsf{w}}([\hat{n}^{-},n^{+}]) for all n^𝔫Θ\hat{n}^{-}\in\mathfrak{n}^{-}_{\Theta}.

Note that π𝔞Θ(Hα)kerψ𝗏(𝗐)=𝗐\pi_{\mathfrak{a}_{\Theta}}(H_{\alpha})\notin\ker\psi_{\mathsf{v}(\mathsf{w})}=\mathsf{w}^{\perp} for all αΘ𝗐\alpha\in\Theta_{\mathsf{w}}. Let xαx_{\alpha} for some αΘ𝗐\alpha\in\Theta_{\mathsf{w}} be a nonzero component in the decomposition of n+n^{+} according to 𝔖𝗐+=αΘ𝗐𝔤α\mathfrak{S}^{+}_{\mathsf{w}}=\bigoplus_{\alpha\in\Theta_{\mathsf{w}}}\mathfrak{g}_{-\alpha}. By Lemma 5.6, there exists xα𝔤α𝔫Θx_{-\alpha}\in\mathfrak{g}_{-\alpha}\subset\mathfrak{n}^{-}_{\Theta} such that [xα,xα]>0Hα[x_{-\alpha},x_{\alpha}]\in\mathbb{R}_{>0}H_{\alpha}. Using the definition of 𝔖𝗐+\mathfrak{S}^{+}_{\mathsf{w}}, we have π𝗐([xα,xα])0\pi_{\mathsf{w}}([x_{-\alpha},x_{\alpha}])\neq 0. Then dLe(xα)=π𝗐([xα,n+])=π𝗐([xα,xα])0dL_{e}(x_{-\alpha})=\pi_{\mathsf{w}}([x_{-\alpha},n^{+}])=\pi_{\mathsf{w}}([x_{-\alpha},x_{\alpha}])\neq 0 so there is a neighborhood ONΘO\subset N^{-}_{\Theta} containing ee such that L1(0)OL^{-1}(0)\cap O is a smooth submanifold (in fact a subvariety) of NΘN^{-}_{\Theta} of strictly lower dimension. However, (NΘ)1OL1(0)O(N^{-}_{\Theta})_{1}\cap O\subset L^{-1}(0)\cap O, but on the other hand ((NΘ)1)((N^{-}_{\Theta})_{1})^{-} contains Λ𝗂Θ𝒪\Lambda_{\mathsf{i}\Theta}\cap\mathcal{O} for some open set 𝒪𝗂Θ\mathcal{O}\subset\mathcal{F}_{\mathsf{i}\Theta}. It follows via the diffeomorphism NΘNΘeN^{-}_{\Theta}\to N^{-}_{\Theta}e^{-} that Λ𝗂Θ𝒪\Lambda_{\mathsf{i}\Theta}\cap\mathcal{O} is contained in a smooth submanifold of 𝗂Θ\mathcal{F}_{\mathsf{i}\Theta} of strictly lower dimension, contradicting Lemma 5.7. ∎

For any normed vector space (V,)(V,\|\cdot\|), let 𝕊(V)\mathbb{S}(V) denote the unit sphere in VV centered at 0V0\in V. Since Eq. 15 is an open condition and 𝕊(𝔫Θ+)\mathbb{S}(\mathfrak{n}^{+}_{\Theta}) is compact, a straightforward compactness argument gives the following corollary.

Corollary 5.9.

There exists C𝓌>0C_{\mathscr{w}}>0 such that the following holds. Let 𝗐𝓌¯\mathsf{w}\in\overline{\mathscr{w}}. There exist nontrivial n1,n2,,nj0(NΘ)1n_{1}^{-},n_{2}^{-},\dotsc,n_{j_{0}}^{-}\in(N^{-}_{\Theta})_{1} for some j0j_{0}\in\mathbb{N} and δ>0\delta>0 such that if η1,η2,,ηj0(NΘ)1\eta_{1}^{-},\eta_{2}^{-},\dotsc,\eta_{j_{0}}^{-}\in(N^{-}_{\Theta})_{1} with dNΘ(ηj,nj)δd_{N^{-}_{\Theta}}(\eta_{j}^{-},n_{j}^{-})\leq\delta for all 1jj01\leq j\leq j_{0}, then for all n+𝕊(𝔖𝗐+)n^{+}\in\mathbb{S}(\mathfrak{S}^{+}_{\mathsf{w}}), there exists 1jj01\leq j\leq j_{0} such that

|π𝗐(Adηj(n+))|C𝓌.\displaystyle\bigl{|}\pi_{\mathsf{w}}\bigl{(}\operatorname{Ad}_{\eta_{j}^{-}}(n^{+})\bigr{)}\bigr{|}\geq C_{\mathscr{w}}.
Remark 5.10.

It is important to note that although Corollary 5.9 is analogous to [64, Lemma 6.4], when projected to 𝗐\mathbb{R}\mathsf{w} for both results, the former has a stronger quantifier “for all n+𝕊(𝔖𝗐+)n^{+}\in\mathbb{S}(\mathfrak{S}^{+}_{\mathsf{w}})”. As a result, Proposition 5.11 also has a stronger quantifier and is hence stronger than the 𝗐\mathbb{R}\mathsf{w} projection of [64, Proposition 6.5]. Note that the stronger proposition is possible only for the “geodesic flow” and is ultimately rooted in Lemma 5.6. Due to this stronger proposition, we do not require the non-concentration property as in [64]. At a conceptual level, this is similar to what happens in [67].

With Lemmas 5.5 and 5.9 in hand, the proof of Proposition 5.11 is nearly the same as that of [64, Proposition 6.5]. Note that in the proposition, the open neighborhood 𝒰m\mathcal{U}_{m} can be chosen to be convex simply by shrinking it if necessary to an appropriate open ball. Also, its dependence on mm is not an issue since the proposition can be upgraded as discussed below.

Proposition 5.11 (LNIC).

There exist ε(0,1)\varepsilon\in(0,1), m0m_{0}\in\mathbb{N}, and j0j_{0}\in\mathbb{N} such that for all mm0m\geq m_{0}, there exists a convex open neighborhood 𝒰m𝖴~1(m)\mathcal{U}_{m}\subset\tilde{\mathsf{U}}_{1}^{(m)} of the center z1𝖱~1z_{1}\in\tilde{\mathsf{R}}_{1} such that there exist sections vj=σxj:𝖴~1(m)𝖴~xj,0𝗀xjv_{j}=\sigma^{-x_{j}}:\tilde{\mathsf{U}}_{1}^{(m)}\to{}^{\mathsf{g}^{-x_{j}}}\tilde{\mathsf{U}}_{x_{j,0}} for some mutually distinct admissible sequences xj=(xj,0,xj,1,,xj,m1,1)x_{j}=(x_{j,0},x_{j,1},\dotsc,x_{j,m-1},1) for all integers 0jj00\leq j\leq j_{0} such that for all 𝗐𝓌¯\mathsf{w}\in\overline{\mathscr{w}}, u𝒰mu\in\mathcal{U}_{m}, and Z𝕊(Tu(𝒰m))Z\in\mathbb{S}(\operatorname{T}_{u}(\mathcal{U}_{m})), there exist 1jj01\leq j\leq j_{0} such that

|(dφj,u)u(Z)|ε\displaystyle|(d\varphi_{j,u})_{u}(Z)|\geq\varepsilon

where we define φj:𝖴~1(m)×𝖴~1(m)\varphi_{j}:\tilde{\mathsf{U}}_{1}^{(m)}\times\tilde{\mathsf{U}}_{1}^{(m)}\to\mathbb{R} by

φj(u,u)\displaystyle\varphi_{j}(u,u^{\prime}) =(τxj𝗐vjτx0𝗐v0)(u)(τxj𝗐vjτx0𝗐v0)(u)\displaystyle=\bigl{(}\tau^{\mathsf{w}}_{x_{j}}\circ v_{j}-\tau^{\mathsf{w}}_{x_{0}}\circ v_{0}\bigr{)}(u)-\bigl{(}\tau^{\mathsf{w}}_{x_{j}}\circ v_{j}-\tau^{\mathsf{w}}_{x_{0}}\circ v_{0}\bigr{)}(u^{\prime})

and we denote φj,u=φj(u,)\varphi_{j,u}=\varphi_{j}(u,\cdot) for all u,u𝖴~1(m)u,u^{\prime}\in\tilde{\mathsf{U}}_{1}^{(m)} and 1jj01\leq j\leq j_{0}.

We immediately derive the following corollary from Proposition 5.11 first on the open neighborhood int(𝖴1)𝒰m𝖴1\operatorname{int}(\mathsf{U}_{1})\cap\mathcal{U}_{m}\subset\mathsf{U}_{1} of the center z1𝖱1z_{1}\in\mathsf{R}_{1} for a given mm0m\geq m_{0} since 𝒰m\mathcal{U}_{m} is convex and then upgrade it to int(U)\operatorname{int}(U) using the topological mixing property of the transition matrix TT (see Subsection 4.3). The mutual disjointness in Proposition 5.12 follows from mutual distinctness of the admissible sequences in Proposition 5.11.

Proposition 5.12.

There exist ε(0,1)\varepsilon\in(0,1), an unbounded subset 𝓂\mathscr{m}\subset\mathbb{N}, and j0j_{0}\in\mathbb{N} such that for all m𝓂m\in\mathscr{m}, there exists a set of Lipschitz sections {vj:UU}j=0j0\{v_{j}:U\to U\}_{j=0}^{j_{0}} of σm\sigma^{m} such that for all 𝗐𝓌¯\mathsf{w}\in\overline{\mathscr{w}} and u,uint(U)u,u^{\prime}\in\operatorname{int}(U), there exist 1jj01\leq j\leq j_{0} such that

|(τm𝗐vjτm𝗐v0)(u)(τm𝗐vjτm𝗐v0)(u)|εd(u,u).\displaystyle|(\tau^{\mathsf{w}}_{m}\circ v_{j}-\tau^{\mathsf{w}}_{m}\circ v_{0})(u)-(\tau^{\mathsf{w}}_{m}\circ v_{j}-\tau^{\mathsf{w}}_{m}\circ v_{0})(u^{\prime})|\geq\varepsilon d(u,u^{\prime}).

Moreover, v0(U),v1(U),,vj0(U)v_{0}(U),v_{1}(U),\dotsc,v_{j_{0}}(U) are mutually disjoint.

Fix ε\varepsilon, 𝓂\mathscr{m}, and j0j_{0} to be the ones provided by Proposition 5.12 henceforth.

6. Dolgopyat operators and the proof of Theorem 4.8

In this section we carry out Dolgopyat’s method to prove Theorem 4.8. Many of the techniques used here go all the way back to Dolgopyat [22]. Since it is sensitive to the setting at hand, some of the arguments require careful treatment in each instance. Over the years, it has been made cleaner and more efficient in some settings such as by Stoyanov [67].

Thanks to Remark 5.10, we are able to follow Stoyanov’s strategy using cylinders and the new distance DD. One especially important advantage of Stoyanov’s version of Dolgopyat’s method is that we do not require the Federer/doubling property. This is not just a convenience—there is no appropriate metric to use on 𝔫Θ+\mathfrak{n}^{+}_{\Theta} which is also compatible with UU and a doubling property is not known yet for PS measures in higher rank to the best of the authors’ knowledge. The reader may also find it useful to consult [43, 62, 63] where many full proofs are provided in a similar framework.

We fix 𝗐𝓌\mathsf{w}\in\mathscr{w} for the rest of the section. Recall that this implies Θ𝗐=Θ\Theta_{\mathsf{w}}=\Theta and hence, Proposition 5.12 applies in this section.

6.1. Preliminary lemmas and constants

We obtain the following eventually contracting property, which also implies the same for the new distance function DD, from the metric Anosov property of the translation flow in Theorem 3.2, and a compactness argument for the lower bound. Another lemma follows from it as in [67, Proposition 3.3].

Lemma 6.1.

There exist c0(0,1)c_{0}\in(0,1) and κ1>κ2>1\kappa_{1}>\kappa_{2}>1 such that for all cylinders 𝙲U\mathtt{C}\subset U with len(𝙲)=j\operatorname{len}(\mathtt{C})=j\in\mathbb{N}, and u,u𝙲u,u^{\prime}\in\mathtt{C}, we have both

c0κ2jd(u,u)d(σj(u),σj(u))c01κ1jd(u,u)\displaystyle c_{0}\kappa_{2}^{j}d(u,u^{\prime})\leq d(\sigma^{j}(u),\sigma^{j}(u^{\prime}))\leq{c_{0}}^{-1}\kappa_{1}^{j}d(u,u^{\prime})
c0κ2jD(u,u)D(σj(u),σj(u))c01κ1jD(u,u).\displaystyle c_{0}\kappa_{2}^{j}D(u,u^{\prime})\leq D(\sigma^{j}(u),\sigma^{j}(u^{\prime}))\leq{c_{0}}^{-1}\kappa_{1}^{j}D(u,u^{\prime}).
Lemma 6.2.

There exist p0p_{0}\in\mathbb{N} and ρ(0,1)\rho\in(0,1) such that for all cylinders 𝙲U\mathtt{C}\subset U with len(𝙲)=l0\operatorname{len}(\mathtt{C})=l\in\mathbb{Z}_{\geq 0} and subcylinders 𝙲,𝙲′′𝙲\mathtt{C}^{\prime},\mathtt{C}^{\prime\prime}\subset\mathtt{C} with len(𝙲)=l+1\operatorname{len}(\mathtt{C}^{\prime})=l+1 and len(𝙲′′)=l+p0\operatorname{len}(\mathtt{C}^{\prime\prime})=l+p_{0}, we have

diam(𝙲′′)ρdiam(𝙲)diam(𝙲).\displaystyle\operatorname{diam}(\mathtt{C}^{\prime\prime})\leq\rho\operatorname{diam}(\mathtt{C})\leq\operatorname{diam}(\mathtt{C}^{\prime}).

We fix constants c0c_{0}, κ1\kappa_{1}, κ2\kappa_{2}, p0p_{0}, and ρ\rho provided by Lemmas 6.1 and 6.2 henceforth. We use these bounds extensively without comments. We also fix p1p_{1}\in\mathbb{N} such that

122ρp11116.\displaystyle\frac{1}{2}-2\rho^{p_{1}-1}\geq\frac{1}{16}. (16)

The following is a Lasota–Yorke [37] type of lemma whose proof is similar to that of [67, Lemma 5.4] and [43, Lemma 3.9]. Its proof gives the explicit constant A0>2c01eT𝗐c0(κ21)max(1,T𝗐κ21)>2A_{0}>2c_{0}^{-1}e^{\frac{T_{\mathsf{w}}}{c_{0}(\kappa_{2}-1)}}\max\bigl{(}1,\frac{T_{\mathsf{w}}}{\kappa_{2}-1}\bigr{)}>2 which we fix henceforth.

Lemma 6.3.

There exists A0>0A_{0}>0 such that for all |a|<a0|a|<a_{0}^{\prime}, |b|>1|b|>1, and kk\in\mathbb{N}, we have

  1. ((1))

    if h𝒞B(U)h\in\mathcal{C}_{B}(U) (resp. h𝒞~B(U)h\in\tilde{\mathcal{C}}_{B}(U)) for some B>0B>0, then a,𝗐k(h)𝒞B(U)\mathcal{L}_{a,\mathsf{w}}^{k}(h)\in\mathcal{C}_{B^{\prime}}(U) (resp. a,𝗐k(h)𝒞~B(U)\mathcal{L}_{a,\mathsf{w}}^{k}(h)\in\tilde{\mathcal{C}}_{B^{\prime}}(U)) for B=A0(Bκ2k+1)B^{\prime}=A_{0}\left(\frac{B}{\kappa_{2}^{k}}+1\right);

  2. ((2))

    if HC(U,)H\in C(U,\mathbb{R}) and hB(U,)h\in B(U,\mathbb{R}) satisfy

    H(u)H(u)2Bh(u)D(u,u)\displaystyle\|H(u)-H(u^{\prime})\|_{2}\leq Bh(u)D(u,u^{\prime})

    for all u,uUju,u^{\prime}\in U_{j}, for all j𝒜j\in\mathcal{A}, for some B>0B>0, then for all j𝒜j\in\mathcal{A}, for all u,uUju,u^{\prime}\in U_{j}, we have

    ξ,𝗐k(H)(u)ξ,𝗐k(H)(u)2A0(Bκ2ka,𝗐k(h)(u)+|b|a,𝗐kH(u))D(u,u)\displaystyle\left\|\mathcal{L}_{\xi,\mathsf{w}}^{k}(H)(u)-\mathcal{L}_{\xi,\mathsf{w}}^{k}(H)(u^{\prime})\right\|_{2}\leq A_{0}\left(\frac{B}{\kappa_{2}^{k}}\mathcal{L}_{a,\mathsf{w}}^{k}(h)(u)+|b|\mathcal{L}_{a,\mathsf{w}}^{k}\|H\|(u)\right)D(u,u^{\prime})

    for all u,uUu,u^{\prime}\in U.

Now, we fix b0=1b_{0}=1 and some other significant positive constants

E\displaystyle E >2A0>4,\displaystyle>2A_{0}>4, (17)
ϵ1\displaystyle\epsilon_{1} <min(δ^,πc0(κ21)2T𝗐),\displaystyle<\min\left(\hat{\delta},\frac{\pi c_{0}(\kappa_{2}-1)}{2T_{\mathsf{w}}}\right), (18)
m\displaystyle m 𝓂 such that κ2m>max(8A0,4Eϵ1ρp1c0,4128Ec0ερ),\displaystyle\in\mathscr{m}\text{ such that }\kappa_{2}^{m}>\max\left(8A_{0},\frac{4E\epsilon_{1}\rho^{p_{1}}}{c_{0}},\frac{4\cdot 128E}{c_{0}\varepsilon\rho}\right), (19)
μ\displaystyle\mu <min(2Eϵ1c0ρp0p1+1κ1m,14,11616e2mT𝗐(ερϵ164)2).\displaystyle<\min\left(\frac{2E\epsilon_{1}c_{0}\rho^{p_{0}p_{1}+1}}{\kappa_{1}^{m}},\frac{1}{4},\frac{1}{16\cdot 16e^{2mT_{\mathsf{w}}}}\left(\frac{\varepsilon\rho\epsilon_{1}}{64}\right)^{2}\right). (20)

Having fixed m𝓂m\in\mathscr{m}, we also fix the corresponding set of Lipschitz sections {vj:UU}j=0j0\{v_{j}:U\to U\}_{j=0}^{j_{0}} of σm\sigma^{m} provided by Proposition 5.12.

6.2. Construction of Dolgopyat operators

Let |b|>b0|b|>b_{0}. We define the set {𝙲1(b),𝙲2(b),,𝙲cb(b)}\{\mathtt{C}_{1}(b),\mathtt{C}_{2}(b),\dotsc,\mathtt{C}_{c_{b}}(b)\} for some cbc_{b}\in\mathbb{N} consisting of maximal cylinders 𝙲U\mathtt{C}\subset U with diam(𝙲)ϵ1|b|\operatorname{diam}(\mathtt{C})\leq\frac{\epsilon_{1}}{|b|} so that U=l=1cb𝙲l(b)¯U=\bigcup_{l=1}^{c_{b}}\overline{\mathtt{C}_{l}(b)}. We define the set {𝙳1(b),𝙳2(b),,𝙳db(b)}\{\mathtt{D}_{1}(b),\mathtt{D}_{2}(b),\dotsc,\mathtt{D}_{d_{b}}(b)\} for some dbd_{b}\in\mathbb{N} consisting of subcylinders 𝙳𝙲l(b)\mathtt{D}\subset\mathtt{C}_{l}(b) with len(𝙳)=len(𝙲l(b))+p0p1\operatorname{len}(\mathtt{D})=\operatorname{len}(\mathtt{C}_{l}(b))+p_{0}p_{1} and 1lcb1\leq l\leq c_{b}. We define the index set Ξ(b)={0,1,,j0}×{1,2,,db}\Xi(b)=\{0,1,\dotsc,j_{0}\}\times\{1,2,\dotsc,d_{b}\}. We define 𝚇j,k(b)=vj(𝙳k(b))¯\mathtt{X}_{j,k}(b)=\overline{v_{j}(\mathtt{D}_{k}(b))} which satisfies 𝚇j,k(b)𝚇j,k(b)=\mathtt{X}_{j,k}(b)\cap\mathtt{X}_{j^{\prime},k^{\prime}}(b)=\varnothing for all distinct (j,k),(j,k)Ξ(b)(j,k),(j^{\prime},k^{\prime})\in\Xi(b). For all JΞ(b)J\subset\Xi(b), we define the function

βJ=χUμ(j,k)Jχ𝚇j,k(b).\displaystyle\beta_{J}=\chi_{U}-\mu\sum_{(j,k)\in J}\chi_{\mathtt{X}_{j,k}(b)}.

We record a number of basic facts derived from Lemmas 6.1 and 6.2.

  • For all 1lcb1\leq l\leq c_{b} and (j,k)Ξ(b)(j,k)\in\Xi(b), we have the diameter bounds:

    ρϵ1|b|diam(𝙲l(b))ϵ1|b|\displaystyle\rho\frac{\epsilon_{1}}{|b|}\leq\operatorname{diam}(\mathtt{C}_{l}(b))\leq\frac{\epsilon_{1}}{|b|} (21)
    ρp0p1+1ϵ1|b|diam(𝙳k(b))ρp1ϵ1|b|\displaystyle\rho^{p_{0}p_{1}+1}\frac{\epsilon_{1}}{|b|}\leq\operatorname{diam}(\mathtt{D}_{k}(b))\leq\rho^{p_{1}}\frac{\epsilon_{1}}{|b|} (22)
    ϵ1c0ρp0p1+1|b|κ1mdiam(𝚇j,k(b))ϵ1ρp1|b|c0κ2m.\displaystyle\frac{\epsilon_{1}c_{0}\rho^{p_{0}p_{1}+1}}{|b|\kappa_{1}^{m}}\leq\operatorname{diam}(\mathtt{X}_{j,k}(b))\leq\frac{\epsilon_{1}\rho^{p_{1}}}{|b|c_{0}\kappa_{2}^{m}}. (23)
  • For all JΞ(b)J\subset\Xi(b), we have βJLD(U,)\beta_{J}\in L_{D}(U,\mathbb{R}) with Lipschitz constant (cf. [67, Lemma 5.2]):

    LipD(βJ)μmin(j,k)Jdiam(𝚇j,k(b))μ|b|κ1mϵ1c0ρp0p1+1.\displaystyle\operatorname{Lip}_{D}(\beta_{J})\leq\frac{\mu}{\min_{(j,k)\in J}\operatorname{diam}(\mathtt{X}_{j,k}(b))}\leq\frac{\mu|b|\kappa_{1}^{m}}{\epsilon_{1}c_{0}\rho^{p_{0}p_{1}+1}}. (24)
  • For all 1lcb1\leq l\leq c_{b}, if u,u𝙲l(b)u,u^{\prime}\in\mathtt{C}_{l}(b), then

    D(vj(u),vj(u))ϵ1|b|c0κ2m.\displaystyle D(v_{j}(u),v_{j}(u^{\prime}))\leq\frac{\epsilon_{1}}{|b|c_{0}\kappa_{2}^{m}}. (25)
Definition 6.4.

For all ξ\xi\in\mathbb{C} with |a|<a0|a|<a_{0}^{\prime} and |b|>b0|b|>b_{0}, and JΞ(b)J\subset\Xi(b), we define the Dolgopyat operators 𝒩a,J𝗐:LD(U,)LD(U,)\mathcal{N}_{a,J}^{\mathsf{w}}:L_{D}(U,\mathbb{R})\to L_{D}(U,\mathbb{R}) by

𝒩a,J𝗐(h)=am(βJh)\displaystyle\mathcal{N}_{a,J}^{\mathsf{w}}(h)=\mathcal{L}_{a}^{m}(\beta_{J}h)

for all hLD(U,)h\in L_{D}(U,\mathbb{R}).

Definition 6.5.

For all |b|>b0|b|>b_{0}, a subset JΞ(b)J\subset\Xi(b) is said to be dense if for all 1lcb1\leq l\leq c_{b}, there exists (j,k)J(j,k)\in J such that 𝙳k(b)𝙲l(b)\mathtt{D}_{k}(b)\subset\mathtt{C}_{l}(b). Denote 𝒥(b)\mathcal{J}(b) to be the set of all dense subsets of Ξ(b)\Xi(b).

6.3. Proof of Theorem 4.8

Properties (1) and (3)(3)(2b) in Theorem 4.8 are derived from Lemma 6.3 using estimates from Eqs. 17, 19, and 20. We omit the proofs of since they are almost identical to those in [67, 43, 62, 63].

Similarly, we also omit the proof of Property (2) in Theorem 4.8 and refer the reader to [67, 62, 63]. However, we mention that the proof of Property (2) uses the Gibbs property of the measure νU𝗐\nu_{U}^{\mathsf{w}} which is automatically satisfied since it comes from a equilibrium state. This replaces the role of the Federer/doubling property.

For the sake of completeness, we include the proof of Property (3)(3)(2a) in Theorem 4.8 where the crucial LNIC from Proposition 5.12 is used. Here we use Θ𝗐=Θ\Theta_{\mathsf{w}}=\Theta.

Lemma 6.6.

Let |b|>b0|b|>b_{0}. Suppose 𝙳k(b),𝙳k(b)𝙲l(b)\mathtt{D}_{k}(b),\mathtt{D}_{k^{\prime}}(b)\subset\mathtt{C}_{l}(b) for some 1k,kdb1\leq k,k^{\prime}\leq d_{b} and 1lcb1\leq l\leq c_{b} such that d(u0,u0)12diam(𝙲l(b))d(u_{0},u_{0}^{\prime})\geq\frac{1}{2}\operatorname{diam}(\mathtt{C}_{l}(b)) for some u0𝙳k(b)u_{0}\in\mathtt{D}_{k}(b) and u0𝙳k(b)u_{0}^{\prime}\in\mathtt{D}_{k^{\prime}}(b). Then, there exists 1jj01\leq j\leq j_{0} such that

ερϵ116|b||(τm𝗐vjτm𝗐v0)(u)(τm𝗐vjτm𝗐v0)(u)|π\displaystyle\frac{\varepsilon\rho\epsilon_{1}}{16}\leq|b|\cdot|(\tau^{\mathsf{w}}_{m}\circ v_{j}-\tau^{\mathsf{w}}_{m}\circ v_{0})(u)-(\tau^{\mathsf{w}}_{m}\circ v_{j}-\tau^{\mathsf{w}}_{m}\circ v_{0})(u^{\prime})|\leq\pi

for all u𝙳k(b)u\in\mathtt{D}_{k}(b) and u𝙳k(b)u^{\prime}\in\mathtt{D}_{k^{\prime}}(b).

Proof.

Let u0u_{0}, u0u_{0}^{\prime}, uu, and uu^{\prime} be as in the lemma. The upper bound always holds on 𝙲l(b)\mathtt{C}_{l}(b) by Eq. 18. Recall that 𝗐𝓌\mathsf{w}\in\mathscr{w} and hence Θ𝗐=Θ\Theta_{\mathsf{w}}=\Theta in this section. Thus, the lemma follows from Proposition 5.12 and the estimate

d(u,u)d(u0,u0)d(u,u0)d(u,u0)ρϵ12|b|2ρp1ϵ1|b|ρϵ116|b|\displaystyle d(u,u^{\prime})\geq d(u_{0},u_{0}^{\prime})-d(u,u_{0})-d(u^{\prime},u_{0}^{\prime})\geq\frac{\rho\epsilon_{1}}{2|b|}-2\frac{\rho^{p_{1}}\epsilon_{1}}{|b|}\geq\frac{\rho\epsilon_{1}}{16|b|}

using Eq. 16. ∎

Now, for all ξ\xi\in\mathbb{C} with |a|<a0|a|<a_{0}^{\prime} and |b|>b0|b|>b_{0}, for all integers 1jj01\leq j\leq j_{0}, for all HL(U,)H\in L(U,\mathbb{C}) and for all h𝒞E|b|(U)h\in\mathcal{C}_{E|b|}(U), we define the functions χj,0[ξ,H,h],χj,1[ξ,H,h]:U\chi_{j,0}^{[\xi,H,h]},\chi_{j,1}^{[\xi,H,h]}:U\to\mathbb{R} by

χj,0[ξ,H,h](u)\displaystyle\chi_{j,0}^{[\xi,H,h]}(u) =|e(τm𝗐,(a)+ibτm𝗐)(vj(u))H(vj(u))+e(τm𝗐,(a)+ibτm𝗐)(v0(u))H(v0(u))|eτm𝗐,(a)(vj(u))h(vj(u))+(1μ)eτm𝗐,(a)(v0(u))h(v0(u)),\displaystyle=\frac{\bigl{|}e^{(\tau_{m}^{\mathsf{w},(a)}+ib\tau^{\mathsf{w}}_{m})(v_{j}(u))}H(v_{j}(u))+e^{(\tau_{m}^{\mathsf{w},(a)}+ib\tau^{\mathsf{w}}_{m})(v_{0}(u))}H(v_{0}(u))\bigr{|}}{e^{\tau_{m}^{\mathsf{w},(a)}(v_{j}(u))}h(v_{j}(u))+(1-\mu)e^{\tau_{m}^{\mathsf{w},(a)}(v_{0}(u))}h(v_{0}(u))},
χj,1[ξ,H,h](u)\displaystyle\chi_{j,1}^{[\xi,H,h]}(u) =|e(τm𝗐,(a)+ibτm𝗐)(vj(u))H(vj(u))+e(τm𝗐,(a)+ibτm𝗐)(v0(u))H(v0(u))|(1μ)eτm𝗐,(a)(vj(u))h(vj(u))+eτm𝗐,(a)(v0(u))h(v0(u))\displaystyle=\frac{\bigl{|}e^{(\tau_{m}^{\mathsf{w},(a)}+ib\tau^{\mathsf{w}}_{m})(v_{j}(u))}H(v_{j}(u))+e^{(\tau_{m}^{\mathsf{w},(a)}+ib\tau^{\mathsf{w}}_{m})(v_{0}(u))}H(v_{0}(u))\bigr{|}}{(1-\mu)e^{\tau_{m}^{\mathsf{w},(a)}(v_{j}(u))}h(v_{j}(u))+e^{\tau_{m}^{\mathsf{w},(a)}(v_{0}(u))}h(v_{0}(u))}

for all uUu\in U. We need another lemma before proving Property (3)(3)(2a) in Theorem 4.8. See [67, 62, 63, 43] for a proof.

Lemma 6.7.

Let |b|>b0|b|>b_{0}. Suppose HL(U,)H\in L(U,\mathbb{C}) and h𝒞E|b|(U)h\in\mathcal{C}_{E|b|}(U) satisfy Properties (3)(3)(1a) and (3)(3)(1b) in Theorem 4.8. Then for all (j,k)Ξ(b)(j,k)\in\Xi(b), we have

12h(vj(u))h(vj(u))2for all u,u𝙳k(b)\displaystyle\frac{1}{2}\leq\frac{h(v_{j}(u))}{h(v_{j}(u^{\prime}))}\leq 2\qquad\text{for all $u,u^{\prime}\in\mathtt{D}_{k}(b)$}

and also either of the alternatives

  1. (1)

    |H(vj(u))|34h(vj(u))|H(v_{j}(u))|\leq\frac{3}{4}h(v_{j}(u)) for all u𝙳k(b)u\in\mathtt{D}_{k}(b),

  2. (2)

    |H(vj(u))|14h(vj(u))|H(v_{j}(u))|\geq\frac{1}{4}h(v_{j}(u)) for all u𝙳k(b)u\in\mathtt{D}_{k}(b).

For any w1,w2{0}w_{1},w_{2}\in\mathbb{C}-\{0\}, let (w1,w2)[0,π]\angle(w_{1},w_{2})\in[0,\pi] denote the angle between w1w_{1} and w2w_{2} viewed as vectors in 2\mathbb{C}\cong\mathbb{R}^{2}.

Lemma 6.8.

Suppose w1,w2{0}w_{1},w_{2}\in\mathbb{C}-\{0\} such that (w1,w2)α\angle(w_{1},w_{2})\geq\alpha and |w1||w2|L\frac{|w_{1}|}{|w_{2}|}\leq L for some α[0,π]\alpha\in[0,\pi] and L1L\geq 1. Then we have

|w1+w2|(1α216L)|w1|+|w2|.\displaystyle|w_{1}+w_{2}|\leq\left(1-\frac{\alpha^{2}}{16L}\right)|w_{1}|+|w_{2}|.
Lemma 6.9.

Let ξ\xi\in\mathbb{C} with |a|<a0|a|<a_{0}^{\prime} and |b|>b0|b|>b_{0}. Suppose HL(U,)H\in L(U,\mathbb{C}) and h𝒞E|b|(U)h\in\mathcal{C}_{E|b|}(U) satisfy Properties (3)(3)(1a) and (3)(3)(1b) in Theorem 4.8. For all integers 1lcb1\leq l\leq c_{b}, there exists (j,k)Ξ(b)(j,k)\in\Xi(b) such that 𝙳k(b)𝙲l(b)\mathtt{D}_{k}(b)\subset\mathtt{C}_{l}(b) and such that χj,0[ξ,H,h](u)1\chi_{j,0}^{[\xi,H,h]}(u)\leq 1 or χj,1[ξ,H,h](u)1\chi_{j,1}^{[\xi,H,h]}(u)\leq 1 for all u𝙳k(b)u\in\mathtt{D}_{k}(b).

Proof.

Let ξ\xi, HH, hh, and ll be as in the lemma. Suppose Alternative (1) in Lemma 6.7 holds for some (j,k)Ξ(b)(j,k)\in\Xi(b). Then it is a straightforward calculation to check that χj,1[ξ,H,h](u)1\chi_{j,1}^{[\xi,H,h]}(u)\leq 1 for all u𝙳k(b)u\in\mathtt{D}_{k}(b), using Eq. 20. Otherwise, Alternative (2) in Lemma 6.7 holds for all (j,k)Ξ(b)(j,k)\in\Xi(b). Choose 1k,kdb1\leq k,k^{\prime}\leq d_{b} satisfying the hypotheses in Lemma 6.6 and 1jj01\leq j\leq j_{0} satisfying the conclusion of Lemma 6.6. Let u𝙳k(b)u\in\mathtt{D}_{k}(b) and u𝙳k(b)u^{\prime}\in\mathtt{D}_{k^{\prime}}(b). Note that |H(v(u))|,|H(v(u))|>0|H(v_{\ell}(u))|,|H(v_{\ell}(u^{\prime}))|>0 for all {0,j}\ell\in\{0,j\}. We would like to apply Lemma 6.8 but first we need to establish bounds on relative angle and relative size. We start with the former. For all {0,j}\ell\in\{0,j\}, let u{u,u}u_{\ell}\in\{u,u^{\prime}\} such that |H(v(u))|=min(|H(v(u))|,|H(v(u))|)|H(v_{\ell}(u_{\ell}))|=\min(|H(v_{\ell}(u))|,|H(v_{\ell}(u^{\prime}))|). Then recalling the hypotheses for HH and hh, and Eqs. 25 and 19, for all {0,j}\ell\in\{0,j\}, we have

|H(v(u))H(v(u))|min(|H(v(u))|,|H(v(u))|)\displaystyle\frac{|H(v_{\ell}(u))-H(v_{\ell}(u^{\prime}))|}{\min\left(|H(v_{\ell}(u))|,|H(v_{\ell}(u^{\prime}))|\right)} E|b|h(v(u))D(v(u),v(u))|H(v(u))|\displaystyle\leq\frac{E|b|h(v_{\ell}(u_{\ell}))D(v_{\ell}(u),v_{\ell}(u^{\prime}))}{|H(v_{\ell}(u_{\ell}))|}
4E|b|ϵ1|b|c0κ2mερϵ1128.\displaystyle\leq 4E|b|\cdot\frac{\epsilon_{1}}{|b|c_{0}\kappa_{2}^{m}}\leq\frac{\varepsilon\rho\epsilon_{1}}{128}.

Using some elementary geometry, the above shows that sin((H(v(u)),H(v(u))))ερϵ1128\sin(\angle(H(v_{\ell}(u)),H(v_{\ell}(u^{\prime}))))\leq\frac{\varepsilon\rho\epsilon_{1}}{128} with (H(v(u)),H(v(u)))[0,π2)\angle(H(v_{\ell}(u)),H(v_{\ell}(u^{\prime})))\in[0,\frac{\pi}{2}), for all {0,j}\ell\in\{0,j\} and hence,

(H(v(u)),H(v(u)))2sin((H(v(u)),H(v(u))))ερϵ164\angle(H(v_{\ell}(u)),H(v_{\ell}(u^{\prime})))\leq 2\sin(\angle(H(v_{\ell}(u)),H(v_{\ell}(u^{\prime}))))\leq\frac{\varepsilon\rho\epsilon_{1}}{64}

for all {0,j}\ell\in\{0,j\}. For notational convenience, we define φ:U\varphi:U\to\mathbb{R} by

φ(w)=b(τm𝗐vj(w)τm𝗐v0(w))for all wU.\varphi(w)=b(\tau^{\mathsf{w}}_{m}\circ v_{j}(w)-\tau^{\mathsf{w}}_{m}\circ v_{0}(w))\qquad\text{for all $w\in U$}.

By Lemma 6.6, we have ερϵ116|φ(u)φ(u)|π\frac{\varepsilon\rho\epsilon_{1}}{16}\leq|\varphi(u)-\varphi(u^{\prime})|\leq\pi. The second inequality is to ensure that we take the correct branch of angle in the following calculations. We will use these bounds to obtain a lower bound for (Vj(u),V0(u))\angle(V_{j}(u),V_{0}(u)) or (Vj(u),V0(u))\angle(V_{j}(u^{\prime}),V_{0}(u^{\prime})) where we define

V(w)=e(τm𝗐,(a)+ibτm𝗐)(v(w))H(v(w))for all wU and {0,j}.\displaystyle V_{\ell}(w)=e^{(\tau_{m}^{\mathsf{w},(a)}+ib\tau^{\mathsf{w}}_{m})(v_{\ell}(w))}H(v_{\ell}(w))\qquad\text{for all $w\in U$ and $\ell\in\{0,j\}$}.

Using the triangle inequality and previously computed bounds, we have

(Vj(u),V0(u))=(eiφ(u)H(vj(u)),H(v0(u)))\displaystyle\angle\left(V_{j}(u),V_{0}(u)\right)=\angle\bigl{(}e^{i\varphi(u)}H(v_{j}(u)),H(v_{0}(u))\bigr{)}
\displaystyle\geq{} (eiφ(u)H(vj(u)),eiφ(u)H(vj(u)))(eiφ(u)H(vj(u)),eiφ(u)H(vj(u)))\displaystyle\angle\bigl{(}e^{i\varphi(u)}H(v_{j}(u)),e^{i\varphi(u^{\prime})}H(v_{j}(u))\bigr{)}-\angle\bigl{(}e^{i\varphi(u^{\prime})}H(v_{j}(u)),e^{i\varphi(u^{\prime})}H(v_{j}(u^{\prime}))\bigr{)}
(H(v0(u)),H(v0(u)))(eiφ(u)H(vj(u)),H(v0(u)))\displaystyle{}-\angle(H(v_{0}(u)),H(v_{0}(u^{\prime})))-\angle\bigl{(}e^{i\varphi(u^{\prime})}H(v_{j}(u^{\prime})),H(v_{0}(u^{\prime}))\bigr{)}
=\displaystyle={} |φ(u)φ(u)|(H(vj(u)),H(vj(u)))(H(v0(u)),H(v0(u)))\displaystyle|\varphi(u)-\varphi(u^{\prime})|-\angle\left(H(v_{j}(u)),H(v_{j}(u^{\prime}))\right)-\angle(H(v_{0}(u)),H(v_{0}(u^{\prime})))
(Vj(u),V0(u))\displaystyle{}-\angle(V_{j}(u^{\prime}),V_{0}(u^{\prime}))
\displaystyle\geq{} ερϵ116ερϵ164ερϵ164(Vj(u),V0(u))\displaystyle\frac{\varepsilon\rho\epsilon_{1}}{16}-\frac{\varepsilon\rho\epsilon_{1}}{64}-\frac{\varepsilon\rho\epsilon_{1}}{64}-\angle(V_{j}(u^{\prime}),V_{0}(u^{\prime}))

and hence,

(Vj(u),V0(u))+(Vj(u),V0(u))ερϵ132\displaystyle\angle\left(V_{j}(u),V_{0}(u)\right)+\angle\left(V_{j}(u^{\prime}),V_{0}(u^{\prime})\right)\geq\frac{\varepsilon\rho\epsilon_{1}}{32}

for all u𝙳k(b)u\in\mathtt{D}_{k}(b) and u𝙳k(b)u^{\prime}\in\mathtt{D}_{k^{\prime}}(b). Without loss of generality, we may assume that (Vj(u),V0(u))ερϵ164\angle(V_{j}(u),V_{0}(u))\geq\frac{\varepsilon\rho\epsilon_{1}}{64} for all u𝙳k(b)u\in\mathtt{D}_{k}(b), which establishes the required bound on relative angle. For the bound on relative size, let (,){(0,j),(j,0)}(\ell,\ell^{\prime})\in\{(0,j),(j,0)\} such that h(v(u0))h(v(u0))h(v_{\ell}(u_{0}))\leq h(v_{\ell^{\prime}}(u_{0})) for some u0𝙳k(b)u_{0}\in\mathtt{D}_{k}(b). Then by Lemma 6.7, we have

|V(u)||V(u)|\displaystyle\frac{|V_{\ell}(u)|}{|V_{\ell^{\prime}}(u)|} =eτm𝗐,(a)(v(u))|H(v(u))|eτm𝗐,(a)(v(u))|H(v(u))|4eτm𝗐,(a)(v(u))τm𝗐,(a)(v(u))h(v(u))h(v(u))16e2mT𝗐\displaystyle=\frac{e^{\tau_{m}^{\mathsf{w},(a)}(v_{\ell}(u))}\big{|}H(v_{\ell}(u))\big{|}}{e^{\tau_{m}^{\mathsf{w},(a)}(v_{\ell^{\prime}}(u))}\big{|}H(v_{\ell^{\prime}}(u))\big{|}}\leq\frac{4e^{\tau_{m}^{\mathsf{w},(a)}(v_{\ell}(u))-\tau_{m}^{\mathsf{w},(a)}(v_{\ell^{\prime}}(u))}h(v_{\ell}(u))}{h(v_{\ell^{\prime}}(u))}\leq 16e^{2mT_{\mathsf{w}}}

for all u𝙳k(b)u\in\mathtt{D}_{k}(b), which establishes the required bound on relative size. Now applying Lemmas 6.8 and 20 and |H|h|H|\leq h on |Vj(u)+V0(u)||V_{j}(u)+V_{0}(u)| gives χj,0[ξ,H,h](u)1\chi_{j,0}^{[\xi,H,h]}(u)\leq 1 or χj,1[ξ,H,h](u)1\chi_{j,1}^{[\xi,H,h]}(u)\leq 1 for all u𝙳k(b)u\in\mathtt{D}_{k}(b). ∎

Lemma 6.10.

There exists a0>0a_{0}>0 such that for all ξ\xi\in\mathbb{C} with |a|<a0|a|<a_{0} and |b|>b0|b|>b_{0}, if HL(U,)H\in L(U,\mathbb{C}) and h𝒞E|b|(U)h\in\mathcal{C}_{E|b|}(U) satisfy Properties (3)(3)(1a) and (3)(3)(1b) in Theorem 4.8, then there exists J𝒥(b)J\in\mathcal{J}(b) such that

|ξ,𝗐m(H)(u)|𝒩a,J𝗐(h)(u)\displaystyle\left|\mathcal{L}_{\xi,\mathsf{w}}^{m}(H)(u)\right|\leq\mathcal{N}_{a,J}^{\mathsf{w}}(h)(u)

for all uUu\in U.

Appendix A Failure of proper discontinuity of ΓΘ(2)×\Gamma\curvearrowright\mathcal{F}_{\Theta}^{(2)}\times\mathbb{R}

In Theorem A.2, we exhibit numerous instances where the action ΓΘ(2)×\Gamma\curvearrowright\mathcal{F}_{\Theta}^{(2)}\times\mathbb{R} is not properly discontinuous. We remind the reader that the action is of course properly discontinuous whenever GG is of rank one, in which case 𝔞Θ=𝔞\mathfrak{a}_{\Theta}=\mathfrak{a}\cong\mathbb{R}, SΘ=SS_{\Theta}=S is compact, and Γ<G\Gamma<G is convex cocompact.

We will use the following lemma. Recall the Weyl group 𝒲=NK(A)/ZK(A)\mathcal{W}=N_{K}(A)/Z_{K}(A). Note that for any loxodromic element gh(intA+)h1g\in h(\operatorname{int}A^{+})h^{-1} for some hGh\in G, its set of fixed points is h𝒲e+h\mathcal{W}e^{+}\subset\mathcal{F} among which the attracting and repelling ones are h+=he+ΛΓh^{+}=he^{+}\in\Lambda_{\Gamma} and h=he=hw0e+ΛΓh^{-}=he^{-}=hw_{0}e^{+}\in\Lambda_{\Gamma}.

Lemma A.1.

Let GG be a semisimple real algebraic group. Let gGg\in G be a loxodromic element so that gh(intA+)h1g\in h(\operatorname{int}A^{+})h^{-1} for some hGh\in G. Let w𝒲=NK(A)/ZK(A)w\in\mathcal{W}=N_{K}(A)/Z_{K}(A) be an element in the Weyl group. Then

β(hw1)+(h,gkh)=kAdw(λ(g))for all k0.\displaystyle\beta_{(hw^{-1})^{+}}(h,g^{k}h)=k\operatorname{Ad}_{w}(\lambda(g))\qquad\text{for all $k\in\mathbb{Z}_{\geq 0}$}.
Proof.

Let GG and ww be as in the lemma. By GG-equivariance, it suffices to show that β(w1)+(e,av)=Adw(v)\beta_{(w^{-1})^{+}}(e,a_{v})=\operatorname{Ad}_{w}(v) for all vint𝔞+v\in\operatorname{int}\mathfrak{a}^{+}. Indeed, we calculate that

avw1=w1wavw1KaAdw(v)\displaystyle a_{-v}w^{-1}=w^{-1}\cdot wa_{-v}w^{-1}\in Ka_{-\operatorname{Ad}_{w}(v)}

and so by definition, β(w1)+(e,av)=σ(av,(w1)+)=Adw(v)\beta_{(w^{-1})^{+}}(e,a_{v})=-\sigma(a_{-v},(w^{-1})^{+})=\operatorname{Ad}_{w}(v). ∎

Theorem A.2.

There exist infinitely many triples (G,Γ,ψ)(G,\Gamma,\psi) where GG is a connected semisimple real algebraic group of rank at least 22, Γ<G\Gamma<G is a Zariski dense Θ\Theta-Anosov subgroup for some subset ΘΠ\Theta\subset\Pi with #Θ2\#\Theta\geq 2, and ψintΘ\psi\in\operatorname{int}\mathcal{L}_{\Theta}^{*}, such that the action ΓΘ(2)×\Gamma\curvearrowright\mathcal{F}_{\Theta}^{(2)}\times\mathbb{R} defined in Eq. 4 is not properly discontinuous.

Proof.

Let G0G_{0} be any rank one connected simple real algebraic group (of which there are infinitely many), and Γ0<G0\Gamma_{0}<G_{0} be any convex cocompact subgroup. We simply write Id:Γ0G0\operatorname{Id}:\Gamma_{0}\to G_{0} for the representation obtained by restricting the identity map IdG0:G0G0\operatorname{Id}_{G_{0}}:G_{0}\to G_{0}. Take a different convex cocompact discrete faithful representation ρ:Γ0G0\rho:\Gamma_{0}\to G_{0} and consider the induced representation

Id×ρ:Γ0G0×G0.\displaystyle\operatorname{Id}\times\rho:\Gamma_{0}\to G_{0}\times G_{0}.

Take the rank two connected semisimple real algebraic group G=G0×G0G=G_{0}\times G_{0}. Denote by 𝒲0={e,w00}\mathcal{W}_{0}=\{e,w_{00}\} and 0\mathcal{F}_{0} the Weyl group and the Furstenberg boundary associated to G0G_{0}, and by 𝒲=𝒲0×𝒲0/2×/2\mathcal{W}=\mathcal{W}_{0}\times\mathcal{W}_{0}\cong\mathbb{Z}/2\mathbb{Z}\times\mathbb{Z}/2\mathbb{Z} and =0×0\mathcal{F}=\mathcal{F}_{0}\times\mathcal{F}_{0} the same for GG, respectively. Note that w0=(w00,w00)𝒲w_{0}=(w_{00},w_{00})\in\mathcal{W}. Take the discrete subgroup Γ=(Id×ρ)(Γ0)<G\Gamma=(\operatorname{Id}\times\rho)(\Gamma_{0})<G which is the self-joining of Γ0\Gamma_{0} via ρ\rho. It is known that such discrete subgroups are Π\Pi-Anosov either by checking that any orbit of ΓG/K\Gamma\curvearrowright G/K gives a quasiisometric embedding of Γ0\Gamma_{0} or by checking the antipodality condition (in the current setting, (2)0(2)×0(2)\mathcal{F}^{(2)}\cong\mathcal{F}_{0}^{(2)}\times\mathcal{F}_{0}^{(2)}). Hence, Γ\Gamma is also Θ\Theta-Anosov for any nonempty subset ΘΠ\Theta\subset\Pi. We also assume that Γ<G\Gamma<G is Zariski dense (cf. [30, Lemma 2.2] for an equivalent criterion). Denote by ΛΓ00\Lambda_{\Gamma_{0}}\subset\mathcal{F}_{0} the limit set of Γ0\Gamma_{0} and let ζρ:ΛΓ00\zeta_{\rho}:\Lambda_{\Gamma_{0}}\to\mathcal{F}_{0} be the induced (continuous) boundary map.

Now, take any loxodromic element γ=(γ0,ρ(γ0))Γ\gamma=(\gamma_{0},\rho(\gamma_{0}))\in\Gamma for some γ0Γ0\gamma_{0}\in\Gamma_{0}, so that γ=h(intA+)h1\gamma=h(\operatorname{int}A^{+})h^{-1} for some hGh\in G. Write ξ+,ξΛΓ0\xi^{+},\xi^{-}\in\Lambda_{\Gamma_{0}} for the attracting and repelling fixed points of γ0\gamma_{0}, respectively, which are distinct. Then, the set of fixed points of γ\gamma is

h𝒲e+={(ξ+,ζρ(ξ+)),(ξ,ζρ(ξ+)),(ξ+,ζρ(ξ)),(ξ,ζρ(ξ))}\displaystyle h\mathcal{W}e^{+}=\{(\xi^{+},\zeta_{\rho}(\xi^{+})),(\xi^{-},\zeta_{\rho}(\xi^{+})),(\xi^{+},\zeta_{\rho}(\xi^{-})),(\xi^{-},\zeta_{\rho}(\xi^{-}))\}\subset\mathcal{F}

among which the attracting and repelling ones are (ξ+,ζρ(ξ+)),(ξ,ζρ(ξ))ΛΓ(\xi^{+},\zeta_{\rho}(\xi^{+})),(\xi^{-},\zeta_{\rho}(\xi^{-}))\in\Lambda_{\Gamma}. Take any Weyl group element w𝒲{e,w0}w\in\mathcal{W}-\{e,w_{0}\}, say w=(e,w00)w=(e,w_{00}) for the sake of concreteness, and also the fixed point x=hw1e+=hwe+=(ξ+,ζρ(ξ))ΛΓx=hw^{-1}e^{+}=hwe^{+}=(\xi^{+},\zeta_{\rho}(\xi^{-}))\in\mathcal{F}-\Lambda_{\Gamma}, where xΛΓ=(Id×ζρ)(ΛΓ0)x\notin\Lambda_{\Gamma}=(\operatorname{Id}\times\zeta_{\rho})(\Lambda_{\Gamma_{0}}) since ξξ+\xi^{-}\neq\xi^{+}. By Lemma A.1, we have

βx(e,γk)=kAdw(λ(γ))Adw(int𝔞+)for all k.\displaystyle\beta_{x}(e,\gamma^{k})=k\operatorname{Ad}_{w}(\lambda(\gamma))\in\operatorname{Ad}_{w}(\operatorname{int}\mathfrak{a}^{+})\qquad\text{for all $k\in\mathbb{N}$}.

Now, simply choose ψintΘ\psi\in\operatorname{int}\mathcal{L}_{\Theta}^{*} with Adw(λ(γ))kerψ\operatorname{Ad}_{w}(\lambda(\gamma))\in\ker\psi so that

ψ(βγkx(e,γk))=ψ(βx(e,γk))=0for all k.\displaystyle\psi\bigl{(}\beta_{\gamma^{k}x}(e,\gamma^{k})\bigr{)}=\psi(\beta_{x}(e,\gamma^{k}))=0\qquad\text{for all $k\in\mathbb{N}$}. (26)

Take any yy\in\mathcal{F} such that (x,y)(2)(x,y)\in\mathcal{F}^{(2)}, say y=hw1e=(ξ,ζρ(ξ+))y=hw^{-1}e^{-}=(\xi^{-},\zeta_{\rho}(\xi^{+})). Now, by compactness of \mathcal{F}, we could pass to subsequences of {γk}k\{\gamma^{k}\}_{k\in\mathbb{N}} to get {kj}j\{k_{j}\}_{j\in\mathbb{N}}\subset\mathbb{N} such that {γkj(x,y)}j\{\gamma^{k_{j}}(x,y)\}_{j\in\mathbb{N}} converges; however, for our choice of (x,y)(x,y), the original sequence is already just a constant sequence. This imples the easier property that ΓΘ(2)\Gamma\curvearrowright\mathcal{F}_{\Theta}^{(2)} is not properly discontinuous. Using Eq. 26 as well, we have γk(x,y,0)=(γkx,γky,ψ(βγkx(e,γk)))=(x,y,0)\gamma^{k}(x,y,0)=\bigl{(}\gamma^{k}x,\gamma^{k}y,\psi\bigl{(}\beta_{\gamma^{k}x}(e,\gamma^{k})\bigr{)}\bigr{)}=(x,y,0) for all kk\in\mathbb{N}, so we similarly conclude that {γk(x,y,0)}j\{\gamma^{k}(x,y,0)\}_{j\in\mathbb{N}} is a constant sequence and hence converges. Thus, ΓΘ(2)×\Gamma\curvearrowright\mathcal{F}_{\Theta}^{(2)}\times\mathbb{R} is not properly discontinuous. ∎

Appendix B Non-obtuseness of Weyl chambers

We give a quick proof of a fundamental proposition below regarding angles in Weyl chambers. Although a stronger statement holds regarding roots, the proposition below suffices for our purposes. We refer the reader to the concise book of Serre [66, Chapter V] for the basics of root systems used here. In fact the proof given here is along the lines of [66, Chapter V, §7, Proposition 3].

We prefer to work on 𝔞\mathfrak{a}^{*} and so we endow it with the inner product induced by 𝔞\mathfrak{a}. Let (𝔞)+(\mathfrak{a}^{*})^{+} be the closed positive Weyl chamber corresponding to Φ+\Phi^{+}. We write sα:=Id𝔞2,αα,αα𝔞s_{\alpha}:=\operatorname{Id}_{\mathfrak{a}^{*}}-2\frac{\langle\cdot,\alpha\rangle}{\langle\alpha,\alpha\rangle}\alpha\in\mathfrak{a}^{**} for the symmetry associated to the root αΦ\alpha\in\Phi. It follows from the definition of root systems that for all α,βΦ\alpha,\beta\in\Phi, the number n(β,α):=2β,αα,αn(\beta,\alpha):=2\frac{\langle\beta,\alpha\rangle}{\langle\alpha,\alpha\rangle}\in\mathbb{Z} is an integer such that sα(β)=βn(β,α)αΦs_{\alpha}(\beta)=\beta-n(\beta,\alpha)\alpha\in\Phi is also a root.

Proposition B.1.

We have v1,v20\langle v_{1},v_{2}\rangle\geq 0 for all v1,v2𝔞+v_{1},v_{2}\in\mathfrak{a}^{+}. Equivalently, if dim(𝔞)1\dim(\mathfrak{a})\neq 1, then αint((𝔞)+)\alpha\notin\operatorname{int}((\mathfrak{a}^{*})^{+}) for all αΠ\alpha\in\Pi.

Proof.

Suppose that dim(𝔞)1\dim(\mathfrak{a})\neq 1. Then, there exist two distinct simple roots α,βΠ\alpha,\beta\in\Pi and for the sake of contradiction, suppose that β,α>0\langle\beta,\alpha\rangle>0. Then n(β,α)>0n(\beta,\alpha)>0 and n(α,β)>0n(\alpha,\beta)>0. It is a standard fact of root systems that n(β,α)n(α,β){0,1,2,3,4}n(\beta,\alpha)n(\alpha,\beta)\in\{0,1,2,3,4\}. Now, by our initial assumption, the product cannot vanish. Next, by the Cauchy–Schwarz inequality, if the product is 44, then α\alpha and β\beta are collinear in which case β=2α\beta=2\alpha or α=2β\alpha=2\beta by integrality of n(β,α)n(\beta,\alpha) and n(α,β)n(\alpha,\beta), contradicting simplicity of α\alpha and β\beta. Thus, n(β,α)n(α,β){1,2,3}n(\beta,\alpha)n(\alpha,\beta)\in\{1,2,3\}. In any case, integrality again implies that n(β,α)=1n(\beta,\alpha)=1 or n(α,β)=1n(\alpha,\beta)=1. Therefore, either sα(β)=βαΦs_{\alpha}(\beta)=\beta-\alpha\in\Phi or sβ(α)=αβΦs_{\beta}(\alpha)=\alpha-\beta\in\Phi, both of which contradict simplicity of α\alpha and β\beta once again. ∎

Appendix C A minor correction

We take the opportunity to make a minor correction in our prequel [18]. In [18, §4.3], the transition matrix TT is not topologically mixing a priori. However, it is topologically transitive (also called irreducible) as a consequence of the density result in [18, Proposition 6.3] which only uses ergodicity from [18, Proposition 3.7] due to Lee–Oh. One can then arrange the Markov section, by inserting extra rectangles if necessary as in the beginning of the proof of [18, Theorem 6.10], so that coprime primitive periods exist for the Poincaré first return map, implying that TT is aperiodic. Then, towards the end of the proof of [18, Theorem 6.10], “for any nNTn\geq N_{T}” should be replaced with “for some coprime n1,n2NTn_{1},n_{2}\geq N_{T}” and again we conclude θ{0,1}\theta\in\{0,1\}.

With the above correction, the topological mixing result in [18, Theorem 8.1] is valid. Only now, one derives as a consequence that TT is indeed topologically mixing.

References

  • Alb [99] P. Albuquerque. Patterson-Sullivan theory in higher rank symmetric spaces. Geom. Funct. Anal., 9(1):1–28, 1999.
  • Bar [10] Thierry Barbot. Three-dimensional Anosov flag manifolds. Geom. Topol., 14(1):153–191, 2010.
  • Ben [97] Y. Benoist. Propriétés asymptotiques des groupes linéaires. Geom. Funct. Anal., 7(1):1–47, 1997.
  • Ben [96] Yves Benoist. Actions propres sur les espaces homogènes réductifs. Ann. of Math. (2), 144(2):315–347, 1996.
  • Bow [70] Rufus Bowen. Markov partitions for Axiom A{\rm A} diffeomorphisms. Amer. J. Math., 92:725–747, 1970.
  • Bow [73] Rufus Bowen. Symbolic dynamics for hyperbolic flows. Amer. J. Math., 95:429–460, 1973.
  • Bow [08] Rufus 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.
  • BR [75] Rufus Bowen and David Ruelle. The ergodic theory of Axiom A flows. Invent. Math., 29(3):181–202, 1975.
  • BCLS [15] Martin Bridgeman, Richard Canary, François Labourie, and Andres Sambarino. The pressure metric for Anosov representations. Geom. Funct. Anal., 25(4):1089–1179, 2015.
  • BILW [05] Marc Burger, Alessandra Iozzi, François Labourie, and Anna Wienhard. Maximal representations of surface groups: symplectic Anosov structures. Pure Appl. Math. Q., 1(3):543–590, 2005.
  • BIW [24] Marc Burger, Alessandra Iozzi, and Anna. Wienhard. Maximal representations and Anosov structures. 2024. In preparation.
  • BLLO [23] Marc Burger, Or Landesberg, Minju Lee, and Hee Oh. The Hopf-Tsuji-Sullivan dichotomy in higher rank and applications to Anosov subgroups. J. Mod. Dyn., 19:301–330, 2023.
  • Car [23] León Carvajales. Growth of quadratic forms under Anosov subgroups. Int. Math. Res. Not. IMRN, (1):785–854, 2023.
  • Cha [94] Christophe Champetier. Petite simplification dans les groupes hyperboliques. Ann. Fac. Sci. Toulouse Math. (6), 3(2):161–221, 1994.
  • Che [02] N. Chernov. Invariant measures for hyperbolic dynamical systems. In Handbook of dynamical systems, Vol. 1A, pages 321–407. North-Holland, Amsterdam, 2002.
  • CF [23] Michael Chow and Elijah Fromm. Joint equidistribution of maximal flat cylinders and holonomies for anosov homogeneous spaces. 2023. Preprint, arXiv:2305.03590.
  • CS [22] Michael Chow and Pratyush Sarkar. Exponential mixing of frame flows for convex cocompact locally symmetric spaces. 2022. Preprint, arXiv:2211.14737.
  • CS [23] Michael Chow and Pratyush Sarkar. Local mixing of one-parameter diagonal flows on Anosov homogeneous spaces. Int. Math. Res. Not. IMRN, (18):15834–15895, 2023.
  • DGK [17] Jeffrey Danciger, François Guéritaud, and Fanny Kassel. Convex cocompact actions in real projective geometry. Ann. Sci. École Norm. Sup., to appear, 2017. Preprint, arXiv:1704.08711.
  • DMS [24] Benjamin Delarue, Daniel Monclair, and Andrew Sanders. Locally homogeneous Axiom A flows I: projective Anosov subgroups and exponential mixing. 2024. Preprint, arXiv:2403.14257.
  • DKO [24] Subhadip Dey, Dongryul M. Kim, and Hee Oh. Ahlfors regularity of patterson-sullivan measures of anosov groups and applications. 2024. Preprint, arXiv:2401.12398.
  • Dol [98] Dmitry Dolgopyat. On decay of correlations in Anosov flows. Ann. of Math. (2), 147(2):357–390, 1998.
  • ELO [23] Samuel Edwards, Minju Lee, and Hee Oh. Anosov groups: local mixing, counting and equidistribution. Geom. Topol., 27(2):513–573, 2023.
  • GGKW [17] François Guéritaud, Olivier Guichard, Fanny Kassel, and Anna Wienhard. Anosov representations and proper actions. Geom. Topol., 21(1):485–584, 2017.
  • GW [12] Olivier Guichard and Anna Wienhard. Anosov representations: domains of discontinuity and applications. Invent. Math., 190(2):357–438, 2012.
  • Hej [76] Dennis A. Hejhal. The Selberg trace formula for PSL(2,R){\rm PSL}(2,R). Vol. I, volume Vol. 548 of Lecture Notes in Mathematics. Springer-Verlag, Berlin-New York, 1976.
  • KLP [17] Michael Kapovich, Bernhard Leeb, and Joan Porti. Anosov subgroups: dynamical and geometric characterizations. Eur. J. Math., 3(4):808–898, 2017.
  • Kat [95] Tosio Kato. Perturbation theory for linear operators. Classics in Mathematics. Springer-Verlag, Berlin, 1995. Reprint of the 1980 edition.
  • [29] Dongryul M. Kim and Hee Oh. Non-concentration property of patterson-sullivan measures for anosov subgroups. 2023. Preprint, arXiv:2304.14911.
  • [30] Dongryul M. Kim and Hee Oh. Rigidity of kleinian groups via self-joinings: measure theoretic criterion. 2023. Preprint, arXiv:2302.03552.
  • KO [24] Dongryul M. Kim and Hee Oh. Relatively anosov groups: finiteness, measure of maximal entropy, and reparameterization. 2024. Preprint, arXiv:2404.09745.
  • [32] Dongryul M. Kim, Hee Oh, and Yahui Wang. Ergodic dichotomoy for subspace flows in higher rank. 2023. Preprint, arXiv:2310.19976.
  • [33] Dongryul M. Kim, Hee Oh, and Yahui Wang. Properly discontinuous actions, growth indicators and conformal measures for transverse subgroups. 2023. Preprint, arXiv:2306.06846.
  • KM [96] D. Y. Kleinbock and G. A. Margulis. Bounded orbits of nonquasiunipotent flows on homogeneous spaces. In Sinaĭ’s Moscow Seminar on Dynamical Systems, volume 171 of Amer. Math. Soc. Transl. Ser. 2, pages 141–172. Amer. Math. Soc., Providence, RI, 1996.
  • Kob [96] Toshiyuki Kobayashi. Criterion for proper actions on homogeneous spaces of reductive groups. J. Lie Theory, 6(2):147–163, 1996.
  • Lab [06] François Labourie. Anosov flows, surface groups and curves in projective space. Invent. Math., 165(1):51–114, 2006.
  • LY [73] A. Lasota and James A. Yorke. On the existence of invariant measures for piecewise monotonic transformations. Trans. Amer. Math. Soc., 186:481–488 (1974), 1973.
  • Lee [13] John M. Lee. Introduction to smooth manifolds, volume 218 of Graduate Texts in Mathematics. Springer, New York, second edition, 2013.
  • [39] Minju Lee and Hee Oh. Ergodic decompositions of geometric measures on anosov homogeneous spaces. arXiv:2010.11337, 2020.
  • [40] Minju Lee and Hee Oh. Invariant measures for horospherical actions and anosov groups. arXiv:2008.05296, 2020.
  • LPS [23] Jialun Li, Wenyu Pan, and Pratyush Sarkar. Exponential mixing of frame flows for geometrically finite hyperbolic manifolds. 2023. Preprint, arXiv:2302.03798.
  • MW [12] François Maucourant and Barak Weiss. Lattice actions on the plane revisited. Geom. Dedicata, 157:1–21, 2012.
  • OW [16] Hee Oh and Dale Winter. Uniform exponential mixing and resonance free regions for convex cocompact congruence subgroups of SL2(){\rm SL}_{2}(\mathbb{Z}). J. Amer. Math. Soc., 29(4):1069–1115, 2016.
  • PP [90] William Parry and Mark Pollicott. Zeta functions and the periodic orbit structure of hyperbolic dynamics. Astérisque, (187-188):268, 1990.
  • Pat [76] S. J. Patterson. The limit set of a Fuchsian group. Acta Math., 136(3-4):241–273, 1976.
  • PS [16] Vesselin Petkov and Luchezar Stoyanov. Ruelle transfer operators with two complex parameters and applications. Discrete Contin. Dyn. Syst., 36(11):6413–6451, 2016.
  • Pol [85] Mark Pollicott. On the rate of mixing of Axiom A flows. Invent. Math., 81(3):413–426, 1985.
  • Pol [86] Mark Pollicott. Meromorphic extensions of generalised zeta functions. Invent. Math., 85(1):147–164, 1986.
  • Pol [87] Mark Pollicott. Symbolic dynamics for Smale flows. Amer. J. Math., 109(1):183–200, 1987.
  • PS [01] Mark Pollicott and Richard Sharp. Error terms for closed orbits of hyperbolic flows. Ergodic Theory Dynam. Systems, 21(2):545–562, 2001.
  • PS [24] Mark Pollicott and Richard Sharp. Zeta functions in higher Teichmüller theory. Math. Z., 306(3):Paper No. 37, 20, 2024.
  • [52] J.-F. Quint. Mesures de Patterson-Sullivan en rang supérieur. Geom. Funct. Anal., 12(4):776–809, 2002.
  • [53] Jean-François Quint. Divergence exponentielle des sous-groupes discrets en rang supérieur. Comment. Math. Helv., 77(3):563–608, 2002.
  • Rat [73] M. Ratner. Markov partitions for Anosov flows on nn-dimensional manifolds. Israel J. Math., 15:92–114, 1973.
  • Rue [87] D. Ruelle. Resonances for Axiom 𝐀{\bf A} flows. J. Differential Geom., 25(1):99–116, 1987.
  • Rue [86] David Ruelle. Resonances of chaotic dynamical systems. Phys. Rev. Lett., 56(5):405–407, 1986.
  • Rue [89] David Ruelle. The thermodynamic formalism for expanding maps. Comm. Math. Phys., 125(2):239–262, 1989.
  • [58] A. Sambarino. Hyperconvex representations and exponential growth. Ergodic Theory Dynam. Systems, 34(3):986–1010, 2014.
  • [59] Andrés Sambarino. Quantitative properties of convex representations. Comment. Math. Helv., 89(2):443–488, 2014.
  • Sam [15] Andrés Sambarino. The orbital counting problem for hyperconvex representations. Ann. Inst. Fourier (Grenoble), 65(4):1755–1797, 2015.
  • Sam [24] Andrés Sambarino. A report on an ergodic dichotomy. Ergodic Theory Dynam. Systems, 44(1):236–289, 2024.
  • Sar [19] Pratyush Sarkar. Uniform exponential mixing for congruence covers of convex cocompact hyperbolic manifolds. 2019. Preprint, arXiv:1903.00825.
  • Sar [23] Pratyush Sarkar. Congruence counting in schottky and continued fractions semigroups of SO(n,1){\rm SO}(n,1). 2023. Preprint, arXiv:2108.00545.
  • SW [21] Pratyush Sarkar and Dale Winter. Exponential mixing of frame flows for convex cocompact hyperbolic manifolds. Compos. Math., 157(12):2585–2634, 2021.
  • Sel [56] A. Selberg. Harmonic analysis and discontinuous groups in weakly symmetric Riemannian spaces with applications to Dirichlet series. J. Indian Math. Soc. (N.S.), 20:47–87, 1956.
  • Ser [01] Jean-Pierre Serre. Complex semisimple Lie algebras. Springer Monographs in Mathematics. Springer-Verlag, Berlin, 2001. Translated from the French by G. A. Jones, Reprint of the 1987 edition.
  • Sto [11] Luchezar Stoyanov. Spectra of Ruelle transfer operators for axiom A flows. Nonlinearity, 24(4):1089–1120, 2011.
  • Sul [79] Dennis Sullivan. The density at infinity of a discrete group of hyperbolic motions. Inst. Hautes Études Sci. Publ. Math., (50):171–202, 1979.
  • Thi [09] Xavier Thirion. Propriétés de mélange du flot des chambres de Weyl des groupes de ping-pong. Bull. Soc. Math. France, 137(3):387–421, 2009.
  • Win [15] Dale Winter. Mixing of frame flow for rank one locally symmetric spaces and measure classification. Israel J. Math., 210(1):467–507, 2015.