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

On the dimension of limit sets on (3)\mathbb{P}(\mathbb{R}^{3}) via stationary measures: variational principles and applications

Yuxiang Jiao, Jialun Li, Wenyu Pan and Disheng Xu
Abstract

In this article, we establish the variational principle of the affinity exponent of Borel Anosov representations. We also establish such a principle of the Rauzy gasket. In [LPX23], they obtain a dimension formula of the stationary measures on (3){\mathbb{P}}({\mathbb{R}}^{3}). Combined with our result, it allows us to study the Hausdorff dimension of limit sets of Anosov representations in SL3()\mathrm{SL}_{3}(\mathbb{R}) and the Rauzy gasket. It yields the equality between the Hausdorff dimensions and the affinity exponents in both settings. In the appendix, we improve the numerical lower bound of the Hausdorff dimension of Rauzy gasket to 1.51.5.

1 Introduction

This article is the second part of [LPX23]. We provide different examples of computing the Hausdorff dimension of limit sets on a projective space using stationary measures.

First, we consider the limit sets of Anosov representations. For a finitely generated hyperbolic group Γ,\Gamma, let |||\cdot| be the word norm on Γ\Gamma with respect to a fixed finite generating set. A homomorphism ρ:ΓSLn()\rho:\Gamma\to\mathrm{SL}_{n}({\mathbb{R}}) is called a Borel Anosov representation if there exists c>0c>0 such that for every 1pn1,1\leqslant p\leqslant n-1,

σp(ρ(γ))σp+1(ρ(γ))cec|γ|,γΓ,\frac{\sigma_{p}(\rho(\gamma))}{\sigma_{p+1}(\rho(\gamma))}\geqslant ce^{c|\gamma|},\quad\forall\gamma\in\Gamma,

where σp(g)\sigma_{p}(g) denotes the pp-th maximal singular value of gSLn().g\in\mathrm{SL}_{n}({\mathbb{R}}).

The concept of Anosov representation was first introduced by Labourie in [Lab06] to study Hitchin components of the representations of surface groups. They are a generalization of convex cocompact representations in PSO(n,1).\mathrm{PSO}(n,1). Given a Borel Anosov representation ρ,\rho, we consider the action of ρ(Γ)\rho(\Gamma) on the projective space (n){\mathbb{P}}({\mathbb{R}}^{n}). It always admits a limit set, denoted by L(ρ(Γ)),L(\rho(\Gamma)), which is the closure of attracting fixed points of ρ(γ)\rho(\gamma) on (n){\mathbb{P}}({\mathbb{R}}^{n}) for γΓ\gamma\in\Gamma.

To compute the Hausdorff dimension of the limit set of ρ(Γ)\rho(\Gamma), the difficulty lies in the fact that ρ(Γ)\rho(\Gamma) usually acts on (n){\mathbb{P}}({\mathbb{R}}^{n}) non-conformally. For example, consider the action of diagonal matrix diag(2,1,1/2)\mathrm{diag}(2,1,1/2) on (3)\mathbb{P}(\mathbb{R}^{3}). The stretch rates in different directions are different. When computing the Hausdorff dimension of the set, one needs to estimate the minimal number of balls to cover it. Naturally, in the non-conformal setting, an efficient way is to arrange the balls according to the stretch rate. Therefore, it leads us to consider the affinity exponent and we expect the Hausdorff dimension of the limit set equals its affinity exponent. The concept of affinity exponent was proposed by Falconer to study the Hausdorff dimension of self-affine fractals [Fal88]. Later, Pozzetti-Sambarino-Wienhard [PSW22] extended this concept to study Anosov representations.

For 0sn1,0\leqslant s\leqslant n-1, let ψs:SLn()\psi_{s}:\mathrm{SL}_{n}({\mathbb{R}})\to{\mathbb{R}} be given by

ψs(g)1is(logσ1(g)logσi+1(g))+(ss)(logσ1(g)logσs+2(g)),gSLn().\psi_{s}(g)\coloneqq\sum_{1\leqslant i\leqslant\lfloor s\rfloor}(\log\sigma_{1}(g)-\log\sigma_{i+1}(g))+(s-\lfloor s\rfloor)(\log\sigma_{1}(g)-\log\sigma_{\lfloor s\rfloor+2}(g)),\quad g\in\mathrm{SL}_{n}({\mathbb{R}}).

Then the affinity exponent of ρ\rho is given by

sA(ρ)sup{s:γΓexp(ψs(ρ(γ)))=}.s_{\mathrm{A}}(\rho)\coloneqq\sup\left\{s:\sum_{\gamma\in\Gamma}\exp(-\psi_{s}(\rho(\gamma)))=\infty\right\}. (1.1)

It is shown in [PSW22] that the affinity exponent is always an upper bound of the Hausdorff dimension of the limit set. To obtain the equality between two notions of dimensions, it remains to show a reversed inequality. An approach to give a lower bound of the dimension of L(ρ(Γ))L(\rho(\Gamma)) is to consider measures supported on it. A probability Borel measure μ\mu on a metric space is called exact dimensional if there exists α\alpha such that

limr0logμ(B(x,r))logr=α,μa.e.x,\lim_{r\to 0}\frac{\log\mu(B(x,r))}{\log r}=\alpha,\quad\mu\mathrm{-a.e.}~{}x,

and α\alpha is called the exact dimension of μ,\mu, which will be denoted by dimμ\dim\mu. Due to a result by Young [You82], the Hausdorff dimension of a set is bounded below by the exact dimension of measures (if exist) supported on it. Therefore, our approach is to construct satationary measures supported on the limit set whose exact dimensions approximate the affinity exponent.

Let us recall the definition of Lyapunov dimension of stationary measures, which is the expected value of exact dimensions. Let ν\nu be a finitely supported probability measure on SLn()\mathrm{SL}_{n}({\mathbb{R}}) with the Lyapunov spectrum λ(ν)={λ1(ν)λn(ν)}.\lambda(\nu)=\left\{\lambda_{1}(\nu)\geqslant\cdots\geqslant\lambda_{n}(\nu)\right\}. We denote χi(ν)=λ1(ν)λi+1(ν)\chi_{i}(\nu)=\lambda_{1}(\nu)-\lambda_{i+1}(\nu) for 1in1.1\leqslant i\leqslant n-1. Let μ\mu be a ν\nu-stationary measure on (n).{\mathbb{P}}({\mathbb{R}}^{n}). The Furstenberg entropy is given by

hF(μ,ν)=logdgμdμ(ξ)(dgμdμ(ξ))dν(g)dμ(ξ).h_{\mathrm{F}}(\mu,\nu)=\int\log\frac{\mathrm{d}g\mu}{\mathrm{d}\mu}(\xi)\left(\frac{\mathrm{d}g\mu}{\mathrm{d}\mu}(\xi)\right)\mathrm{d}\nu(g)\mathrm{d}\mu(\xi). (1.2)

Assume further that χn1(ν)>0,\chi_{n-1}(\nu)>0, then the Lyapunov dimension of μ\mu is defined to be

dimLYμ=d+hF(μ,ν)(χ1(ν)++χd(ν))χd+1(ν),\dim_{\mathrm{LY}}\mu=d+\frac{h_{\mathrm{F}}(\mu,\nu)-(\chi_{1}(\nu)+\cdots+\chi_{d}(\nu))}{\chi_{d+1}(\nu)}, (1.3)

where dd is the maximal integer such that χ1(ν)++χd(ν)hF(μ,ν).\chi_{1}(\nu)+\cdots+\chi_{d}(\nu)\leqslant h_{\mathrm{F}}(\mu,\nu).

To achieve the affinity exponent in our setting, we will approximate it by Lyapunov dimensions of stationary measures. We call it a variational principle of critical exponent. Such variational principle has been considered in several different contexts to obtain a lower bound of the Hausdorff dimension of dynamically invariant sets: Morris-Shmerkin [MS19] and Morris-Sert [MS23] on self-affine IFSs on n{\mathbb{R}}^{n} and He-Jiao-Xu [HJX23] on Diff(SS1)\mathrm{Diff}(\SS^{1}).

In conclusion, our approach can be roughly summarized by the following inequalities

sA(ρ)(1)dimL(ρ(Γ))(2)supμdimμ=(3)supμdimLYμ(4)sA(ρ);s_{\mathrm{A}}(\rho)\stackrel{{\scriptstyle\text{(1)}}}{{\geqslant}}\dim L(\rho(\Gamma))\stackrel{{\scriptstyle\text{(2)}}}{{\geqslant}}\sup_{\mu}\dim\mu\stackrel{{\scriptstyle\text{(3)}}}{{=}}\sup_{\mu}\dim_{\mathrm{LY}}\mu\stackrel{{\scriptstyle\text{(4)}}}{{\geqslant}}s_{\mathrm{A}}(\rho);

here (1) is established in [PSW22], (2) is established in [You82], [Rap21] and [LL23b], (3) is established in [HS17] and [LPX23] for special cases. We obtain the following variational principle, which establishes (4). We say ρ:ΓSLn()\rho:\Gamma\to\mathrm{SL}_{n}({\mathbb{R}}) is Zariski dense if ρ(Γ)\rho(\Gamma) is Zariski dense in SLn().\mathrm{SL}_{n}({\mathbb{R}}).

Theorem 1.1.

Let ρ:ΓSLn()\rho:\Gamma\to\mathrm{SL}_{n}({\mathbb{R}}) be a Zariski dense Borel Anosov representation. For every ϵ>0,\epsilon>0, there exists a finitely supported probability measure ν\nu on ρ(Γ)\rho(\Gamma) whose support generates a Zariski dense subgroup such that the unique ν\nu-stationary measure μ\mu satisfying

dimLYμsA(ρ)ϵ.\dim_{\mathrm{LY}}\mu\geqslant s_{\mathrm{A}}(\rho)-\epsilon.

To obtain the approximations, we need to find a probability measure ν\nu on ρ(Γ)\rho(\Gamma) with large entropy and controlled Lyapunov exponents. In our setting, the Furstenberg entropy equals the random walk entropy (2.1), which characterizes the freeness of the semigroup generated by suppν.\operatorname{\mathrm{supp}}\nu. We aim to find a finite subset of ρ(Γ)\rho(\Gamma) which freely generates a free semigroup, and consider the uniform random walk on this subset.

At the core of the proof, we use a geometric group theoretic argument to construct free semigroups, which is different from the one in the IFS setting. Thanks to the hyperbolicity of the group, we can always approximate the group by free semigroups, in the sense of the growth rate of groups. However, the approximating semigroups should satisfy some additional conditions coming from the dynamics. Our key argument, as presented in Sections 3 and 4, establishes such a desired construction.

A direct application of Theorem 1.1 is computing the Hausdorff dimension of limit sets of Anosov representations in SL3().\mathrm{SL}_{3}({\mathbb{R}}). Let us recall the dimension formula of stationary measures established in the first part of our paper [LPX23].

Theorem 1.2 ([LPX23, Theorem 1.10]).

Let ν\nu be a Zariski dense, finitely supported probability measure on SL3()\mathrm{SL}_{3}(\mathbb{R}) that satisfies the exponential separation condition, and μ\mu be its Furstenberg measure on (3)\mathbb{P}(\mathbb{R}^{3}). Then we have dimμ=dimLYμ.\dim\mu=\dim_{\mathrm{LY}}\mu.

As a consequence, we can derive the Hausdorff dimension of limit sets, as shown in [LPX23].

Theorem 1.3 ([LPX23, Theorem 1.3]).

Let Γ\Gamma be a hyperbolic group and ρ:ΓSL3()\rho:\Gamma\rightarrow\mathrm{SL}_{3}(\mathbb{R}) be an irreducible Anosov representation, then the Hausdorff dimension of the limit set L(ρ(Γ))L(\rho(\Gamma)) in (3)\mathbb{P}(\mathbb{R}^{3}) equals the affinity exponent sA(ρ)s_{\mathrm{A}}(\rho).

In a similar vein of Theorem 1.1, we also obtain a variational principle on the flag variety, Proposition 4.4. An application is to obtain an estimation of the Hausdorff dimension of limit sets of Borel Anosov representations on the flag variety by Ledrappier-Lessa [LL23a] .

Rauzy gasket

Another example we consider is the Rauzy gasket which is a fractal set on (3){\mathbb{P}}({\mathbb{R}}^{3}) formed by projective actions of SL3()\mathrm{SL}_{3}({\mathbb{R}}). We also establish the identity between its Hausdorff dimension and its affinity exponent: we show the affinity exponent is an upper bound of its Hausdorff dimension and a variational principle of the affinity exponent.

Let Δ\Delta be the projectivization of {(x,y,z):x,y,z0}\{(x,y,z):x,y,z\geqslant 0\} in (3).{\mathbb{P}}({\mathbb{R}}^{3}). Let Γ\Gamma_{\mathscr{R}} be the semigroup generated by

A1=(111010001),A2=(100111001),A3=(100010111),A_{1}=\begin{pmatrix}1&1&1\\ 0&1&0\\ 0&0&1\end{pmatrix},\ A_{2}=\begin{pmatrix}1&0&0\\ 1&1&1\\ 0&0&1\end{pmatrix},\ A_{3}=\begin{pmatrix}1&0&0\\ 0&1&0\\ 1&1&1\end{pmatrix},

and we call it the Rauzy semigroup. Then as ΓSL3()\Gamma_{\mathscr{R}}\subset\mathrm{SL}_{3}(\mathbb{R}), the semigroup Γ\Gamma_{\mathscr{R}} acts on (3)\mathbb{P}(\mathbb{R}^{3}). Due to the choice of Δ\Delta, the semigroup Γ\Gamma_{\mathscr{R}} preserves Δ\Delta. The Rauzy gasket XX is the unique attractor of the Rauzy semigroup, which can be defined formally as

nij{1,2,3}(Ai1AinΔ).\bigcap_{n\rightarrow\infty}\bigcup_{i_{j}\in\{1,2,3\}}(A_{i_{1}}\cdots A_{i_{n}}\Delta).

The Rauzy gasket, depicted in Figure 1, was first introduced in 1991 by Arnoux and Rauzy [AR91] in the context of interval exchange transformations. They conjectured that the gasket has Lebesgue measure 0. Levitt [Lev93] rediscovered the gasket in 1993 and confirmed the Arnoux-Rauzy conjecture, the proof of which is essentially due to Yoccoz. Later, in the study of Novikov’s problem, Dynnikov and De Leo [DD09] provided a numerical estimate of the Hausdorff dimension of the Rauzy gasket. They suggested lower and upper bounds are 1.71.7 and 1.81.8, respectively. Meanwhile, Arnoux asked whether the Hausdorff dimension is less than or equal to 22 [AS13]. Avila, Hubert, and Skripchenko [AHS16] provided a positive answer to this question, and Gutiérrez-Romo and Matheus in [GRM20] proved the lower bound is greater than 1.191.19. Recently, Pollicott and Sewell [PS21] used a renewal theoretical argument to show that the Hausdorff dimension of the Rauzy gasket is less than 1.74071.7407. See also [Fou20] for a weaker upper bound.

Refer to caption
Figure 1: Figure in [AS13]

Let ρ\rho be the natural embedding of Rauzy semigroup Γ\Gamma_{\mathscr{R}} into SL3().\mathrm{SL}_{3}({\mathbb{R}}). The affinity exponent sA(Γ)s_{\mathrm{A}}(\Gamma_{\mathscr{R}}) of the Rauzy gasket is defined as the critical exponent of (1.1), which also works for the Rauzy semigroup. The following theorem confirms a folk-lore conjecture of the Hausdorff dimension of the Rauzy gasket.

Theorem 1.4.

The Hausdorff dimension of the Rauzy gasket XX is equal to its affinity exponent sA(Γ).s_{\mathrm{A}}(\Gamma_{\mathscr{R}}).

The result and the proof of Theorem have an interesting outgrowth: we improve the numerical lower bound obtained in [GRM20] to 3/23/2, and the argument is versatile and allows us to deal with a (semi)group which contains a large reducible subsemigroup.

Corollary 1.5.

We have dimX3/2\dim X\geqslant 3/2.

Unlike Anosov representations, the Rauzy semigroup is non-uniformly hyperbolic. The estimate of the upper bound of the Hausdorff dimension requires the study of the points where the action lacks hyperbolicity. To obtain a lower bound of the Hausdorff dimension, we also establish a variational principle for the Rauzy semigroup. We build on the variational principle for IFS setting [MS23]. Like Anosov representations, we aim to find free semigroups. In this process, the difficulty lies in that non-uniform hyperbolicity makes us lose control of word lengths. We make use of the prefix argument to overcome the issue. This prefix argument also occurs in [HJX23].

Remark 1.6.

Recently Natalia Jurga also obtained 1.4 independently.

Organization.

In Section 2, we discuss different notions of entropy. We establish a geometric group theoretic lemma in Section 3 and the variational principle for Anosov representations in Section 4. Section 5 is devoted to the study of the Hausdorff dimension of the Rauzy gasket.

Acknowledgement.

We would like to thank Wenyuan Yang for carefully explaining the basic ideas and arguments in [Yan19], which is useful for Section 3. We would like to thank Cagri Sert, François Ledrappier, and Pablo Lessa for helpful discussions. Part of the work was done in the conference “Beyond uniform hyperbolicity” at the Banach Center in Będlewo, Poland, in 2023. We thank the organizers and the hospitality of the center.

2 Preliminaries

2.1 Actions and random walks of SLn()\mathrm{SL}_{n}({\mathbb{R}})

Consider the nn-dimensional Euclidean space n\mathbb{R}^{n} and denote by \|\cdot\| the Euclidean norm. By abuse of the notation, we denote by \|\cdot\| the norm on 2n\wedge^{2}\mathbb{R}^{n} induced by the one in n\mathbb{R}^{n}. Let (n){\mathbb{P}}({\mathbb{R}}^{n}) be the projective space. The metric on (n){\mathbb{P}}({\mathbb{R}}^{n}) is given by

d(v,w):=vwvwfor anyv,w(n),d(\mathbb{R}v,\mathbb{R}w):=\frac{\|v\wedge w\|}{\|v\|\|w\|}\,\,\,\text{for any}\,\,\,\mathbb{R}v,\mathbb{R}w\in\mathbb{P}(\mathbb{R}^{n}),

which is bi-Lipschitz equivalent to the SO(n)\mathrm{SO}(n)-invariant metric on (n).{\mathbb{P}}({\mathbb{R}}^{n}).

Let 𝔞={λ=diag(λ1,,λn):λi,iλi=0}\mathfrak{a}=\left\{\lambda=\mathrm{diag}(\lambda_{1},\cdots,\lambda_{n}):\lambda_{i}\in{\mathbb{R}},\sum_{i}\lambda_{i}=0\right\} be a Cartan algebra of 𝔰𝔩n()\mathfrak{sl}_{n}({\mathbb{R}}) and 𝔞+={λ𝔞:λ1λn}\mathfrak{a}^{+}=\left\{\lambda\in\mathfrak{a}:\lambda_{1}\geqslant\cdots\geqslant\lambda_{n}\right\} be a positive Weyl chamber. Set A+=exp𝔞+A^{+}=\exp\mathfrak{a}^{+} and K=SOn().K=\mathrm{SO}_{n}({\mathbb{R}}).

For every gSLn(),g\in\mathrm{SL}_{n}({\mathbb{R}}), it admits the Cartan decomposition g=k~gagkgKA+K.g=\widetilde{k}_{g}a_{g}k_{g}\in KA^{+}K. Here, ag=diag(σ1(g),,σn(g))a_{g}=\mathrm{diag}(\sigma_{1}(g),\cdots,\sigma_{n}(g)) where σ1(g)σn(g)\sigma_{1}(g)\geqslant\cdots\geqslant\sigma_{n}(g) are singular values of gg as the notation before. The Cartan projection of gg is defined to be κ(g)diag(logσ1(g),logσn(g))𝔞+.\kappa(g)\coloneqq\mathrm{diag}(\log\sigma_{1}(g),\cdots\log\sigma_{n}(g))\in\mathfrak{a}^{+}. Let e1,,ene_{1},\cdots,e_{n} be the standard orthonormal basis of n\mathbb{R}^{n} and Ei=ei(n)E_{i}={\mathbb{R}}e_{i}\in{\mathbb{P}}({\mathbb{R}}^{n}) be the corresponding point in the projective space. We consider the following notions as in our first paper [LPX23]:

  • Vg+k~gE1(n)V_{g}^{+}\coloneqq\widetilde{k}_{g}E_{1}\in\mathbb{P}(\mathbb{R}^{n}), which is an attracting point of gg;

  • Hgkg1(E2En)(d)H_{g}^{-}\coloneqq k_{g}^{-1}(E_{2}\oplus\cdots\oplus E_{n})\subset\mathbb{P}(\mathbb{R}^{d}), which is a repelling hyperplane of gg

    (if σ1>σ2\sigma_{1}>\sigma_{2}, then Vg+V_{g}^{+} and HgH_{g}^{-} are uniquely defined);

  • b(g,ϵ){x(n):d(x,Hg)>ϵ}b(g^{-},\epsilon)\coloneqq\{x\in\mathbb{P}(\mathbb{R}^{n}):d(x,H_{g}^{-})>\epsilon\} for any ϵ>0\epsilon>0;

  • B(g+,ϵ){x(n):d(x,Vg+)ϵ}B(g^{+},\epsilon)\coloneqq\{x\in\mathbb{P}(\mathbb{R}^{n}):d(x,V_{g}^{+})\leqslant\epsilon\} for any ϵ>0\epsilon>0.

We have the following useful lemma [BQ16, Lemma 14.2] for later use.

Lemma 2.1.

For any gSLn()g\in\mathrm{SL}_{n}(\mathbb{R}) and V=v(n)V=\mathbb{R}v\in\mathbb{P}(\mathbb{R}^{n}), we have

d(V,Hg)gvgvd(V,Hg)+σ2(g)σ1(g),d(gV,Vg+)d(V,Hg)σ2(g)σ1(g).\displaystyle d(V,H_{g}^{-})\leqslant\frac{\|gv\|}{\|g\|\|v\|}\leqslant d(V,H_{g}^{-})+\frac{\sigma_{2}(g)}{\sigma_{1}(g)},\quad d(gV,V_{g}^{+})d(V,H_{g}^{-})\leqslant\frac{\sigma_{2}(g)}{\sigma_{1}(g)}.

Let ν\nu be a finitely supported probability measure on SLn().\mathrm{SL}_{n}({\mathbb{R}}). It induces random walks on the projective space and the flag variety. Recall the flag variety of n{\mathbb{R}}^{n} as

(n){ξ=(ξ1ξ2ξiξn1):ξi is a linear subspace of n of dimension i}\mathcal{F}({\mathbb{R}}^{n})\coloneqq\left\{\,\xi=(\xi^{1}\subset\xi^{2}\subset\cdots\subset\xi^{i}\subset\cdots\subset\xi^{n-1}):\xi^{i}\text{ is a linear subspace of }{\mathbb{R}}^{n}\text{ of dimension }i\,\right\}

and that SLn()\mathrm{SL}_{n}(\mathbb{R}) acts on (n)\mathcal{F}({\mathbb{R}}^{n}) canonically.

We denote GνG_{\nu} to be the group generated by suppν.\operatorname{\mathrm{supp}}\nu. If we further assume that GνG_{\nu} is Zariski dense in SLn().\mathrm{SL}_{n}({\mathbb{R}}). Then the Lyapunov spectrum λ(ν)={λ1(ν)λn(ν)}\lambda(\nu)=\left\{\lambda_{1}(\nu)\geqslant\cdots\geqslant\lambda_{n}(\nu)\right\} is simple. Moreover, the random walks induced by ν\nu on the projective space (n){\mathbb{P}}({\mathbb{R}}^{n}) and the flag variety (n)\mathcal{F}({\mathbb{R}}^{n}) both have a unique stationary measure ([Fur63, GR85, GM89], see also [BQ16, Chapter 10]).

2.2 Different notions of entropies

The Furstenberg entropy is mysterious and might be difficult to compute. We recall another notion of the entropy associated to the random walk. The random walk entropy of ν\nu is

hRW(ν)=limk1kH(νk).h_{\mathrm{RW}}(\nu)=\lim_{k\to\infty}\frac{1}{k}H(\nu^{*k}). (2.1)
Remark 2.2.

In [BHR19] and [Rap21], the notion of the entropy they used is the following: for a discrete measure ν=piδgi\nu=\sum p_{i}\delta_{g_{i}}, H(ν)H(p)=pilogpiH(\nu)\coloneqq H(p)=-\sum p_{i}\log p_{i}. This notion works well in many settings of IFSs. For general semigroup actions, the random walk entropy hRW(ν)h_{\mathrm{RW}}(\nu) as in [HS17] is more precise. In particular, if suppν\operatorname{\mathrm{supp}}\nu freely generates a free semigroup then H(ν)=hRW(ν)H(\nu)=h_{\mathrm{RW}}(\nu). This kind of entropy was first studied by Avez in [Ave72] to study the structure of the group action on the boundary.

To obtain a more calculable dimension formula, it is expected to show that the Furstenberg entropy in 1.2 is equal to the random walk entropy. In the following proposition we will see that for some concrete examples, we do have this equality. We say a representation ρ:ΓSLn()\rho:\Gamma\to\mathrm{SL}_{n}({\mathbb{R}}) is Zariski dense if ρ(Γ)\rho(\Gamma) is Zariski dense.

Proposition 2.3.

Let Γ\Gamma be a hyperbolic group and ρ:ΓSLn()\rho:\Gamma\rightarrow\mathrm{SL}_{n}(\mathbb{R}) be a Zariski dense Borel Anosov representation. Let ν\nu be a finitely supported probability measure on ρ(Γ)\rho(\Gamma) such that GνG_{\nu} is Zariski dense. Then the unique ν\nu-stationary measure μ\mu on (n)\mathbb{P}(\mathbb{R}^{n}) satisfies

hF(μ,ν)=hRW(ν).h_{\mathrm{F}}(\mu,\nu)=h_{\mathrm{RW}}(\nu).
Remark 2.4.

The Zariski density assumptions on ρ\rho and GνG_{\nu} are both nonnecessary. It is shown in [LL23a] that for every ν\nu supported on ρ(Γ)\rho(\Gamma) with non-elementary GνG_{\nu}, the equality of entropies holds.

To show Propostion 2.3, we may consider the random walk on the flag variety.

Proposition 2.5.

Let ν\nu be a probability measure on SLn()\mathrm{SL}_{n}(\mathbb{R}) such that GνG_{\nu} is a Zariski dense discrete subgroup. Let μ=μ(ν)\mu_{\mathcal{F}}=\mu_{\mathcal{F}}(\nu) and μ=μ(ν)\mu=\mu(\nu) be the unique ν\nu-stationary measure on (n)\mathcal{F}({\mathbb{R}}^{n}) and (n)\mathbb{P}(\mathbb{R}^{n}) respectively. Then we have

hRW(ν)=hF(μ,ν)hF(μ,ν).h_{\mathrm{RW}}(\nu)=h_{\mathrm{F}}(\mu_{\mathcal{F}},\nu)\geqslant h_{\mathrm{F}}(\mu,\nu).
Proof.

Proposition 2.5 follows from [Fur02, Theorem 2.31] (originally in [KV83, Theorem 3.2], [Led85, Section 3.2]). In order to apply their result, we need to use [Fur02, Theorem 2.21] (originally in [KV83], [Led85]) to obtain that (suppμ,μ)(\operatorname{\mathrm{supp}}\mu_{\mathcal{F}},\mu_{\mathcal{F}}) is the Poisson boundary of (Gν,ν)(G_{\nu},\nu). Then the Furstenberg entropy of the Poisson boundary is exactly the random walk entropy for discrete GνG_{\nu} due to [Fur02, Theorem 2.31]. ∎

Proof of 2.3.

By Proposition 2.5 and that the image of an Anosov representation is discrete, it remains to prove that hF(μ,ν)=hF(μ,ν)h_{\mathrm{F}}(\mu_{\mathcal{F}},\nu)=h_{\mathrm{F}}(\mu,\nu).

Consider the canonical projection

π:(n)(n),ξ=(ξ1ξ2ξn1)ξ1.\pi:\mathcal{F}({\mathbb{R}}^{n})\to\mathbb{P}(\mathbb{R}^{n}),~{}\xi=(\xi^{1}\subset\xi^{2}\subset\cdots\subset\xi^{n-1})\mapsto\xi^{1}.

Due to uniqueness of the Furstenberg measure on (n)\mathbb{P}(\mathbb{R}^{n}), we know that π(μ)=μ\pi_{*}(\mu_{\mathcal{F}})=\mu. By classical Rokhlin’s disintegration theorem, we can define a desintegration {μξ}\{\mu^{\xi}\} of the measure μ\mu_{\mathcal{F}} over μ\mu, where μξ\mu^{\xi} is a well-defined probabilty measure on π1(ξ)\pi^{-1}(\xi) for μ\mu a.e. ξ\xi.

Let L(ρ(Γ))L_{\mathcal{F}}(\rho(\Gamma)) and L(ρ(Γ))L(\rho(\Gamma)) be the limit sets on (n)\mathcal{F}({\mathbb{R}}^{n}) and (n)\mathbb{P}(\mathbb{R}^{n}) (the closure of attracting fixed points of proximal elements) respectively. [BQ16, Lemma 9.5] tells us that L(ρ(Γ))L_{\mathcal{F}}(\rho(\Gamma)) (resp. L(ρ(Γ))L(\rho(\Gamma))) is the unique ρ(Γ)\rho(\Gamma)-minimal closed invariant subset on (n)\mathcal{F}({\mathbb{R}}^{n}) (resp. (n)\mathbb{P}(\mathbb{R}^{n})). Hence μ\mu_{\mathcal{F}} (resp. μ\mu) is supported on the limit set L(ρ(Γ))L_{\mathcal{F}}(\rho(\Gamma)) (resp. L(ρ(Γ))L(\rho(\Gamma))). Before continuing, we need a lemma about the structure of the limit sets.

Lemma 2.6.

Let ρ:ΓSLn()\rho:\Gamma\rightarrow\mathrm{SL}_{n}(\mathbb{R}) be a Borel Anosov representation. Then the canonical projection π:L(ρ(Γ))L(ρ(Γ))\pi:L_{\mathcal{F}}(\rho(\Gamma))\rightarrow L(\rho(\Gamma)) has a trivial fibre.

Proof.

Because ρ:ΓSLn()\rho:\Gamma\rightarrow\mathrm{SL}_{n}({\mathbb{R}}) is Borel Anosov, we can apply [Can, Theorem 31.1]. Then there exists a continuous ρ\rho equivariant map (ξ1,ξ)(\xi^{1},\xi^{\prime}) from Γ\partial\Gamma to (n)×Grass(n1,n)\mathbb{P}(\mathbb{R}^{n})\times\mathrm{Grass}(n-1,\mathbb{R}^{n}), such that ξ1\xi^{1} satisfies the Cartan property in [Can, Chapter 30] and ξ1(x)ξ(x),\xi^{1}(x)\subset\xi^{\prime}(x),

ξ1(x)ξ(y)=n,xy.\xi^{1}(x)\oplus\xi^{\prime}(y)=\mathbb{R}^{n},\quad\forall x\neq y.

It follows that ξ1\xi^{1} is injective from Γ\partial\Gamma to (n)\mathbb{P}(\mathbb{R}^{n}). Moreover, the image of ξ1\xi^{1} is exactly L(ρ(Γ)).L(\rho(\Gamma)).

From Borel Anosov property, ξ1\xi^{1} can be extended to a limit map ξ:Γ\xi:\partial\Gamma\rightarrow\mathcal{F} given by ξ(x)=(ξ1(x)ξ2(x)ξn1(x))\xi(x)=(\xi^{1}(x)\subset\xi^{2}(x)\subset\cdots\subset\xi^{n-1}(x)). 111The existence of the limit map ξk(x)\xi^{k}(x) is from PkP_{k}-Anosov and the consistence condition ξk(x)ξk+1(x)\xi^{k}(x)\subset\xi^{k+1}(x) can be deduced from the Cartan property of the limit map (see for example Section 30 and 31 in [Can]), that is ξk(x)=limnUk(γn)\xi^{k}(x)=\lim_{n\rightarrow\infty}U_{k}(\gamma_{n}). Here γn\gamma_{n} is a sequence converges to the boundary point xx and Uk(γn)=kγn(E1Ek)U_{k}(\gamma_{n})=k_{\gamma_{n}}(E_{1}\oplus\cdots\oplus E_{k}) from the Cartan decomposition γn=kγnaγnkγn\gamma_{n}=k_{\gamma_{n}}a_{\gamma_{n}}k^{\prime}_{\gamma_{n}}. The image of ξ\xi is L(ρ(Γ))L_{\mathcal{F}}(\rho(\Gamma))\subset\mathcal{F}. The map ξ1\xi^{1} being injective implies that the natural projection of L(ρ(Γ))L_{\mathcal{F}}(\rho(\Gamma)) to L(ρ(Γ))L(\rho(\Gamma)) has trivial fiber. ∎

By Lemma 2.6, we know that the fiber is trivial for all ξL(ρ(Γ))\xi\in L(\rho(\Gamma)). Hence we have the relation of relative measure-preserving of ((n),μ)((n),μ)(\mathcal{F}({\mathbb{R}}^{n}),\mu_{\mathcal{F}})\rightarrow(\mathbb{P}(\mathbb{R}^{n}),\mu), that is for ν\nu a.e. gg and μ\mu a.e. ξ\xi

gμξ=gδπ1ξ=δπ1(gξ)=μgξ.g\mu^{\xi}=g\delta_{\pi^{-1}\xi}=\delta_{\pi^{-1}(g\xi)}=\mu^{g\xi}.

By [Fur02, Proposition 2.25] (see also [KV83]), we obtain

hF(μ,ν)=hF(μ,ν)h_{\mathrm{F}}(\mu_{\mathcal{F}},\nu)=h_{\mathrm{F}}(\mu,\nu)

Then by Proposition 2.5, the proof is complete. ∎

3 Free sub-semigroups in hyperbolic groups

This section is devoted to establish the geometric group theoretical preparation for the later proof of variational principle. In this section, Γ\Gamma is a finitely generated group with a fixed symmetric generating set 𝒮\mathcal{S}. For every gΓ,g\in\Gamma, let |g||g| denote the word length of gg with respect to 𝒮.\mathcal{S}. For a positive integer L,L, let A(L){gΓ:|g|=L}A(L)\coloneqq\left\{g\in\Gamma:|g|=L\right\} refer to an annulus. The following is the main proposition of this section.

Proposition 3.1.

Let Γ\Gamma be a non-elementary torsion-free hyperbolic group and ρ:ΓSLn()\rho:\Gamma\to\mathrm{SL}_{n}({\mathbb{R}}) be a faithful representation. Let 𝐆{\mathbf{G}} be the Zariski closure of ρ(Γ)\rho(\Gamma) which is assumed to be Zariski connected. Then there exists a finite subset FΓF\subset\Gamma with #F3,\#F\geqslant 3, constants C1,C2,L0>0C_{1},C_{2},L_{0}>0 and m+m\in{\mathbb{Z}}_{+} such that the following holds.

For every subset SA(L)S\subset A(L) for some LL0L\geqslant L_{0} there exists a subset SSS^{\prime}\subset S with #SC11#S\#S^{\prime}\geqslant C_{1}^{-1}\#S and FFF^{\prime}\subset F with #F=#F2\#F^{\prime}=\#F-2 satisfying

  1. (1)

    {ρ(f)m:fF}\left\{\rho(f)^{m}:f\in F^{\prime}\right\} generates a semigroup whose Zarski closure is 𝐆.{\mathbf{G}}.

  2. (2)

    S~{sfς:sS,fF,ς=m,2m}Γ\widetilde{S}\coloneqq\left\{sf^{\varsigma}:s\in S^{\prime},f\in F^{\prime},\varsigma=m,2m\right\}\subset\Gamma freely generates a free semigroup.

  3. (3)

    For any sequence of elements s~1,,s~kS~,\widetilde{s}_{1},\cdots,\widetilde{s}_{k}\in\widetilde{S}, we have

    |s~1s~k|i=1k|s~i|kC2.|\widetilde{s}_{1}\cdots\widetilde{s}_{k}|\geqslant\sum_{i=1}^{k}|\widetilde{s}_{i}|-kC_{2}. (3.1)

3.1 Preliminaries on geometric group theory

Recall that Γ\Gamma is a finitely generated group and 𝒮\mathcal{S} is a fixed symmetric generating set. Let X=Cay(Γ,𝒮)X=\operatorname{\mathrm{Cay}}(\Gamma,\mathcal{S}) be the Cayley graph of Γ\Gamma with respect to 𝒮.\mathcal{S}. Endow XX with the graph metric d𝒮,d_{\mathcal{S}}, which makes XX a proper geodesic space. We abbreviate d𝒮d_{\mathcal{S}} to dd in this section. Then Γ\Gamma has a natural left action on (X,d)(X,d) by isometries. Letting oo be the point in XX corresponding to the identity in Γ,\Gamma, we fix oo as the base point. Then the word length |g||g| is equal to d(o,go).d(o,go).

For every subset YXY\subset X and r>0,r>0, we use 𝒩r(Y)\mathcal{N}_{r}(Y) to denote the rr-neighborhood of Y.Y. For two subsets Y1,Y2X,Y_{1},Y_{2}\subset X, we use dH(Y1,Y2)d_{\mathrm{H}}(Y_{1},Y_{2}) to denote the Hausdorff distance between Y1,Y2Y_{1},Y_{2} with respect to d,d, given by

dH(Y1,Y2)inf{0<r:Y1𝒩r(Y2) and Y2𝒩r(Y1)}.d_{\mathrm{H}}(Y_{1},Y_{2})\coloneqq\inf\left\{0<r\leqslant\infty:Y_{1}\subset\mathcal{N}_{r}(Y_{2})\text{ and }Y_{2}\subset\mathcal{N}_{r}(Y_{1})\right\}.

For a subset YXY\subset X and a point xX,x\in X, we denote πY(x)\pi_{Y}(x) to be the projection of xx on Y,Y, that is πY(x){yY:d(x,y)=d(x,Y)}.\pi_{Y}(x)\coloneqq\left\{y\in Y:d(x,y)=d(x,Y)\right\}. For another subset ZX,Z\subset X, the projection of ZZ on YY is πY(Z)zZπY(z).\pi_{Y}(Z)\coloneqq\cup_{z\in Z}\pi_{Y}(z).

For a rectifiable path αX,\alpha\subset X, we denote α\alpha_{-} and α+\alpha_{+} to be the initial and terminal points of α,\alpha, respectively. The length of α\alpha is denoted by (α).\ell(\alpha). For every pair x,yX,x,y\in X, we denote [x,y][x,y] to be a choice of a geodesic between xx and y.y.

A path α\alpha is called a cc-quasi-geodesic for c1c\geqslant 1 if

(β)cd(β,β+)+c\ell(\beta)\leqslant c\cdot d(\beta_{-},\beta_{+})+c

for every rectifiable subpath βα.\beta\subset\alpha. Morse lemma states that every cc-quasi-geodesic α\alpha is contained in the cc^{\prime}-neighborhood of [α,α+][\alpha_{-},\alpha_{+}] in a δ\delta-hyperbolic space, where cc^{\prime} only depends on cc and δ.\delta. A subset YXY\subset X is called cc-quasi-convex for c0c\geqslant 0 if for every x1,x2Y,x_{1},x_{2}\in Y, we have [x1,x2]𝒩c(Y).[x_{1},x_{2}]\subset\mathcal{N}_{c}(Y). A quasi-geodesic can also be interpreted as a quasi-convex subset.

Recall that Γ\Gamma is a (δ\delta-)hyperbolic group if and only if XX is a (δ\delta-)hyperbolic space. That is, for every geodesic triangle in X,X, every edge is contained in the δ\delta-neighborhood of the other two edges. In the following of this section, we always assume that Γ\Gamma is a finitely generated δ\delta-hyperbolic group. Every infinite order element gΓg\in\Gamma is loxodromic in the following sence

limn+1n|gn|=limn+1nd(o,gno)>0.\lim_{n\to+\infty}\frac{1}{n}|g^{n}|=\lim_{n\to+\infty}\frac{1}{n}d(o,g^{n}o)>0.

In fact, for every loxodromic element g,g, the map ngnon\mapsto g^{n}o is a quasi-isometric embedding from {\mathbb{Z}} to (X,d).(X,d). That is, there exists c1,c2,c3,c4>0c_{1},c_{2},c_{3},c_{4}>0 such that for every m,n,m,n\in{\mathbb{Z}},

c1|mn|c2d(gmo,gno)c3|mn|+c4.c_{1}|m-n|-c_{2}\leqslant d(g^{m}o,g^{n}o)\leqslant c_{3}|m-n|+c_{4}.

For every loxodromic element gΓ,g\in\Gamma, the set

E(g){hΓ:dH(hgo,go)<}E(g)\coloneqq\left\{h\in\Gamma:d_{\mathrm{H}}(h\left\langle g\right\rangle o,\left\langle g\right\rangle o)<\infty\right\}

is a subgroup of Γ\Gamma satisfying [E(g):g]<.[E(g):\left\langle g\right\rangle]<\infty. We use Ax(g)=E(g)oX\operatorname{\mathrm{Ax}}(g)=E(g)o\subset X to denote the axis corresponding to g,g, which is a quasi convex subset. We also remark that if g1E(g2)g_{1}\in E(g_{2}) is also loxodromic then E(g1)=E(g2).E(g_{1})=E(g_{2}). Two loxodromic elements g1,g2g_{1},g_{2} are called independent222This definition of independence is different with the one in [Yan19]. But this is enough for our applications for hyperbolic groups. if E(g1)E(g2),E(g_{1})\neq E(g_{2}), namely E(g1)E(g2)E(g_{1})\cap E(g_{2}) is finite. Moreover, we have the following bounded intersection property.

Lemma 3.2.

Let Ax(g1),Ax(g2)X\operatorname{\mathrm{Ax}}(g_{1}),\operatorname{\mathrm{Ax}}(g_{2})\subset X be different axes, then for every r>0,r>0, 𝒩r(Ax(g1))𝒩r(Ax(g2))\mathcal{N}_{r}(\operatorname{\mathrm{Ax}}(g_{1}))\cap\mathcal{N}_{r}(\operatorname{\mathrm{Ax}}(g_{2})) is bounded.

Proof.

Since dH(Ax(gi),gio)<,d_{\mathrm{H}}(\operatorname{\mathrm{Ax}}(g_{i}),\left\langle g_{i}\right\rangle o)<\infty, it suffices to show that 𝒩r(g1o))𝒩r(g2o)\mathcal{N}_{r}(\left\langle g_{1}\right\rangle o))\cap\mathcal{N}_{r}(\left\langle g_{2}\right\rangle o) is bounded. Otherwise, there exists infinitely many pairs of integers (n1,n2)(n_{1},n_{2}) such that d(g1n1o,g2n2o)2r.d(g_{1}^{n_{1}}o,g_{2}^{n_{2}}o)\leqslant 2r. Then there exists (n1,n2)(n1,n2)(n_{1},n_{2})\neq(n_{1}^{\prime},n_{2}^{\prime}) such that

g2n2g1n1=g2n2g1n1.g_{2}^{-n_{2}}g_{1}^{n_{1}}=g_{2}^{-n_{2}^{\prime}}g_{1}^{n_{1}^{\prime}}.

This implies that g1n1n1=g2n2n2g_{1}^{n_{1}^{\prime}-n_{1}}=g_{2}^{n_{2}^{\prime}-n_{2}}, which contradicts Ax(g1)Ax(g2).\operatorname{\mathrm{Ax}}(g_{1})\neq\operatorname{\mathrm{Ax}}(g_{2}).

In the case of Γ\Gamma is torsion-free, every nontrivial element is loxodromic. By a classification of virtually cyclic group [Hem04, Lemma 11.4], E(g)E(g) is cyclic for every nontrivial element gΓ.g\in\Gamma. Moreover, g1g_{1} and g2g_{2} are independent if and only if E(g1)E(g2)={id}.E(g_{1})\cap E(g_{2})=\left\{\mathrm{id}\right\}.

3.2 An extension lemma

Now we give the main technical tool for showing Proposition 3.1. It is a variant of [Yan19, Lemma 2.19], which states more generally for the group with contracting elements. A direct proof for the case of hyperbolic groups is given in this section, which is also inspired by the work of W. Yang.

For a subset SΓ,S\subset\Gamma, we say SS is RR-separated for R>0R>0 if {so:sS}\left\{so:s\in S\right\} is RR-separated in X.X.

Proposition 3.3.

Let Γ\Gamma be a non-elementary hyperbolic group. Let FΓF\subset\Gamma be a finite subset of pairwise independent loxodromic elements with #F3.\#F\geqslant 3. Then there exist m0,R,C1,L0>0m_{0},R,C_{1},L_{0}>0 such that the following holds.

For every RR-separated subset SA(L)S\subset A(L) for some LL0L\geqslant L_{0} there exists a subset SSS^{\prime}\subset S with #SC11#S\#S^{\prime}\geqslant C_{1}^{-1}\#S and FFF^{\prime}\subset F with #F=#F2\#F^{\prime}=\#F-2 such that for every mm0,m\geqslant m_{0},

S~{sfς:sS,fF,ς=m,2m}Γ\widetilde{S}\coloneqq\left\{sf^{\varsigma}:s\in S^{\prime},f\in F^{\prime},\varsigma=m,2m\right\}\subset\Gamma

freely generates a free semigroup. Furthermore, there exists C2>0C_{2}>0 only depends on FF such that for any sequence of elements s~1,,s~kS~,\widetilde{s}_{1},\cdots,\widetilde{s}_{k}\in\widetilde{S}, we have

|s~1s~k|i=1k|s~i|kC2.|\widetilde{s}_{1}\cdots\widetilde{s}_{k}|\geqslant\sum_{i=1}^{k}|\widetilde{s}_{i}|-kC_{2}. (3.2)

Before going through the proof, we need some preparation on hyperbolic groups. Recall that XX is a δ\delta-hyperbolic space. Let x,yz=(d(x,z)+d(y,z)d(x,y))/2\left\langle x,y\right\rangle_{z}=(d(x,z)+d(y,z)-d(x,y))/2 be the Gromov product, then d(z,[x,y])10δx,yzd(z,[x,y]).d(z,[x,y])-10\delta\leqslant\left\langle x,y\right\rangle_{z}\leqslant d(z,[x,y]).

Definition 3.4.

A (τ,D)(\tau,D)-chain is a sequence of points x0,x1,,xnXx_{0},x_{1},\cdots,x_{n}\in X such that

  • xi1,xi+1xiτ\left\langle x_{i-1},x_{i+1}\right\rangle_{x_{i}}\leqslant\tau for every 1in1,1\leqslant i\leqslant n-1, and

  • d(xi1,xi)Dd(x_{i-1},x_{i})\geqslant D for every 1in.1\leqslant i\leqslant n.

By connecting two consecutive points in a (τ,D),(\tau,D), we obtain a quasi-geodesic. The following lemma shows such path is indeed a uniform quasi-geodesic. See also [Gou22, Section 3B].

Lemma 3.5.

For every τ>0,\tau>0, there exists D=D(τ),c=c(τ)>0D=D(\tau),c=c(\tau)>0 such that for every (τ,D)(\tau,D)-chain x0,x1,,xn,x_{0},x_{1},\cdots,x_{n}, the path i=1n[xi1,xi]\bigcup_{i=1}^{n}[x_{i-1},x_{i}] is a cc-quasi-geodesic.

The following it a bounded projection property for different axes. A general version for spaces with contracting elements can be found in [Yan19, Lemma 2.17].

Lemma 3.6.

Let f1,f2Γf_{1},f_{2}\in\Gamma be independent loxodromic elements. There exists τ1>0\tau_{1}>0 such that for every gΓ,g\in\Gamma, we have

min{d(o,πAx(f1)(go)),d(o,πAx(f2)(go))}τ1.\min\left\{d(o,\pi_{\operatorname{\mathrm{Ax}}(f_{1})}(go)),d(o,\pi_{\operatorname{\mathrm{Ax}}(f_{2})}(go))\right\}\leqslant\tau_{1}.
Proof.

Let ziπAx(fi)(go),z_{i}\in\pi_{\operatorname{\mathrm{Ax}}(f_{i})}(go), we first show that [o,zi][zi,go][o,z_{i}][z_{i},go] is a c1c_{1}-quasi-geodesic for some c1>0c_{1}>0 only depends on Ax(f1)\operatorname{\mathrm{Ax}}(f_{1}) and Ax(f2).\operatorname{\mathrm{Ax}}(f_{2}). Note that there exists c2>0c_{2}>0 such that Ax(fi)\operatorname{\mathrm{Ax}}(f_{i}) is c2c_{2}-quasi-convex. Fixing an i{1,2},i\in\left\{1,2\right\}, for every x[o,zi]x\in[o,z_{i}] and y[zi,go],y\in[z_{i},go], we have

d(y,x)d(y,Ax(fi))c2=d(y,zi)c2.d(y,x)\geqslant d(y,\operatorname{\mathrm{Ax}}(f_{i}))-c_{2}=d(y,z_{i})-c_{2}.

Hence d(x,zi)+d(zi,y)d(x,y)+2d(zi,y)3d(x,y)+2c2d(x,z_{i})+d(z_{i},y)\leqslant d(x,y)+2d(z_{i},y)\leqslant 3d(x,y)+2c_{2} is a (3+2c2)(3+2c_{2})-quasi-geodesic.

By Morse lemma, [o,zi]𝒩c([o,go])[o,z_{i}]\subset\mathcal{N}_{c}([o,go]) where c=c(c1)>0.c=c(c_{1})>0. If both d(z1,o)d(z_{1},o) and d(z2,o)d(z_{2},o) are larger than τ1,\tau_{1}, we can choose zi[o,zi]z_{i}^{\prime}\in[o,z_{i}] with d(o,zi)=τ1.d(o,z_{i}^{\prime})=\tau_{1}. Let wi[o,go]w_{i}\in[o,go] with d(wi,zi)c,d(w_{i},z_{i}^{\prime})\leqslant c, then |d(wi,o)τ1|c.|d(w_{i},o)-\tau_{1}|\leqslant c. Hence w1,w2𝒩3c(Ax(f1))𝒩3c(Ax(f2)).w_{1},w_{2}\in\mathcal{N}_{3c}(\operatorname{\mathrm{Ax}}(f_{1}))\cap\mathcal{N}_{3c}(\operatorname{\mathrm{Ax}}(f_{2})). Since d(w1,o)τ1c,d(w_{1},o)\geqslant\tau_{1}-c, this contradicts Lemma 3.2 for a sufficiently large τ1.\tau_{1}.

Definition 3.7.

Let gΓg\in\Gamma and τ>0,\tau>0, we call a loxodromic element fΓf\in\Gamma is τ\tau-contracting for gg if for every hE(f),h\in E(f), we have go,hooτ\left\langle go,ho\right\rangle_{o}\leqslant\tau and g1o,hooτ.\left\langle g^{-1}o,ho\right\rangle_{o}\leqslant\tau.

Lemma 3.8.

Let FΓF\subset\Gamma be a finite subset of pairwise independent loxodromic elements with #F3.\#F\geqslant 3. There exists τ>0\tau>0 such that for every gΓ,g\in\Gamma, there exists FgFF_{g}\subset F with #Fg=#F2\#F_{g}=\#F-2 such that every element in FgF_{g} is τ\tau-contracting for g.g.

Proof.

By the previous lemma, there exists τ1>0\tau_{1}>0 such that for every f1,f2Ff_{1},f_{2}\in F and gΓ,g\in\Gamma, we have min{d(o,πAx(f1)(go)),d(o,πAx(f2)(go))}τ1.\min\left\{d(o,\pi_{\operatorname{\mathrm{Ax}}(f_{1})}(go)),d(o,\pi_{\operatorname{\mathrm{Ax}}(f_{2})}(go))\right\}\leqslant\tau_{1}. Note that there exists c1>0c_{1}>0 such that for every fF,f\in F, Ax(f)\operatorname{\mathrm{Ax}}(f) is c1c_{1}-quasi-convex. We take τ=τ1+c1+10δ.\tau=\tau_{1}+c_{1}+10\delta.

For each gΓg\in\Gamma and fFf\in F satisfying d(o,πAx(f)(go))τ1.d(o,\pi_{\operatorname{\mathrm{Ax}}(f)}(go))\leqslant\tau_{1}. For every hE(f),h\in E(f), we have [o,ho]𝒩c1(Ax(f)).[o,ho]\in\mathcal{N}_{c_{1}}(\operatorname{\mathrm{Ax}}(f)). Hence

d(go,[o,ho])d(go,Ax(f))c1d(go,o)d(o,πAx(f)(go))c1d(o,go)τ1c1.d(go,[o,ho])\geqslant d(go,\operatorname{\mathrm{Ax}}(f))-c_{1}\geqslant d(go,o)-d(o,\pi_{\operatorname{\mathrm{Ax}}(f)}(go))-c_{1}\geqslant d(o,go)-\tau_{1}-c_{1}.

Then o,hogod(go,[o,ho])10δd(o,go)τ1c110δ=d(o,go)τ.\left\langle o,ho\right\rangle_{go}\geqslant d(go,[o,ho])-10\delta\geqslant d(o,go)-\tau_{1}-c_{1}-10\delta=d(o,go)-\tau. We obtain

go,hoo=d(o,go)o,hogoτ.\left\langle go,ho\right\rangle_{o}=d(o,go)-\left\langle o,ho\right\rangle_{go}\leqslant\tau.

To complete the proof, we apply the argument to both gg and g1.g^{-1}. By the previous lemma, there are at least #F2\#F-2 elements in FF which are τ\tau-contracting for g.g.

Now we are at the stage of proving Proposition 3.3.

Proof of Proposition 3.3..

We apply the previous lemma to FF and obtain a constant τ>0.\tau>0. Let C1=#F(#F1)/2C_{1}=\#F(\#F-1)/2. By the pigeonhole principle, there exists FFF^{\prime}\subset F with #F=#F2\#F^{\prime}=\#F-2 such that

S={sS:Fs=F}S^{\prime}=\left\{s\in S:F_{s}=F^{\prime}\right\}

has the cardinality at least C11#S,C_{1}^{-1}\#S, where FsFF_{s}\subset F is the subset given by the previous lemma consisting of τ\tau-contracting elements.

By Lemma 3.5, there exists D=D(τ)>0D=D(\tau)>0 and c1=c1(τ)>0c_{1}=c_{1}(\tau)>0 such that every (τ,D)(\tau,D)-chain forms a c1c_{1}-quasi-geodesic. By Morse Lemma, every cc-quasi-geodesic α\alpha is contained in 𝒩C([α,α+])\mathcal{N}_{C}([\alpha_{-},\alpha_{+}]) for some C>0.C>0. Now we take C2=2C,R=4C+1,C_{2}=2C,R=4C+1, L0=D+4C.L_{0}=D+4C. The choice of m0m_{0} will be given later. Let c2>0c_{2}>0 such that every axis Ax(f)\operatorname{\mathrm{Ax}}(f) is c2c_{2}-quasi-convex for fF.f\in F. Now we show that S~\widetilde{S} freely generates a free semigroup.

Let k1k\geqslant 1 and siS,fiF,ςi{m,2m}s_{i}\in S^{\prime},f_{i}\in F^{\prime},\varsigma_{i}\in\left\{m,2m\right\} for 1ik.1\leqslant i\leqslant k. We consider

xi=s1f1ς1sifiςio,yi=s1f1ς1sio.x_{i}=s_{1}f_{1}^{\varsigma_{1}}\cdots s_{i}f_{i}^{\varsigma_{i}}o,\quad y_{i}=s_{1}f_{1}^{\varsigma_{1}}\cdots s_{i}o.

We claim that x0,y1,x1,,yk,xkx_{0},y_{1},x_{1},\cdots,y_{k},x_{k} is a (τ,D)(\tau,D)-chain. Assume that m0m_{0} is large enough guaranteeing |fm|>D|f^{m}|>D for every fFf\in F and mm0.m\geqslant m_{0}. Since |si|=LD+4C|s_{i}|=L\geqslant D+4C and |fiςi|D|f_{i}^{\varsigma_{i}}|\geqslant D for each i,i, the second condition in Definition 3.4 is verified. Besides, for each i,i, we have

xi1,xiyi=si1o,fiςiooτ,yi,yi+1xi=fiςio,si+1ooτ\left\langle x_{i-1},x_{i}\right\rangle_{y_{i}}=\left\langle s_{i}^{-1}o,f_{i}^{\varsigma_{i}}o\right\rangle_{o}\leqslant\tau,\quad\left\langle y_{i},y_{i+1}\right\rangle_{x_{i}}=\left\langle f_{i}^{-\varsigma_{i}}o,s_{i+1}o\right\rangle_{o}\leqslant\tau

by the τ\tau-contracting property.

Hence α=[x0,y1][y1,x1][xk1,yk][yk,xk]\alpha=[x_{0},y_{1}][y_{1},x_{1}]\cdots[x_{k-1},y_{k}][y_{k},x_{k}] is a c1c_{1}-quasi geodesic, which is contained in the CC-neighborhood of [α,α+]=[o,xk].[\alpha_{-},\alpha_{+}]=[o,x_{k}]. Then we have two estimates on the length of d(o,xk).d(o,x_{k}). Firstly, since |si|D+4C|s_{i}|\geqslant D+4C and |fiςi|D,|f_{i}^{\varsigma_{i}}|\geqslant D, we have

d(o,xk)i=1k(d(xi1,yi)+d(yi,xi))4kCk(2D+2C)4kC>0.d(o,x_{k})\geqslant\sum_{i=1}^{k}(d(x_{i-1},y_{i})+d(y_{i},x_{i}))-4kC\geqslant k(2D+2C)-4kC>0. (3.3)

This implies that xko.x_{k}\neq o. Besides, we have

d(o,xk)i=1kd(xi1,xi)2kC=i=1kd(xi1,xi)kC2,d(o,x_{k})\geqslant\sum_{i=1}^{k}d(x_{i-1},x_{i})-2kC=\sum_{i=1}^{k}d(x_{i-1},x_{i})-kC_{2},

which gives the desired estimate (3.2).

Checking the freeness.

We consider two such sequences si,fi,ςis_{i},f_{i},\varsigma_{i} for 1ik1\leqslant i\leqslant k and sj,fj,ςjs_{j}^{\prime},f_{j}^{\prime},\varsigma_{j}^{\prime} for 1j.1\leqslant j\leqslant\ell. We get two sequences of points xi,yix_{i},y_{i} and xj,yjx_{j}^{\prime},y_{j}^{\prime} in X.X. Assume that xk=x,x_{k}=x_{\ell}^{\prime}, we are going to show that k=k=\ell and si=si,fi=fi,s_{i}=s_{i}^{\prime},f_{i}=f_{i}^{\prime}, ςi=ςi.\varsigma_{i}=\varsigma_{i}^{\prime}. By an inductive argument, it suffices to check for i=1.i=1. Let α\alpha be a fixed geodesic connecting oo and xk=x.x_{k}=x_{\ell}^{\prime}. Then there exist z1,z1αz_{1},z_{1}^{\prime}\in\alpha such that d(y1,z1),d(y1,z1)C.d(y_{1},z_{1}),d(y_{1}^{\prime},z_{1}^{\prime})\leqslant C. Since d(y1,o)=d(y1,o)=L,d(y_{1},o)=d(y_{1}^{\prime},o)=L, we have d(z1,z1)=|d(o,z1)d(o,z1)|2C.d(z_{1},z_{1}^{\prime})=|d(o,z_{1})-d(o,z_{1}^{\prime})|\leqslant 2C. This implies that d(y1,y1)4C.d(y_{1},y_{1}^{\prime})\leqslant 4C. By the RR-separation, y1=y1.y_{1}=y_{1}^{\prime}.

Now we consider a geodesic β\beta connecting y1=y1y_{1}=y_{1}^{\prime} and xk=x.x_{k}=x_{\ell}^{\prime}. Without loss of generality, we assume that d(y1,x1)d(y1,x1).d(y_{1},x_{1})\leqslant d(y_{1}^{\prime},x_{1}^{\prime}). Take a point wπβ(x1)w\in\pi_{\beta}(x_{1}) then d(x1,w)C.d(x_{1},w)\leqslant C. Letting x[y1,x1]x^{\prime}\in[y_{1}^{\prime},x_{1}^{\prime}] with d(y1,x1)=d(x1,y1),d(y_{1}^{\prime},x_{1}^{\prime})=d(x_{1},y_{1}), we have d(x,w)3Cd(x^{\prime},w)\leqslant 3C since [y1,x1]𝒩C(β).[y_{1}^{\prime},x_{1}^{\prime}]\in\mathcal{N}_{C}(\beta). This implies w𝒩3C([y1,x1])𝒩3C([y1,x1]).w\in\mathcal{N}_{3C}([y_{1},x_{1}])\cap\mathcal{N}_{3C}([y_{1}^{\prime},x_{1}^{\prime}]). By the quasi-convexity of axes, we have [y1,x1]s1𝒩c2(Ax(f1))[y_{1},x_{1}]\subset s_{1}\mathcal{N}_{c_{2}}(\operatorname{\mathrm{Ax}}(f_{1})) and [y1,x1]s1𝒩c2(Ax(f1)).[y_{1}^{\prime},x_{1}^{\prime}]\subset s_{1}\mathcal{N}_{c_{2}}(\operatorname{\mathrm{Ax}}(f_{1}^{\prime})). Therefore, both y1y_{1} and ww are contained in s1(𝒩3C+c2(Ax(f1))𝒩3C+c2(Ax(f1))).s_{1}(\mathcal{N}_{3C+c_{2}}(\operatorname{\mathrm{Ax}}(f_{1}))\cap\mathcal{N}_{3C+c_{2}}(\operatorname{\mathrm{Ax}}(f_{1}^{\prime}))). Note that d(y1,w)d(y1,x1)d(x1,w)|f1ς1|C,d(y_{1},w)\geqslant d(y_{1},x_{1})-d(x_{1},w)\geqslant|f_{1}^{\varsigma_{1}}|-C, we conclude that

diam𝒩3C+c2(Ax(f1))𝒩3C+c2(Ax(f1))d(y1,w)|f1ς1|C.\operatorname{\mathrm{diam}}\mathcal{N}_{3C+c_{2}}(\operatorname{\mathrm{Ax}}(f_{1}))\cap\mathcal{N}_{3C+c_{2}}(\operatorname{\mathrm{Ax}}(f_{1}^{\prime}))\geqslant d(y_{1},w)\geqslant|f_{1}^{\varsigma_{1}}|-C.

By Lemma 3.2, the (3C+c2)(3C+c_{2})-neighborhoods of axes of different elements in FF have uniformly bounded intersections. We take m0m_{0} sufficiently large at beginning such that |fm|C|f^{m}|-C is strictly larger than diameters of all such intersections for every fF,mm0f\in F,m\geqslant m_{0}, this forces f1=f1.f_{1}=f_{1}^{\prime}.

Finally, we should show ς1=ς1.\varsigma_{1}=\varsigma_{1}^{\prime}. Otherwise, we assume that ς1=m\varsigma_{1}=m and ς1=2m.\varsigma_{1}^{\prime}=2m. In this case, we have

y2=s1f1ms2o,x1=s1f1mo,x1=s1f12mo,y2=s1f12ms2o.y_{2}=s_{1}f_{1}^{m}s_{2}o,\quad x_{1}=s_{1}f_{1}^{m}o,\quad x_{1}^{\prime}=s_{1}f_{1}^{2m}o,\quad y_{2}^{\prime}=s_{1}f_{1}^{2m}s_{2}^{\prime}o.

By the τ\tau-contracting property, these points form a (τ,D)(\tau,D)-chain. This leads to that

xk,yk,,y2,x1,x1,y2,,y,x.x_{k},y_{k},\cdots,y_{2},x_{1},x_{1}^{\prime},y_{2}^{\prime},\cdots,y_{\ell}^{\prime},x_{\ell}^{\prime}.

is also a (τ,D)(\tau,D)-chain. Applying the estimate in (3.3), this contradicts xk=x.x_{k}=x_{\ell}^{\prime}.

3.3 Proof of Proposition 3.1.

To prove Proposition 3.1, we need to find a finite subset consisting of pairwise independent loxodromic elements satisfying the desired condition on Zariski closures.

Lemma 3.9.

Let Γ\Gamma be a non-elementary torsion-free hyperbolic group and ρ:ΓSLn()\rho:\Gamma\to\mathrm{SL}_{n}({\mathbb{R}}) be a faithful representation. Let 𝐆{\mathbf{G}} be the Zariski closure of ρ(Γ)\rho(\Gamma) and assume 𝐆{\mathbf{G}} is Zariski connected. For every k0,k\geqslant 0, there exists a finite subset FΓF\subset\Gamma consisting of pairwise independent nontrivial elements such that for every subset FFF^{\prime}\subset F with #FFk,\#F^{\prime}\geqslant F-k, ρ(F)\rho(F^{\prime}) generates a semigroup whose Zarski closure is 𝐆.{\mathbf{G}}.

Proof.

We will find F=FkF=F_{k} for each k0k\geqslant 0 by an induction on k.k.

For the case of k=0,k=0, applying [MS23, Lemma 3.6], we can find a finite subset F0ΓF_{0}\subset\Gamma such that ρ(F0)\rho(F_{0}) generates a semigroup whose Zariski closure is 𝐆.{\mathbf{G}}. Furthermore, we can assume that elements in F0F_{0} are pairwise independent by the following process. If f1,f2F0f_{1},f_{2}\in F_{0} are not independent, then there exists fΓf\in\Gamma and m1,m2m_{1},m_{2}\in{\mathbb{Z}} such that fi=fmif_{i}=f^{m_{i}} since Γ\Gamma is torsion free. We can remove f1,f2f_{1},f_{2} and add ff in F0.F_{0}.

Assume that Fk1F_{k-1} is found. Recall that for every fΓ,f\in\Gamma, E(f)E(f) is a cyclic group containing all elements which are not independent of f.f. Notice that Zariski closure of ρ(E(f))\rho(E(f)) is commutative. On the other hand, applying Tits alternative for hyperbolic groups, Γ\Gamma contains a nonabelian free subgroup. This implies that ρ(E(f))\rho(E(f)) is not Zariski dense in 𝐆.{\mathbf{G}}. By the connectivity of 𝐆,{\mathbf{G}}, the set S=ΓfFk1E(f)S=\Gamma\setminus\bigcup_{f\in F_{k-1}}E(f) is Zariski dense in 𝐆.{\mathbf{G}}. Applying [MS23, Lemma 3.6] and the argument for k=0k=0 case, we can find a finite subset F~S\widetilde{F}\subset S consisting of pairwise independent elements such that ρ(F~)\rho(\widetilde{F}) generates a semigroup whose Zariski closure is 𝐆.{\mathbf{G}}. We take Fk=Fk1F~,F_{k}=F_{k-1}\cup\widetilde{F}, which gives a desired construction for the kk-case. ∎

Proof of 3.1.

Applying Lemma 3.9 to the case of k=2,k=2, we can find a finite subset FΓF\subset\Gamma consisting of pairwise independent elements such that for every FFF^{\prime}\subset F with #F=#F2,\#F^{\prime}=\#F-2, ρ(F)\rho(F^{\prime}) generates a semigroup whose Zariski closure is 𝐆.{\mathbf{G}}. We apply Proposition 3.3 to F.F. Let m0m_{0} be the constant given by this proposition. In order to find mm0m\geqslant m_{0} satisfying the first condition in the proposition, we need the following lemma.

Lemma 3.10.

For every gSLn(),g\in\mathrm{SL}_{n}(\mathbb{R}), there exists a positive integer \ell such that for every mm coprime with ,\ell, the Zariski closure of gm\left\langle g^{m}\right\rangle is equal to the Zariski closure of g.\left\langle g\right\rangle.

Proof.

Let 𝐇{\mathbf{H}} be the Zariski closure of g.\left\langle g\right\rangle. Let \ell be the number of connected components of 𝐇.{\mathbf{H}}. Let 𝐇{\mathbf{H}}^{\prime} be the Zariski closure of gm\left\langle g^{m}\right\rangle where mm is coprime with .\ell. Then 𝐇{\mathbf{H}}^{\prime} is a finite index algebraic subgroup of 𝐇{\mathbf{H}} and hence a union of some connected components of 𝐇.{\mathbf{H}}. Note that the action of g\left\langle g\right\rangle given by left translation is transitive among connected components of 𝐇{\mathbf{H}} and so does gm\left\langle g^{m}\right\rangle, due to mm coprime with \ell. Hence 𝐇=𝐇.{\mathbf{H}}^{\prime}={\mathbf{H}}.

We take mm0m\geqslant m_{0} to be a sufficiently large prime number such that for every fF,f\in F, the Zariski closure of ρ(f)m\left\langle\rho(f)^{m}\right\rangle equals to the Zariski closure of ρ(f).\left\langle\rho(f)\right\rangle. Then for every FFF^{\prime}\subset F with #F=#F2,\#F^{\prime}=\#F-2, the Zariski closure of the semigroup generated by {ρ(f)m:fF}\left\{\rho(f)^{m}:f\in F^{\prime}\right\} equals to that of ρ(F),\rho(F), which is 𝐆.{\mathbf{G}}. The first condition holds.

Note that for every finite subset SΓ,S\subset\Gamma, it contains an RR-separated subset of cardinality at least (#𝒮+1)R#S,(\#\mathcal{S}+1)^{-R}\#S, where 𝒮\mathcal{S} is the symmetric generating set of Γ.\Gamma. Then the last two conditions follow from Proposition 3.3 by enlarging C1C_{1} suitably. ∎

4 Variational principle for Anosov representations

In this section, we will show a variational principle for dimensions of limit sets of Borel Anosov representations. Let Γ\Gamma be a finitely generated hyperbolic group with a fixed finite symmetric generating set 𝒮Γ.\mathcal{S}\subset\Gamma. Let ρ:ΓSLn()\rho:\Gamma\to\mathrm{SL}_{n}({\mathbb{R}}) be an irreducible Borel Anosov representation. Recall that κ(ρ(γ))=diag(logσ1(ρ(γ)),,logσn(ρ(γ)))\kappa(\rho(\gamma))=\mathrm{diag}(\log\sigma_{1}(\rho(\gamma)),\cdots,\log\sigma_{n}(\rho(\gamma))) is the Cartan projection of ρ(γ).\rho(\gamma).

4.1 The variational principle to positive linear functions on 𝔞+\mathfrak{a}^{+}

Recall the Cartan algebra 𝔞={λ=diag(λ1,,λn):λi,λi=0}\mathfrak{a}=\{\lambda=\mathrm{diag}(\lambda_{1},\cdots,\lambda_{n}):\lambda_{i}\in\mathbb{R},\ \sum\lambda_{i}=0\} and a positive Weyl chamber 𝔞+\mathfrak{a}^{+} mentioned in Section 2.1. Let ψ\psi be a linear function on 𝔞\mathfrak{a} which is positive with respect to 𝔞+.\mathfrak{a}^{+}. Specifically, ψ\psi can be expressed as

ψ(λ)=i=1n1aiαi(λ),\psi(\lambda)=\sum_{i=1}^{n-1}a_{i}\cdot\alpha_{i}(\lambda),

where a1,,an10a_{1},\cdots,a_{n-1}\geqslant 0 are not all zero and αi(λ)=λiλi+1\alpha_{i}(\lambda)=\lambda_{i}-\lambda_{i+1} are simple roots.

Recall that for a finitely supported probability measure ν\nu on SLn(),\mathrm{SL}_{n}({\mathbb{R}}), the Lyapunov spectrum of ν\nu is λ(ν)=(λ1(ν),,λn(ν)).\lambda(\nu)=(\lambda_{1}(\nu),\cdots,\lambda_{n}(\nu)). We also view λ(ν)\lambda(\nu) as an element in 𝔞+\mathfrak{a}^{+} using the isomorphism 𝔞n\mathfrak{a}\cong{\mathbb{R}}^{n} and ψ\psi acts on λ(ν).\lambda(\nu). Then ψ(λ(ν))\psi(\lambda(\nu)) is a nonnegative number.

Throughout this subsection, we further assume that Γ\Gamma is torsion-free and ρ\rho is faithful. Note that Γ\Gamma is non-cyclic and hence non-elementary since ρ\rho is irreducible.

Proposition 4.1.

Let 𝐆{\mathbf{G}} be the Zariski closure of ρ(Γ)\rho(\Gamma) which is assumed to be Zariski connected. Let ψ\psi be a linear function as above. If the series γΓexp(ψ(κ(ρ(γ))))\sum_{\gamma\in\Gamma}\exp(-\psi(\kappa(\rho(\gamma)))) diverges, then there exists c>0c>0 such that the following holds. For every ϵ>0,\epsilon>0, there exists infinitely many positive integers NN with a finitely supported probability measure ν\nu on ρ(Γ)\rho(\Gamma) such that

  • GνG_{\nu} is Zariski dense in 𝐆.{\mathbf{G}}.

  • λp(ν)λp+1(ν)cN\lambda_{p}(\nu)-\lambda_{p+1}(\nu)\geqslant cN for every p=1,,n1.p=1,\cdots,n-1.

  • hRW(ν)(1ϵ)Nh_{\mathrm{RW}}(\nu)\geqslant(1-\epsilon)N and ψ(λ(ν))(1+ϵ)N.\psi(\lambda(\nu))\leqslant(1+\epsilon)N.

We first recall an estimate on the lost of singular values under composition. The following lemma is a direct consequence of combining Lemmas 2.5 and A.7 in [BPS19].

Lemma 4.2.

Let 1pn.1\leqslant p\leqslant n. Given c>0,c>0, then there exists δ>0\delta>0 such that the following holds. Let (γk)k𝒮(\gamma_{k})_{k\in{\mathbb{N}}}\in\mathcal{S}^{\mathbb{N}} satisfying for every m,\ell\leqslant m,

σp(ρ(γ+1γm))σp+1(ρ(γ+1γm))cec(m).\frac{\sigma_{p}(\rho(\gamma_{\ell+1}\cdots\gamma_{m}))}{\sigma_{p+1}(\rho(\gamma_{\ell+1}\cdots\gamma_{m}))}\geqslant c\cdot e^{c(m-\ell)}.

Then for every km,\ell\leqslant k\leqslant m, we have

σp(ρ(γ+1γm))δσp(ρ(γ+1γk))σp(ρ(γk+1γm)),\sigma_{p}(\rho(\gamma_{\ell+1}\cdots\gamma_{m}))\geqslant\delta\cdot\sigma_{p}(\rho(\gamma_{\ell+1}\cdots\gamma_{k}))\sigma_{p}(\rho(\gamma_{k+1}\cdots\gamma_{m})),
σp+1(ρ(γ+1γm))δ1σp+1(ρ(γ+1γk))σp+1(ρ(γk+1γm)).\sigma_{p+1}(\rho(\gamma_{\ell+1}\cdots\gamma_{m}))\leqslant\delta^{-1}\cdot\sigma_{p+1}(\rho(\gamma_{\ell+1}\cdots\gamma_{k}))\sigma_{p+1}(\rho(\gamma_{k+1}\cdots\gamma_{m})).
Proof of Proposition 4.1.

Applying Proposition 3.1, we obtain a finite subset FΓF\subset\Gamma with #F3,\#F\geqslant 3, constants C1,C2,L0>0C_{1},C_{2},L_{0}>0 and a positive integer m.m. For every ϵ>0\epsilon>0 sufficiently small, there are infinitely many integers NN such that

S1={γΓ:ψ(κ(ρ(γ)))N}S_{1}=\left\{\gamma\in\Gamma:\psi(\kappa(\rho(\gamma)))\leqslant N\right\}

has cardinality at least e(1ϵ)Ne^{(1-\epsilon)N} due to the divergence of the series. Since ψ\psi is positive, we have ψ(κ(ρ(γ)))c1(logσp(ρ(γ))logσp+1(ρ(γ)))\psi(\kappa(\rho(\gamma)))\geqslant c_{1}(\log\sigma_{p}(\rho(\gamma))-\log\sigma_{p+1}(\rho(\gamma))) for some 1pn11\leqslant p\leqslant n-1 and c1>0.c_{1}>0. Because ρ\rho is Borel Anosov, there exists c2>0c_{2}>0 such that for |γ||\gamma| large enough,

ψ(κ(ρ(γ)))c1(logσp(ρ(γ))logσp+1(ρ(γ)))c2|γ|.\psi(\kappa(\rho(\gamma)))\geqslant c_{1}(\log\sigma_{p}(\rho(\gamma))-\log\sigma_{p+1}(\rho(\gamma)))\geqslant c_{2}|\gamma|.

Hence S1{γΓ:|γ|c21N}.S_{1}\subset\left\{\gamma\in\Gamma:|\gamma|\leqslant c_{2}^{-1}N\right\}. Let c3=(2#log𝒮)1,c_{3}=(2\#\log\mathcal{S})^{-1}, then

#{γΓ:|γ|c3N}2(#𝒮)c3N12e(1ϵ)N\#\left\{\gamma\in\Gamma:|\gamma|\leqslant c_{3}N\right\}\leqslant 2(\#\mathcal{S})^{c_{3}N}\leqslant\frac{1}{2}e^{(1-\epsilon)N}

providing ϵ\epsilon small and NN large. Hence there exists c3NLc21Nc_{3}N\leqslant L\leqslant c_{2}^{-1}N such that

S2{γS1:|γ|=L}S_{2}\coloneqq\left\{\gamma\in S_{1}:|\gamma|=L\right\}

has cardinality at least c2(2N)1e(1ϵ)Ne(12ϵ)Nc_{2}(2N)^{-1}e^{(1-\epsilon)N}\geqslant e^{(1-2\epsilon)N} assuming NN large. By Proposition 3.1 , there exists S3S2S_{3}\subset S_{2} and FFF^{\prime}\subset F with #S3C11#S2\#S_{3}\geqslant C_{1}^{-1}\#S_{2} and #F=#F2\#F^{\prime}=\#F-2 such that

S~{sfς:sS3,fF,ς=m,2m}\widetilde{S}\coloneqq\left\{sf^{\varsigma}:s\in S_{3},f\in F^{\prime},\varsigma=m,2m\right\}

freely generates a free semigroup. Assuming NN large, we have #S~#S3e(13ϵ)N.\#\widetilde{S}\geqslant\#S_{3}\geqslant e^{(1-3\epsilon)N}. Letting ν\nu be the uniform measure on ρ(S~),\rho(\widetilde{S}), we now verify that this is a desired construction.

  • Note that ρ(f)m=ρ(sfm)1ρ(sf2m)Gν\rho(f)^{m}=\rho(sf^{m})^{-1}\rho(sf^{2m})\in G_{\nu} for every fF.f\in F^{\prime}. By the first condition in 3.1, we have GνG_{\nu} is Zariski dense in 𝐆.{\mathbf{G}}.

  • By the definition of Borel Anosov representations, we can take c4>0c_{4}>0 such that for every 1pn11\leqslant p\leqslant n-1 and |γ||\gamma| large enough, logσp(ρ(γ))logσp+1(ρ(γ))c4|γ|\log\sigma_{p}(\rho(\gamma))-\log\sigma_{p+1}(\rho(\gamma))\geqslant c_{4}|\gamma|. Assuming NN is large enough only depends on m,m, for every s~S~\widetilde{s}\in\widetilde{S} we have

    |s~|c3N2mmaxfF|f|c3N+C2and|s~|c21N+2mmaxfF|f|c21N|\widetilde{s}|\geqslant c_{3}N-2m\max_{f\in F}|f|\geqslant c_{3}^{\prime}N+C_{2}\quad\text{and}\quad|\widetilde{s}|\leqslant c_{2}^{-1}N+2m\max_{f\in F}|f|\leqslant c_{2}^{\prime-1}N

    where c2=c2/2,c3=c3/2c_{2}^{\prime}=c_{2}/2,c_{3}^{\prime}=c_{3}/2 and C2C_{2} is the constant given by Proposition 3.1. Then for every s~1,s~kS~,\widetilde{s}_{1},\cdots\widetilde{s}_{k}\in\widetilde{S}, due to (3.1), we have |s~1s~k|c3kN.|\widetilde{s}_{1}\cdots\widetilde{s}_{k}|\geqslant c_{3}^{\prime}kN. Hence for every 1pn1,1\leqslant p\leqslant n-1, we have

    (logσplogσp+1)(ρ(s~1s~k))c4c3kNc2c3c4i=1k|s~i|.(\log\sigma_{p}-\log\sigma_{p+1})(\rho(\widetilde{s}_{1}\cdots\widetilde{s}_{k}))\geqslant c_{4}c_{3}^{\prime}kN\geqslant c_{2}^{\prime}c_{3}^{\prime}c_{4}\sum_{i=1}^{k}|\widetilde{s}_{i}|. (4.1)

    Recall that suppν=S~\operatorname{\mathrm{supp}}\nu=\widetilde{S}. By the first inequality in (4.1) we have

    λp(ν)λp+1(ν)=limk1k[logσp(ρ(γ))logσp+1(ρ(γ))]dνk(γ)c4c3N.\lambda_{p}(\nu)-\lambda_{p+1}(\nu)=\lim_{k\to\infty}\frac{1}{k}\int[\log\sigma_{p}(\rho(\gamma))-\log\sigma_{p+1}(\rho(\gamma))]\mathrm{d}\nu^{*k}(\gamma)\geqslant c_{4}c_{3}^{\prime}N.

    Taking c=c4c3,c=c_{4}c_{3}^{\prime}, we obtain the conclusion.

  • Since S~\widetilde{S} freely generates a free semigroup and ν\nu is the uniform measure on S~,\widetilde{S}, we have

    hRW(ν)=log#S~(13ϵ)N.h_{\mathrm{RW}}(\nu)=\log\#\widetilde{S}\geqslant(1-3\epsilon)N.

    Shrinking ϵ>0,\epsilon>0, we obtain the first estimate.

    In order to estimate ψ(λ(ν)),\psi(\lambda(\nu)), we need an almost additivity property of logσp.\log\sigma_{p}. Recall the second inequality in (4.1) and the constant cancellation property in (3.1), which verifies the condition of 4.2. By applying Lemma 4.2, there exists δ>0\delta>0 only depending on C2C_{2} and c2c3c4c_{2}^{\prime}c_{3}^{\prime}c_{4} such that

    |logσp(ρ(s~1s~k))i=1klogσp(s~i)|klogδ\left|\log\sigma_{p}(\rho(\widetilde{s}_{1}\cdots\widetilde{s}_{k}))-\sum_{i=1}^{k}\log\sigma_{p}(\widetilde{s}_{i})\right|\leqslant-k\log\delta

    for every s~1,,s~kS~\widetilde{s}_{1},\cdots,\widetilde{s}_{k}\in\widetilde{S} and 1pn.1\leqslant p\leqslant n. Since ψ\psi is a linear function, we have

    |ψ(κ(ρ(s~1s~k)))i=1kψ(κ(ρ(s~i)))|kC3logδ,\left|\psi(\kappa(\rho(\widetilde{s}_{1}\cdots\widetilde{s}_{k})))-\sum_{i=1}^{k}\psi(\kappa(\rho(\widetilde{s}_{i})))\right|\leqslant-kC_{3}\log\delta,

    where C3>0C_{3}>0 only depends on ψ.\psi. For each s~S~,\widetilde{s}\in\widetilde{S}, write s~=sfς,\widetilde{s}=sf^{\varsigma}, where sS3s\in S_{3} and fF,ς{m,2m}.f\in F,\varsigma\in\left\{m,2m\right\}. Then there exists C4>0C_{4}>0 only depending on FF and mm such that |logσp(ρs~)logσp(ρs)|C4|\log\sigma_{p}(\rho\widetilde{s})-\log\sigma_{p}(\rho s)|\leqslant C_{4} for every p.p. Hence ψ(κ(ρs~))N+C3C4\psi(\kappa(\rho\widetilde{s}))\leqslant N+C_{3}C_{4} since sS3S1.s\in S_{3}\subset S_{1}.

    Then for sufficiently large N,N, we have

    ψ(κ(ρ(s~1s~k)))kN+kC3C4kC3logδ(1+ϵ)kN.\psi(\kappa(\rho(\widetilde{s}_{1}\cdots\widetilde{s}_{k})))\leqslant kN+kC_{3}C_{4}-kC_{3}\log\delta\leqslant(1+\epsilon)kN.

    This implies that

    ψ(λ(ν))=limk1kψ(κ(ρ(γ)))dνk(γ)(1+ϵ)N.\psi(\lambda(\nu))=\lim_{k\to\infty}\frac{1}{k}\int\psi(\kappa(\rho(\gamma)))\mathrm{d}\nu^{*k}(\gamma)\leqslant(1+\epsilon)N.\qed

4.2 The variational principle of critical exponent

This section is devoted to prove Theorem 1.1. We will also present a version for flag varieties. In this section, we consider a Borel Anosov representation ρ:ΓSLn()\rho:\Gamma\to\mathrm{SL}_{n}({\mathbb{R}}) and let 𝐆{\mathbf{G}} be the Zariski closure of ρ(Γ).\rho(\Gamma). Recall the affinity exponent sA(ρ)s_{A}(\rho) given in (1.1), which is expressed as

sA(ρ)sup{s:γΓexp(ψs(ρ(γ)))=}s_{\mathrm{A}}(\rho)\coloneqq\sup\left\{s:\sum_{\gamma\in\Gamma}\exp(-\psi_{s}(\rho(\gamma)))=\infty\right\}

where

ψs(g)1is(logσ1(g)logσi+1(g))+(ss)(logσ1(g)logσs+2(g)),gSLn().\psi_{s}(g)\coloneqq\sum_{1\leqslant i\leqslant\lfloor s\rfloor}(\log\sigma_{1}(g)-\log\sigma_{i+1}(g))+(s-\lfloor s\rfloor)(\log\sigma_{1}(g)-\log\sigma_{\lfloor s\rfloor+2}(g)),\quad g\in\mathrm{SL}_{n}({\mathbb{R}}).

Note that ψs(g)\psi_{s}(g) is a linear function on Cartan projection κ(g).\kappa(g). By abuse of the notation, we also denote ψs\psi_{s} to be the linear function on 𝔞\mathfrak{a} satisfying ψs(κ(g))=ψs(g).\psi_{s}(\kappa(g))=\psi_{s}(g). Then ψs\psi_{s} is positive with respect to 𝔞+.\mathfrak{a}^{+}.

Let 𝒫f.s.(ρ)\mathscr{P}_{\mathrm{f.s.}}(\rho) be the family of finitely supported probability measures on ρ(Γ).\rho(\Gamma). Let 𝒫f.s.𝐆(ρ)\mathscr{P}_{\mathrm{f.s.}}^{\mathbf{G}}(\rho) be the family of ν𝒫f.s.(ρ)\nu\in\mathscr{P}_{\mathrm{f.s.}}(\rho) satisfying GνG_{\nu} is Zariski dense in 𝐆.{\mathbf{G}}. We will also consider 𝒫f.s.𝐆0(ρ)\mathscr{P}_{\mathrm{f.s.}}^{{\mathbf{G}}^{0}}(\rho) where 𝐆0{\mathbf{G}}^{0} denotes the identity (Zariski-)component of 𝐆.{\mathbf{G}}. For ν𝒫f.s(ρ),\nu\in\mathscr{P}_{\mathrm{f.s}}(\rho), let μ\mu be an ergodic stationary measure of ν\nu on (d).{\mathbb{P}}({\mathbb{R}}^{d}). Recall the Lyapunov dimension of μ\mu in (1.3). Then Theorem 1.1 is interpreted as the following.

Proposition 4.3.

Let ρ:ΓSLn()\rho:\Gamma\to\mathrm{SL}_{n}({\mathbb{R}}) be a Zariski dense Borel Anosov representation (that is 𝐆=SLn(){\mathbf{G}}=\mathrm{SL}_{n}({\mathbb{R}})). Then

sA(ρ)\displaystyle s_{A}(\rho) sup{dimLYμ:μ is the unique stationary measure on (n) of some ν𝒫f.s.𝐆(ρ)}\displaystyle\leqslant\sup\left\{\dim_{\mathrm{LY}}\mu:\mu\text{ is the unique stationary measure on ${\mathbb{P}}({\mathbb{R}}^{n})$ of some }\nu\in\mathscr{P}_{\mathrm{f.s.}}^{\mathbf{G}}(\rho)\right\}

Besides, we also have a variational principle for dimensions on the flag variety. We consider

ψF,s(g)inf{1i<jnaijlogσi(g)σj(g):0aij1,1i<jnaij=s}.\psi_{F,s}(g)\coloneqq\inf\left\{\sum_{1\leqslant i<j\leqslant n}a_{ij}\log\frac{\sigma_{i}(g)}{\sigma_{j}(g)}:0\leqslant a_{ij}\leqslant 1,\sum_{1\leqslant i<j\leqslant n}a_{ij}=s\right\}.

Note that ψF,s(g)\psi_{F,s}(g) is increasing with respect to s.s. The affinity exponent on the flag variety is given by

sA,F(ρ)sup{s:γΓexp(ψF,s(ρ(γ)))=}.s_{A,F}(\rho)\coloneqq\sup\left\{s:\sum_{\gamma\in\Gamma}\exp(-\psi_{F,s}(\rho(\gamma)))=\infty\right\}.

For instance, in the case of n=3,n=3, the function ψF,s\psi_{F,s} is given by

ψF,s(g)={smin{logσ1(g)σ2(g),logσ2(g)σ3(g)},s1;(s1)(logσ1(g)σ2(g)+logσ2(g)σ3(g))+(2s)min{logσ1(g)σ2(g),logσ2(g)σ3(g)},1<s2;logσ1(g)σ2(g)+logσ2(g)σ3(g)+(s2)logσ1(g)σ3(g),2<s3.\psi_{F,s}(g)=\left\{\begin{aligned} &s\min\left\{\log\frac{\sigma_{1}(g)}{\sigma_{2}(g)},\log\frac{\sigma_{2}(g)}{\sigma_{3}(g)}\right\},&s\leqslant 1;\\ &(s-1)\left(\log\frac{\sigma_{1}(g)}{\sigma_{2}(g)}+\log\frac{\sigma_{2}(g)}{\sigma_{3}(g)}\right)+(2-s)\min\left\{\log\frac{\sigma_{1}(g)}{\sigma_{2}(g)},\log\frac{\sigma_{2}(g)}{\sigma_{3}(g)}\right\},&1<s\leqslant 2;\\ &\log\frac{\sigma_{1}(g)}{\sigma_{2}(g)}+\log\frac{\sigma_{2}(g)}{\sigma_{3}(g)}+(s-2)\log\frac{\sigma_{1}(g)}{\sigma_{3}(g)},&2<s\leqslant 3.\end{aligned}\right.

In general, ψF,s(g)\psi_{F,s}(g) is always a minimum of finitely many linear functions on κ(g)\kappa(g) which is positive with respect to 𝔞+.\mathfrak{a}^{+}. Specifically, ψF,s(g)=min{ψF,k(κ(g)):k𝒦s},\psi_{F,s}(g)=\min\left\{\psi_{F,k}(\kappa(g)):k\in\mathcal{K}_{s}\right\}, where each ψF,k(κ(g))\psi_{F,k}(\kappa(g)) is of the form

1i<jnaij(logσi(g)logσj(g))\sum_{1\leqslant i<j\leqslant n}a_{ij}(\log\sigma_{i}(g)-\log\sigma_{j}(g))

satisfying 0aij1,Σ1i<jnaij=s0\leqslant a_{ij}\leqslant 1,\Sigma_{1\leqslant i<j\leqslant n}a_{ij}=s and at most one of aija_{ij} is neither 0 nor 1.1. For given ss and n,n, there are at most finitely many such linear functionals. Hence #𝒦s\#\mathcal{K}_{s} is finite.

For ν𝒫f.s(ρ),\nu\in\mathscr{P}_{\mathrm{f.s}}(\rho), let μ\mu be an ergodic stationary measure of ν\nu on (n).\mathcal{F}({\mathbb{R}}^{n}). Then the Lyapunov dimension of μ\mu is given by

dimLYμ=sup{1i<jndij:0dij1,1i<jndij(λi(ν)λj(ν))=hF(μ,ν)}.\dim_{\mathrm{LY}}\mu=\sup\left\{\sum_{1\leqslant i<j\leqslant n}d_{ij}:0\leqslant d_{ij}\leqslant 1,\sum_{1\leqslant i<j\leqslant n}d_{ij}(\lambda_{i}(\nu)-\lambda_{j}(\nu))=h_{\mathrm{F}}(\mu,\nu)\right\}.
Proposition 4.4.

Let ρ:ΓSLn()\rho:\Gamma\to\mathrm{SL}_{n}({\mathbb{R}}) be a Borel Anosov representation. Then

sA,F(ρ)\displaystyle s_{A,F}(\rho) sup{dimLYμ:μ is an ergodic stationary measure on (n) of some ν𝒫f.s.𝐆0(ρ)}\displaystyle\leqslant\sup\left\{\dim_{\mathrm{LY}}\mu:\mu\text{ is an ergodic stationary measure on $\mathcal{F}({\mathbb{R}}^{n})$ of some }\nu\in\mathscr{P}_{\mathrm{f.s.}}^{{\mathbf{G}}^{0}}(\rho)\right\}

We will prove the case on flag varieties first. The case on the projective space is very similar and we will only indicate where we need to modify in the proof.

Proof of Proposition 4.4.

Since ρ\rho is Anosov, kerρ\ker\rho is finite. Replacing Γ\Gamma by Γ/kerρ,\Gamma/\ker\rho, we can assume that ρ\rho is faithful. Besides, by Selberg’s lemma, ρ(Γ)\rho(\Gamma) is virtually torsion-free. Replacing Γ\Gamma by a finite index torsion-free subgroup, we may assume that Γ\Gamma is torsion free and the Zariski closure of ρ(Γ)\rho(\Gamma) is Zariski connected. This process does not affect the value of affinity exponent and the Zariski closure of ρ(Γ)\rho(\Gamma) is indeed the identity component of the original one. Then the proposition is a direct consequence of the following lemma.

Lemma 4.5.

Let Γ\Gamma be a non-elementary torsion-free hyperbolic group and ρ:ΓSLn()\rho:\Gamma\to\mathrm{SL}_{n}({\mathbb{R}}) a faithful Borel Anosov representation such that the Zariski closure 𝐆{\mathbf{G}} of ρ(Γ)\rho(\Gamma) is Zariski connected. Let s>0s>0 such that the series γΓexp(ψF,s(ρ(γ)))\sum_{\gamma\in\Gamma}\exp(-\psi_{F,s}(\rho(\gamma))) diverges. Then for every ϵ>0\epsilon>0 sufficiently small, there exists ν𝒫f.s.𝐆(ρ)\nu\in\mathscr{P}_{\mathrm{f.s.}}^{\mathbf{G}}(\rho) and an ergodic ν\nu-stationary measure μ\mu on (n)\mathcal{F}({\mathbb{R}}^{n}) satisfying

dimLYμsϵ.\dim_{\mathrm{LY}}\mu\geqslant s-\epsilon.
Proof.

Recall that ψF,s(ρ(γ))=min{ψF,k(κ(ρ(γ))):k𝒦s}\psi_{F,s}(\rho(\gamma))=\min\left\{\psi_{F,k}(\kappa(\rho(\gamma))):k\in\mathcal{K}_{s}\right\} is a minimum of finitely many linear functions. Then

k𝒦sγΓexp(ψF,k(κ(ρ(γ))))γΓexp(ψF,s(ρ(γ)))=.\sum_{k\in\mathcal{K}_{s}}\sum_{\gamma\in\Gamma}\exp(-\psi_{F,k}(\kappa(\rho(\gamma))))\geqslant\sum_{\gamma\in\Gamma}\exp(-\psi_{F,s}(\rho(\gamma)))=\infty.

This implies that there exists k𝒦sk\in\mathcal{K}_{s} satisfying γΓexp(ψF,k(ρ(γ)))\sum_{\gamma\in\Gamma}\exp(-\psi_{F,k}(\rho(\gamma))) diverges. We fix a such ψF,k\psi_{F,k} in latter discussions and assume that

ψF,k(g)=1i<jnaij(logσi(g)logσj(g)),gSLn().\psi_{F,k}(g)=\sum_{1\leqslant i<j\leqslant n}a_{ij}(\log\sigma_{i}(g)-\log\sigma_{j}(g)),\quad\forall g\in\mathrm{SL}_{n}({\mathbb{R}}).

Now we apply Proposition 4.1 to ψF,k.\psi_{F,k}. We obtain a constant c>0c>0 and there exists a positive integer NN and ν𝒫f.s.(ρ)\nu\in\mathscr{P}_{\mathrm{f.s.}}(\rho) satisfying

  • suppν\operatorname{\mathrm{supp}}\nu generates a semigroup whose Zariski closure is 𝐆.{\mathbf{G}}.

  • λp(ν)λp+1(ν)cN\lambda_{p}(\nu)-\lambda_{p+1}(\nu)\geqslant cN for every p=1,,n1.p=1,\cdots,n-1.

  • hRW(ν)(112cϵ)Nh_{\mathrm{RW}}(\nu)\geqslant(1-\frac{1}{2}c\epsilon)N and ψF,k(λ(ν))(1+12cϵ)N.\psi_{F,k}(\lambda(\nu))\leqslant(1+\frac{1}{2}c\epsilon)N.

Since ν\nu has a simple Lyapunov spectrum, there exists a ν\nu-stationary measure μ\mu on (n)\mathcal{F}({\mathbb{R}}^{n}) which corresponds to the distribution of Oseledec’s splitting. Furthermore, (suppμ,μ)(\operatorname{\mathrm{supp}}\mu,\mu) is the Poisson boundary for (Γν,ν)(\Gamma_{\nu},\nu) by [Fur02, Theorem 2.21], where Γν\Gamma_{\nu} is the group generated by suppν.\operatorname{\mathrm{supp}}\nu. By [Fur02, Theorem 2.31], hF(μ,ν)=hRW(ν)(112cϵ)N.h_{\mathrm{F}}(\mu,\nu)=h_{\mathrm{RW}}(\nu)\geqslant(1-\frac{1}{2}c\epsilon)N. Now we estimate dimLYμ.\dim_{\mathrm{LY}}\mu.

Note that

ψF,k(λ(ν))=1i<jnaij(λi(ν)λj(ν))(1+12cϵ)N.\psi_{F,k}(\lambda(\nu))=\sum_{1\leqslant i<j\leqslant n}a_{ij}(\lambda_{i}(\nu)-\lambda_{j}(\nu))\leqslant(1+\frac{1}{2}c\epsilon)N.

Assuming ai0j0>0a_{i_{0}j_{0}}>0 for some 1i0<j0n,1\leqslant i_{0}<j_{0}\leqslant n, we take

aij={aijϵ,i=i0,j=j0;aij,otherwise.a_{ij}^{\prime}=\left\{\begin{aligned} &a_{ij}-\epsilon,&i=i_{0},j=j_{0};\\ &a_{ij},&\text{otherwise}.\end{aligned}\right.

Then

1i<jnaij(λi(ν)λj(ν))(1+12cϵ)Nϵ(λi0(ν)λj0(ν))(112cϵ)NhF(μ,ν).\sum_{1\leqslant i<j\leqslant n}a_{ij}^{\prime}(\lambda_{i}(\nu)-\lambda_{j}(\nu))\leqslant(1+\frac{1}{2}c\epsilon)N-\epsilon(\lambda_{i_{0}}(\nu)-\lambda_{j_{0}}(\nu))\leqslant(1-\frac{1}{2}c\epsilon)N\leqslant h_{\mathrm{F}}(\mu,\nu).

Hence

dimLYμ1i<jnaij=1i<jnaijϵ=sϵ.\dim_{\mathrm{LY}}\mu\geqslant\sum_{1\leqslant i<j\leqslant n}a_{ij}^{\prime}=\sum_{1\leqslant i<j\leqslant n}a_{ij}-\epsilon=s-\epsilon.

We obtain the desired conclusion. ∎

Proof of Proposition 4.3.

The only difference is to show the identity between the Furstenberg entropy and the random walk entropy. Since we assume additionally that ρ(Γ)\rho(\Gamma) is Zariski dense in SLn(),\mathrm{SL}_{n}({\mathbb{R}}), the equality of two notions of entropies follows from Proposition 2.3. ∎

5 Hausdorff dimension of the Rauzy Gasket

We will verify the equality of the Haussdorff dimension of the Rauzy gasket with its affinity exponent. To simplify notations, we abbreviate Γ\Gamma_{\mathscr{R}} to Γ\Gamma in this section.

5.1 Preliminaries and notation

Recall that Δ\Delta is the projectivization of {(x,y,z):x,y,z0}\{(x,y,z):x,y,z\geqslant 0\} in (3).{\mathbb{P}}({\mathbb{R}}^{3}). There is a natural bijection between Δ\Delta and the euclidean triangle Δ~={(x,y,z)3:x+y+z=1,x,y,z0}\widetilde{\Delta}=\{(x,y,z)\in\mathbb{R}^{3}:x+y+z=1,x,y,z\geqslant 0\} by the projective map, which also preserves lines, hence triangles. The euclidean distance dEd_{E} on Δ~\widetilde{\Delta} from 3\mathbb{R}^{3} is bi-Lipschitz equivalent to the projective distance dd coming from Δ\Delta. Since Lipschitz constants do not affect the statements of lemmas, we identify Δ\Delta and Δ~\widetilde{\Delta} in the following and we do not distinguish the euclidean metric and the projective metric. Moreover, the area of triangles in Δ\Delta will be understood as the area of the corresponding triangle in Δ~\widetilde{\Delta} in this section.

Recall that the Rauzy gasket X(3)X\subset\mathbb{P}(\mathbb{R}^{3}) is a projective fractal set defined in the introduction. We may consider the classical coding of XX by infinite words as the case of IFSs. Let Λ{1,2,3}\Lambda\coloneqq\left\{1,2,3\right\} be the set of symbols. We have the following basic fact [AS13, Lemma 3].

Lemma 5.1.

For every 𝐢=(i1,i2,)Λ,\mathbf{i}=(i_{1},i_{2},\cdots)\in\Lambda^{\mathbb{N}}, we have limndiamAi1AinΔ=0.\lim_{n\to\infty}\operatorname{\mathrm{diam}}A_{i_{1}}\cdots A_{i_{n}}\Delta=0.

This fact allows us to define the coding map

Φ:ΛΔ,𝐢=(i1,i2,)nAi1AinΔ.\Phi:\Lambda^{\mathbb{N}}\to\Delta,\quad\mathbf{i}=(i_{1},i_{2},\cdots)\mapsto\cap_{n\in{\mathbb{N}}}A_{i_{1}}\cdots A_{i_{n}}\Delta.

Then the image of Φ\Phi is exactly the Rauzy Gasket X.X.

For any γΓ\gamma\in\Gamma we denote by Δγ\Delta_{\gamma} the image γΔ\gamma\cdot\Delta. We also use |γ||\gamma| for the word length of γ\gamma with respect to the standard generator set {Ai}\{A_{i}\} of Γ\Gamma. For later use, we consider the following notations. Recall that by freeness of Γ\Gamma, any γΓ\gamma\in\Gamma can be decomposed uniquely as the following γ=Ai1Ai|γ|\gamma=A_{i_{1}}\cdots A_{i_{|\gamma|}}. For any n|γ|n\leqslant|\gamma|, we denote by γnAi1Ain\gamma_{n}\coloneqq A_{i_{1}}\cdots A_{i_{n}}. We say that the last nn digits of γ\gamma are not the same for some n|γ|n\leqslant|\gamma|, if in the decomposition, i|γ|n+1,,i|γ|i_{|\gamma|-n+1},\dots,i_{|\gamma|} are not the same. Recall sA(Γ)s_{\mathrm{A}}(\Gamma) defined in the introduction is the affinity exponent of Γ\Gamma.

We will also consider the transpose action of Γ\Gamma. For any γ=Ai1Ai|γ|Γ\gamma=A_{i_{1}}\cdots A_{i_{|\gamma|}}\in\Gamma, the transpose action γt:(3)(3)\gamma^{t}:\mathbb{P}(\mathbb{R}^{3})\to\mathbb{P}(\mathbb{R}^{3}) of γ\gamma is defined by γt:=Ai|γ|tAi1t\gamma^{t}:=A_{i_{|\gamma|}}^{t}\cdots A_{i_{1}}^{t}, where the transposes of AiA_{i}’s are

A1t=(100110101),A2t=(110010011),A3t=(101011001).A_{1}^{t}=\begin{pmatrix}1&0&0\\ 1&1&0\\ 1&0&1\end{pmatrix},\ A_{2}^{t}=\begin{pmatrix}1&1&0\\ 0&1&0\\ 0&1&1\end{pmatrix},\ A_{3}^{t}=\begin{pmatrix}1&0&1\\ 0&1&1\\ 0&0&1\end{pmatrix}.

It is not hard to check that for any γΓ\gamma\in\Gamma, the transpose action γt\gamma^{t} preserves Δ\Delta since the entries of the matrix presentation of γt\gamma^{t} are all non-negative. For any n|γ|n\leqslant|\gamma|, we denote by γnt:=Ai|γ|tAi|γ|n+1t\gamma_{n}^{t}:=A_{i_{|\gamma|}}^{t}\cdots A_{i_{|\gamma|-n+1}}^{t}. Notice that γnt\gamma_{n}^{t} is not equal to (γn)t(\gamma_{n})^{t}, it should be identified as (γt)n(\gamma^{t})_{n}.

5.2 The upper bound of the Hausdorff dimension

The goal of this section is to show the upper bound of the Hausdorff dimension of the Rauzy gasket, that is dimXsA(Γ)\dim X\leqslant s_{\mathrm{A}}(\Gamma). The following elementary lemma in linear algebra is the key observation, which plays a crucial role in estimating the upper bound of dimX.\dim X.

Recall that {e1,e2,e3}\{e_{1},e_{2},e_{3}\} is the standard orthonormal basis of 3{\mathbb{R}}^{3} and Ei=ei(3).E_{i}={\mathbb{R}}e_{i}\in{\mathbb{P}}({\mathbb{R}}^{3}).

Lemma 5.2.

For every nn\in\mathbb{N}, there exists ϵn>0\epsilon_{n}>0 such that for any γΓ\gamma\in\Gamma, if the last nn digits of γ\gamma are not the same, then for i=1,2,3i=1,2,3

γeiϵnσ1(γ).\|\gamma e_{i}\|\geqslant\epsilon_{n}\sigma_{1}(\gamma).
Proof.

The idea to show the lemma is to consider the transpose action of Γ\Gamma. We list some basic facts on the action of AitA_{i}^{t} on Δ.\Delta.

Lemma 5.3.

For every ij{1,2,3},i\neq j\in\left\{1,2,3\right\}, the following holds:

  1. (1)

    AitA_{i}^{t} preserves Δ.\Delta.

  2. (2)

    AitA_{i}^{t} preserves Δ,\Delta^{\prime}, where Δ\Delta^{\prime} is the open projective triangle in (3)\mathbb{P}(\mathbb{R}^{3}) with vertices (011),(101),(110).\begin{pmatrix}0\\ 1\\ 1\end{pmatrix},\begin{pmatrix}1\\ 0\\ 1\end{pmatrix},\begin{pmatrix}1\\ 1\\ 0\end{pmatrix}.

  3. (3)

    AitEj=Ej.A_{i}^{t}E_{j}=E_{j}.

  4. (4)

    AitEi=(111)Δ.A_{i}^{t}E_{i}=\begin{pmatrix}1\\ 1\\ 1\end{pmatrix}\in\Delta^{\prime}.

Combining (2)(3)(4) in the lemma above, we obtain

Lemma 5.4.

For any i{1,2,3}i\in\{1,2,3\}, for any γ=Ai1Ai|γ|Γ\gamma=A_{i_{1}}\cdots A_{i_{|\gamma|}}\in\Gamma, if there exists some ij(γ)=ii_{j}(\gamma)=i, then γtEiΔ.\gamma^{t}E_{i}\in\Delta^{\prime}.

We are back to the proof of Lemma 5.2. Let γΓ\gamma\in\Gamma be an element such that the last nn digits of γ\gamma are not the same. Due to 2.1, we have

γeiγd(Ei,Hγ).\|\gamma e_{i}\|\geqslant\|\gamma\|d(E_{i},H_{\gamma^{-}}). (5.1)

Recall the relation (V(γt)+)=Hγ.(V_{(\gamma^{t})^{+}})^{\perp}=H_{\gamma^{-}}. So it is sufficient to show the angle between EiE_{i} and V(γt)+V_{(\gamma^{t})^{+}} is bounded away from π/2\pi/2. Note that Δ\Delta has a special geometry property: EiE_{i}^{\perp} is the span {Ej:ji}\left\{E_{j}:j\neq i\right\}, which corresponds to an edge of Δ.\Delta. In order to show the angle between V(γt)+V_{(\gamma^{t})^{+}} and EiE_{i} is bounded away from π/2,\pi/2, it suffices to show that d(V(γt)+,Δ)d(V_{(\gamma^{t})^{+}},\partial\Delta) is lower bounded by a positive constant only depending on n.n. To determine the position of V(γt)+,V_{(\gamma^{t})^{+}}, we use the fact that it is the attracting fixed point of γtγ\gamma^{t}\gamma on the projective plane. Hence V(γt)+γtγΔ.V_{(\gamma^{t})^{+}}\in\gamma^{t}\gamma\Delta.

Lemma 5.5.

γtγΔγntΔΔ¯.\gamma^{t}\gamma\Delta\subset\gamma_{n}^{t}\Delta\cap\overline{\Delta^{\prime}}.

Proof.

Since AiA_{i} and AitA_{i}^{t} preserve Δ\Delta for i=1,2,3,i=1,2,3, we obtain γtγΔγntΔ.\gamma^{t}\gamma\Delta\subset\gamma_{n}^{t}\Delta. Now we show γtγΔΔ¯.\gamma^{t}\gamma\Delta\subset\overline{\Delta^{\prime}}. There are two possible cases. If all of A1,A2,A3A_{1},A_{2},A_{3} occur in γ,\gamma, then γtEiΔ\gamma^{t}E_{i}\in\Delta^{\prime} for each i=1,2,3i=1,2,3 by Lemma 5.4. Therefore γtγΔγtΔΔ.\gamma^{t}\gamma\Delta\subset\gamma^{t}\Delta\subset\Delta^{\prime}.

Otherwise, there are at most two of AiA_{i} occur in the γ.\gamma. Recall the assumption that the last nn digits of γ\gamma are not the same. Without loss of generality, we can assume that both A1A_{1} and A2A_{2} occur in the last nn digits of γ.\gamma. Now we consider the region

z{(xyz)Δ:zx+y}.\nabla_{z}\coloneqq\left\{\begin{pmatrix}x\\ y\\ z\end{pmatrix}\in\Delta:z\leqslant x+y\right\}.

Then A1,A2,A1t,A2tA_{1},A_{2},A_{1}^{t},A_{2}^{t} preserve z.\nabla_{z}. Moreover, A1ΔzA_{1}\Delta\subset\nabla_{z} and A2Δz.A_{2}\Delta\subset\nabla_{z}. Since A3A_{3} does not occur in γ,\gamma, we have γtγΔz.\gamma^{t}\gamma\Delta\subset\nabla_{z}. Finally, we notice that the vertices of γntΔ\gamma_{n}^{t}\Delta satisfy γntE1,γntE2Δ\gamma_{n}^{t}E_{1},\gamma_{n}^{t}E_{2}\in\Delta^{\prime} and γntE3=E3\gamma_{n}^{t}E_{3}=E_{3} by Lemma 5.4. This gives γtγΔγntΔz=γntΔΔ¯.\gamma^{t}\gamma\Delta\subset\gamma_{n}^{t}\Delta\cap\nabla_{z}=\gamma_{n}^{t}\Delta\cap\overline{\Delta^{\prime}}.

Refer to caption
Figure 2: The pattern of γntΔΔ¯\gamma^{t}_{n}\Delta\cap\overline{\Delta^{\prime}}.

We complete the proof of Lemma 5.2. By an inductive argument on n,n, we can assume that there are exactly two of {1,2,3}\left\{1,2,3\right\} occur in i|γ|n+1,i|γ|.i_{|\gamma|-n+1},\cdots i_{|\gamma|}. Without loss of generality, we assume these two digits are 11 and 2.2. By Lemma 5.4, the vertices of γntΔ\gamma_{n}^{t}\Delta satisfy

γntE1Δ,γntE2Δ,γntE3=E3.\gamma_{n}^{t}E_{1}\in\Delta^{\prime},\quad\gamma_{n}^{t}E_{2}\in\Delta^{\prime},\quad\gamma_{n}^{t}E_{3}=E_{3}.

Notice that set γntΔΔ¯\gamma_{n}^{t}\Delta\cap\overline{\Delta^{\prime}} is a closed quadrilateral (see Figure 2), which does not intersect the boundary of Δ\Delta. Thus for every such γn\gamma_{n}, we have d(Δ,γntΔΔ¯)>0d(\partial\Delta,\gamma_{n}^{t}\Delta\cap\overline{\Delta^{\prime}})>0. Since there are only finitely many such γn\gamma_{n} for a given positive integer n,n, there exists dn>0d_{n}>0 only depending on nn such that

d(Δ,γntΔΔ¯)>dnd(\partial\Delta,\gamma_{n}^{t}\Delta\cap\overline{\Delta^{\prime}})>d_{n}

for all such γnt\gamma_{n}^{t}. Recalling V(γt)+γtγΔγntΔΔ,V_{(\gamma^{t})^{+}}\in\gamma^{t}\gamma\Delta\subset\gamma_{n}^{t}\Delta\cap\Delta^{\prime}, we have d(V(γt)+,Δ)>dn.d(V_{(\gamma^{t})^{+}},\partial\Delta)>d_{n}.

The following lemma shows some basic estimates of the diameter and the area of Δγ\Delta_{\gamma}.

Lemma 5.6.

There exists C2>1C_{2}>1 such that if the last 22 digits of γ\gamma are not the same, then

  1. (1)

    diam(Δγ)C2σ2(γ)σ1(γ)\operatorname{\mathrm{diam}}(\Delta_{\gamma})\leqslant C_{2}\cdot\frac{\sigma_{2}(\gamma)}{\sigma_{1}(\gamma)};

  2. (2)

    Area(Δγ)C2σ1(γ)3\mathrm{Area}(\Delta_{\gamma})\leqslant C_{2}\cdot\sigma_{1}(\gamma)^{-3}.

Proof.
  1. (1)

    It suffices to show d(γEi,γEj)C2σ2(γ)σ1(γ)d(\gamma E_{i},\gamma E_{j})\leqslant C_{2}\cdot\frac{\sigma_{2}(\gamma)}{\sigma_{1}(\gamma)}. We have

    d(γEi,γEj)\displaystyle d(\gamma E_{i},\gamma E_{j}) =\displaystyle= γeiγejγeiγejσ1(γ)σ2(γ)eiejγeiγej\displaystyle\frac{\|\gamma e_{i}\wedge\gamma e_{j}\|}{\|\gamma e_{i}\|\cdot\|\gamma e_{j}\|}\leqslant\frac{\sigma_{1}(\gamma)\sigma_{2}(\gamma)\|e_{i}\wedge e_{j}\|}{\|\gamma e_{i}\|\cdot\|\gamma e_{j}\|}
    \displaystyle\leqslant σ1(γ)σ2(γ)ϵ22σ1(γ)2 (by Lemma 5.2)\displaystyle\frac{\sigma_{1}(\gamma)\sigma_{2}(\gamma)}{\epsilon_{2}^{2}\sigma_{1}(\gamma)^{2}}\text{ (by Lemma \ref{lem: crucial proj geo lem}) }
    =\displaystyle= ϵ22σ2(γ)σ1(γ).\displaystyle\epsilon_{2}^{-2}\cdot\frac{\sigma_{2}(\gamma)}{\sigma_{1}(\gamma)}.
  2. (2)

    We use the following elementary geometric fact: Let x,y,zx,y,z be three points in 3{o}\mathbb{R}^{3}\setminus\{o\}, then the area of the triangle formed by x,y,zx,y,z (which we denote by (xyz)(xyz)) is equal to

    Area(xyz)=xyz2dE(o,(xyz)),\mathrm{Area}(xyz)=\frac{\|x\wedge y\wedge z\|}{2d_{E}(o,(xyz))},

    where dE(o,(xyz))d_{E}(o,(xyz)) is the distance from the origin oo to the two plane of (xyz)(xyz). This is because the numerator gives the volume of the polyhedron of oxyzoxyz.

    For γΓ\gamma\in\Gamma, recall that the area of Δγ\Delta_{\gamma} is understood in the area of the corresponding subset of Δ~\widetilde{\Delta} in the euclidean space. Let xγ,yγ,zγx_{\gamma},y_{\gamma},z_{\gamma} be the corresponding points in Δ~\widetilde{\Delta} of vertices of Δγ.\Delta_{\gamma}. Then

    Area(xγyγzγ)=xγyγzγ2d(o,(xγyγzγ))=xγyγzγ2d(o,Δ~).\mathrm{Area}(x_{\gamma}y_{\gamma}z_{\gamma})=\frac{\|x_{\gamma}\wedge y_{\gamma}\wedge z_{\gamma}\|}{2d(o,(x_{\gamma}y_{\gamma}z_{\gamma}))}=\frac{\|x_{\gamma}\wedge y_{\gamma}\wedge z_{\gamma}\|}{2d(o,\widetilde{\Delta})}.

    We know that

    xγ=γe1/ω(γe1),x_{\gamma}=\gamma e_{1}/\omega(\gamma e_{1}),

    with ω(v)=v1+v2+v3\omega(v)=v_{1}+v_{2}+v_{3}, similarily for yγ,zγy_{\gamma},z_{\gamma}. Therefore,

    xγyγzγ=γe1γe2γe3/1i3ω(γei)\|x_{\gamma}\wedge y_{\gamma}\wedge z_{\gamma}\|=\|\gamma e_{1}\wedge\gamma e_{2}\wedge\gamma e_{3}\|/\prod_{1\leqslant i\leqslant 3}\omega(\gamma e_{i})

    We actually have γe1γe2γe3=e1e2e3=1\|\gamma e_{1}\wedge\gamma e_{2}\wedge\gamma e_{3}\|=\|e_{1}\wedge e_{2}\wedge e_{3}\|=1. Due to γEiΔ\gamma E_{i}\in\Delta, we also have

    ω(γei)γei\omega(\gamma e_{i})\geqslant\|\gamma e_{i}\|

    Then by Lemma 5.2 we get the proof of (2).∎

The following geometric lemma is essentially proved in [PS21, Lemma 4.1].

Lemma 5.7.

For any δ>0\delta>0, there exists cδ>0c_{\delta}>0 such that for any γΓ\gamma\in\Gamma, there exists a finite open cover {Di(γ):i=1,,k}\{D_{i}(\gamma):i=1,\cdots,k\} of Δγ\Delta_{\gamma} with diamDi(γ)diamΔγ\operatorname{\mathrm{diam}}D_{i}(\gamma)\leqslant\operatorname{\mathrm{diam}}\Delta_{\gamma} such that

idiam1+δDi(γ)cδdiam1δΔγAreaδΔγ.\sum_{i}\operatorname{\mathrm{diam}}^{1+\delta}D_{i}(\gamma)\leqslant c_{\delta}\cdot\operatorname{\mathrm{diam}}^{1-\delta}\Delta_{\gamma}\cdot\mathrm{Area}^{\delta}\Delta_{\gamma}.
Proof.

As in the proof of [PS21, Lemma 4.1], we know that every Δγ\Delta_{\gamma} can be covered by O(diam2(Δγ)/Area(Δγ))O(\operatorname{\mathrm{diam}}^{2}(\Delta_{\gamma})/\mathrm{Area}(\Delta_{\gamma})) disks {Di:i=1,,k}\{D_{i}:i=1,\cdots,k\} of diameter O(Area(Δγ)/diam(Δγ))O(\mathrm{Area}(\Delta_{\gamma})/\operatorname{\mathrm{diam}}(\Delta_{\gamma})), then we get the proof. ∎

We back to the proof of dimH(X)sA(Γ)\dim_{\mathrm{H}}(X)\leqslant s_{\mathrm{A}}(\Gamma). Recall PΓ(s)=γΓφs(γ)P_{\Gamma}(s)=\sum_{\gamma\in\Gamma}\varphi_{s}(\gamma) is the Poincaré series of Γ\Gamma, where φs(γ)\varphi_{s}(\gamma) is defined by

φs(γ)={(σ2σ1)s(γ),0<s1;(σ2σ1)(γ)(σ3σ1)s1(γ),1<s2.\varphi_{s}(\gamma)=\left\{\begin{aligned} &\Big{(}\frac{\sigma_{2}}{\sigma_{1}}\Big{)}^{s}(\gamma),&0<s\leqslant 1;\\ &\Big{(}\frac{\sigma_{2}}{\sigma_{1}}\Big{)}(\gamma)\Big{(}\frac{\sigma_{3}}{\sigma_{1}}\Big{)}^{s-1}(\gamma),&1<s\leqslant 2.\end{aligned}\right.

By definition of sAs_{\mathrm{A}}, it suffices to show if PΓ(s)<P_{\Gamma}(s)<\infty then for any ϵ>0\epsilon>0, dimXs+ϵ\dim X\leqslant s+\epsilon.

Definition 5.8.

For xXx\in X, we say xx is nice if every (i1(x),i2(x),)Φ1(x)Λ(i_{1}(x),i_{2}(x),\cdots)\in\Phi^{-1}(x)\in\Lambda^{\mathbb{N}} is not ending by a single element in Λ={1,2,3}\Lambda=\{1,2,3\}.

We remark that if xx is nice, then xx is uniquely coding. This is because the only possibility of Φ(𝐢)=Φ(𝐢)\Phi(\mathbf{i})=\Phi(\mathbf{i}^{\prime}) is

𝐢=(w,j1,j2,j2,) and 𝐢=(w,j2,j1,j1,),\mathbf{i}=(w,j_{1},j_{2},j_{2},\cdots)\text{ and }\mathbf{i}^{\prime}=(w,j_{2},j_{1},j_{1},\cdots),

where wΛNw\in\Lambda^{N} for some NN\in{\mathbb{N}} and j1j2Λ.j_{1}\neq j_{2}\in\Lambda.

Then the Rauzy Gasket XX can be decomposed as the set of nice points, which we denote it by XniceX_{\mathrm{nice}}, and a countable set. It suffices to show the Hausdorff outer measure Hs+ϵ(Xnice)H^{s+\epsilon}(X_{\mathrm{nice}}) is 0.

We consider

Γm{γΓ:the last 2 digits of γ are not the same and diamΔγ1/m}.\Gamma_{m}\coloneqq\left\{\gamma\in\Gamma:\text{the last $2$ digits of $\gamma$ are not the same and }\operatorname{\mathrm{diam}}\Delta_{\gamma}\leqslant 1/m\right\}.

Now we construct two families of covers 𝒰m\mathcal{U}_{m} and 𝒰m\mathcal{U}_{m}^{\prime} for mm\in\mathbb{N} as

𝒰m{Di(γ):γΓm},𝒰m{Δγ:γΓm},\mathcal{U}_{m}\coloneqq\left\{D_{i}(\gamma):\gamma\in\Gamma_{m}\right\},\quad\mathcal{U}_{m}^{\prime}\coloneqq\left\{\Delta_{\gamma}:\gamma\in\Gamma_{m}\right\},

where {Di(γ)}\{D_{i}(\gamma)\} is the finite open cover of Δγ\Delta_{\gamma} we obtained from Lemma 5.7. Let

Ym=1U𝒰mUandYm=1U𝒰mU.Y\coloneqq\bigcap_{m=1}^{\infty}\bigcup_{U\in\mathcal{U}_{m}}U\quad\text{and}\quad Y^{\prime}\coloneqq\bigcap_{m=1}^{\infty}\bigcup_{U\in\mathcal{U}^{\prime}_{m}}U.

Then the sequence of covers 𝒰m\mathcal{U}_{m} (resp. 𝒰m\mathcal{U}_{m}^{\prime}) is a family of Vitali covers of the set YY (resp. YY^{\prime}).333Recall that a Vitali cover 𝒱\mathcal{V} of a set EE is a family of sets so that, for every δ>0\delta>0 and every xEx\in E, there is some UU in the family 𝒱\mathcal{V} with diamU<δ\operatorname{\mathrm{diam}}U<\delta and xUx\in U. Notice that YY contains YY^{\prime} by the construction of Di(γ)D_{i}(\gamma)’s.

Lemma 5.9.

YXniceY\supset X_{\mathrm{nice}}.

Proof.

We claim that for any 𝐢{1,2,3}\mathbf{i}\in\{1,2,3\}^{\mathbb{N}} which is not ending by a single element, the point x=Φ(𝐢)x=\Phi(\mathbf{i}) is contained in YY. Since 𝐢=(i1(x),)\mathbf{i}=(i_{1}(x),\cdots) is not ending by a single element in {1,2,3}\{1,2,3\}, there exists infinitely many \ell such that i1(x),i(x)i_{\ell-1}(x),i_{\ell}(x) are not the same (otherwise it will be ended by only one element). Collect all such \ell and consider all the elements γ=Ai1Ai\gamma=A_{i_{1}}\cdots A_{i_{\ell}}. Then we get there are infinitely many γΓ\gamma\in\Gamma such that the last two digits of γ\gamma are not the same and xΔγx\in\Delta_{\gamma}. By 5.1, diamΔγ0\operatorname{\mathrm{diam}}\Delta_{\gamma}\to 0 as \ell tending to infinity. Therefore xYYx\in Y^{\prime}\subset Y. ∎

As a consequence, it suffices to show the Hausdorff outer measure

Hs+ϵ(Y)=limδ0Hδs+ϵ(Y)=0.H^{s+\epsilon}(Y)=\lim_{\delta\to 0}H^{s+\epsilon}_{\delta}(Y)=0.

Recall that diamU1/m\operatorname{\mathrm{diam}}U\leqslant 1/m for every U𝒰m.U\in\mathcal{U}_{m}. Then for s1s\geqslant 1, we have

Hs+ϵ(Y)\displaystyle H^{s+\epsilon}(Y) \displaystyle\leqslant lim supδ0Hδs+ϵ(Y)lim supmU𝒰mdiam(U)s+ϵ\displaystyle\limsup_{\delta\to 0}H^{s+\epsilon}_{\delta}(Y)~{}\leqslant~{}\limsup_{m\to\infty}\sum_{U\in\mathcal{U}_{m}}\operatorname{\mathrm{diam}}(U)^{s+\epsilon}
=\displaystyle= lim supmγΓmidiam(Di(γ))s+ϵ\displaystyle\limsup_{m\to\infty}\sum_{\gamma\in\Gamma_{m}}\sum_{i}\operatorname{\mathrm{diam}}(D_{i}(\gamma))^{s+\epsilon}
\displaystyle\leqslant lim supmcs+ϵ1γΓmdiam2sϵΔγAreas+ϵ1Δγ(by Lemma 5.7)\displaystyle\limsup_{m\to\infty}c_{s+\epsilon-1}\sum_{\gamma\in\Gamma_{m}}\operatorname{\mathrm{diam}}^{2-s-\epsilon}\Delta_{\gamma}\cdot\mathrm{Area}^{s+\epsilon-1}\Delta_{\gamma}\qquad(\text{by Lemma \ref{lem: 4.1 PS}})
\displaystyle\leqslant lim supmcs+ϵ1C2γΓmφs+ϵ(γ) (by Lemma 5.6 and the definition of φs).\displaystyle\limsup_{m\to\infty}c_{s+\epsilon-1}C_{2}\sum_{\gamma\in\Gamma_{m}}\varphi_{s+\epsilon}(\gamma)\text{ \quad(by Lemma \ref{lem: proj geo lem} and the definition of $\varphi_{s}$}).

Since φs+ϵ(γ)<\sum\varphi_{s+\epsilon}(\gamma)<\infty, the right-hand side of the last inequality is 0 when mm\to\infty, which is exactly what we want. For the case s<1s<1 (we actually know s1.72s\sim 1.72 by numerical test), we do not need to use Lemma 5.7 and consider YY^{\prime} instead of YY by the same argument. We still get the proof.

5.3 The lower bound of the Hausdorff dimension

In this section, we will show that dimXsA(Γ).\dim X\geqslant s_{\mathrm{A}}(\Gamma). Note that XX contains the boundary of Δ,\Delta, which has the Hausdorff dimension equal to 1.1. Therefore, we already have an a priori estimate that dimX1\dim X\geqslant 1, which allows us to assume, without loss of generality, that sA(Γ)>1s_{\mathrm{A}}(\Gamma)>1. In other words, we assume the existence of a value s>1s>1 for which the series γφs(γ)\sum_{\gamma}\varphi_{s}(\gamma) diverges. This technical assumption simplifies subsequent discussions. Recall that for every exact dimensional probability measure μ\mu supported on X,X, we have dimμdimX.\dim\mu\leqslant\dim X. The inequality dimXsA(Γ)\dim X\geqslant s_{\mathrm{A}}(\Gamma) is a direct consequence of the following variational principle of the affinity exponent.

Lemma 5.10.

Let s>1s>1 such that the series γΓφs(γ)\sum_{\gamma\in\Gamma}\varphi_{s}(\gamma) diverges, then for every ϵ>0,\epsilon>0, there exists a finitely supported measure ν\nu supported on Γ\Gamma and a ν\nu-stationary measure μ\mu supported on XX satisfying dimμsϵ\dim\mu\geqslant s-\epsilon.

Our idea is a stopping time argument which is partially inspired by the study of the variational principle of iterated function systems [FH09] and self-affine measures [MS23]. Specifically, our strategy is to combine our dimension formula of stationary measures with a modification of the proof of Theorem 3.1 in [MS23]. The main difference of our proof from [MS23] is that due to the presence of unipotent elements, the word lengths of elements may lose control. So instead of considering the word length, we only consider κ(γ).\kappa(\gamma). This is enough to estimate the Lyapunov exponents. Then we use a combinatorial argument to make elements generating a free semigroup. This generates a large enough random walk entropy. Therefore, we can find a stationary measure with a large dimension approximating the affinity exponent. A similar argument to address the issue of words with uncontrolled length also appears in [HJX23].

In order to apply the dimension formula (Theorem 1.2), we need to show the Zariski density of Γ\Gamma at first.

Lemma 5.11.

The semigroup Γ\Gamma is Zariski dense in SL3()\mathrm{SL}_{3}(\mathbb{R}).

Proof.

Let 𝐇{\mathbf{H}} be the Zariski closure of semigroup Γ\Gamma in SL3()\mathrm{SL}_{3}(\mathbb{R}), which is a real algebraic group. Let 𝔥\mathfrak{h} be the Lie algebra of the connected component of 𝐇{\mathbf{H}}. By [Bor91, Page 106, Remark], the fact that A1A_{1} is unipotent implies that the one-parameter unipotent group {s,(1ss010001)}\left\{s\in\mathbb{R},\ \begin{pmatrix}1&s&s\\ 0&1&0\\ 0&0&1\end{pmatrix}\right\} is in 𝐇{\mathbf{H}} and X1=(011000000)X_{1}=\begin{pmatrix}0&1&1\\ 0&0&0\\ 0&0&0\end{pmatrix} is in 𝔥\mathfrak{h}. Similarly, the nilpotent elements X2=(000101000)X_{2}=\begin{pmatrix}0&0&0\\ 1&0&1\\ 0&0&0\end{pmatrix}, X3=(000000110)X_{3}=\begin{pmatrix}0&0&0\\ 0&0&0\\ 1&1&0\end{pmatrix} are also in 𝔥\mathfrak{h}. Then we can play with these elements in the Lie algebra 𝔥\mathfrak{h} and prove that they generate all 𝔰𝔩3\mathfrak{sl}_{3}. We have Y3=[X1,X2]=(101011000)Y_{3}=[X_{1},X_{2}]=\begin{pmatrix}1&0&1\\ 0&-1&-1\\ 0&0&0\end{pmatrix}, Y1=[X2,X3]=(000110101)Y_{1}=[X_{2},X_{3}]=\begin{pmatrix}0&0&0\\ 1&1&0\\ -1&0&-1\end{pmatrix}, Y2=[X3,X1]=(110000011)Y_{2}=[X_{3},X_{1}]=\begin{pmatrix}-1&-1&0\\ 0&0&0\\ 0&1&1\end{pmatrix} also in 𝔥\mathfrak{h}. Then [Y1,X1]+Y2+Y3+2X1=(004000000)[Y_{1},X_{1}]+Y_{2}+Y_{3}+2X_{1}=\begin{pmatrix}0&0&4\\ 0&0&0\\ 0&0&0\end{pmatrix}, [Y2,X2]+Y3+Y1+2X2=(000400000)[Y_{2},X_{2}]+Y_{3}+Y_{1}+2X_{2}=\begin{pmatrix}0&0&0\\ 4&0&0\\ 0&0&0\end{pmatrix} also in 𝔥\mathfrak{h}. From these elements, we can obtain the whole 𝔰𝔩(3,)\mathfrak{sl}(3,\mathbb{R}) and then 𝐇{\mathbf{H}} must be the whole group SL3()\mathrm{SL}_{3}(\mathbb{R}). ∎

Recall that for s>1,s>1, the linear function ψs\psi_{s} on 𝔞\mathfrak{a} is given by

ψs(λ)=(λ1λ2)+(s1)(λ1λ3),λ𝔞.\psi_{s}(\lambda)=(\lambda_{1}-\lambda_{2})+(s-1)(\lambda_{1}-\lambda_{3}),\quad\forall\lambda\in\mathfrak{a}.

Now we construct a good set coming from the divergent series γφs(γ)=γexp(ψs(κ(γ))).\sum_{\gamma}\varphi_{s}(\gamma)=\sum_{\gamma}\exp(-\psi_{s}(\kappa(\gamma))).

Definition 5.12.

For β>0,x𝔞+\beta>0,x\in\mathfrak{a}^{+} and n,n\in{\mathbb{N}}, a subset SΓS\subset\Gamma is called (n,β,x)(n,\beta,x)-approximate if #Se(1β)n\#S\geqslant e^{(1-\beta)n} and for every γS,\gamma\in S, 1nκ(γ)xβ.\|\frac{1}{n}\kappa(\gamma)-x\|\leqslant\beta.

Lemma 5.13.

For any β>0\beta>0, there exists x𝔞+x\in\mathfrak{a}^{+} satisfying |ψs(x)1|β|\psi_{s}(x)-1|\leqslant\beta and infinitely many nn\in{\mathbb{N}} such that there exists an (n,β,x)(n,\beta,x)-approximate subset SΓ.S\subset\Gamma.

Proof.

Note that ψs(κ(γ))(s1)logσ1(γ)=(s1)logγ.\psi_{s}(\kappa(\gamma))\geqslant(s-1)\log\sigma_{1}(\gamma)=(s-1)\log\left\|\gamma\right\|. Hence for every t>0,t>0, the set {γΓ:ψs(κ(γ))<t}\left\{\gamma\in\Gamma:\psi_{s}(\kappa(\gamma))<t\right\} is finite by the discreteness of Γ.\Gamma. Combining with the hypothesis that the series γΓexp(ψs(κ(γ)))\sum_{\gamma\in\Gamma}\exp(-\psi_{s}(\kappa(\gamma))) diverges, there are infinitely many nn such that

{γ:ψs(κ(γ))[(1β/10)n,(1+β/10)n]}\left\{\gamma:\psi_{s}(\kappa(\gamma))\in[(1-\beta/10)n,(1+\beta/10)n]\right\}

contains at least e(1β/10)ne^{(1-\beta/10)n} elements.

Since we have ψs(κ(γ))(s1)logσ1(γ)\psi_{s}(\kappa(\gamma))\geqslant(s-1)\log\sigma_{1}(\gamma) and ψs(κ(γ))(s1)logσ3(γ),\psi_{s}(\kappa(\gamma))\geqslant-(s-1)\log\sigma_{3}(\gamma), the set ψs1([1β/10,1+β/10])𝔞+\psi_{s}^{-1}([1-\beta/10,1+\beta/10])\cap\mathfrak{a}^{+} is compact. Therefore we can cover it by finitely many balls of radius β/100.\beta/100. By a pigeonhole principle, there exists a center of some ball, say x𝔞+,x\in\mathfrak{a}^{+}, such that for infinitely many n,n,

{γ:|ψs(κ(γ))n1|β10,κ(γ)nxβ100}\left\{\gamma:|\frac{\psi_{s}(\kappa(\gamma))}{n}-1|\leqslant\frac{\beta}{10},~{}\|\frac{\kappa(\gamma)}{n}-x\|\leqslant\frac{\beta}{100}\right\}

contains at least e(1β)ne^{(1-\beta)n} elements. Using the fact that |ψs(κ(γ))|10κ(γ),|\psi_{s}(\kappa(\gamma))|\leqslant 10\|\kappa(\gamma)\|, we obtain |ψs(x)1|β,|\psi_{s}(x)-1|\leqslant\beta, which is exactly what we want. ∎

We will modify an (n,β,x)(n,\beta,x)-approximate subset to a set which freely generates a free semigroup. Since Γ\Gamma is itself a free semigroup, a subset of Γ\Gamma can freely generate a free semigroup by avoiding the “prefix relations”. We consider the following concepts.

Definition 5.14.
  1. (1)

    An element j1Γj_{1}\in\Gamma is called starting with j2Γj_{2}\in\Gamma if there is j3Γ{id}j_{3}\in\Gamma\setminus\{\mathrm{id}\} such that j1=j2j3j_{1}=j_{2}j_{3}. We also say j2j_{2} is a prefix of j1j_{1} if j1j_{1} is starting with j2.j_{2}.

  2. (2)

    An element j1Γj_{1}\in\Gamma is called ending with j2Γ{id}j_{2}\in\Gamma\setminus\{\mathrm{id}\} if there is j3Γj_{3}\in\Gamma such that j1=j3j2j_{1}=j_{3}j_{2}.

  3. (3)

    An element jj is called minimal in a subset SS of Γ\Gamma if there is no element jSj^{\prime}\in S such that jj is starting with jj^{\prime}.

Within a set SS, a minimal element of SS is never a prefix of another minimal element. So the set of minimal elements SminS_{\min} of SS will freely generate a free semigroup. Moreover, the subset SminΓS_{\min}^{*\ell}\subset\Gamma satisfies that there is no pair of elements such that one is a prefix of the other. In the following lemma, using hyperbolicity and discreteness, we get a lower bound of minimal elements in a set with approximately the same sizes of Cartan projections.

Lemma 5.15.

There exists C>0C>0 such that the following hold. For every β>0,\beta>0, nn\in{\mathbb{N}} and x𝔞+.x\in\mathfrak{a}^{+}. For any element jj in the set

𝒥={γΓ:κ(γ)nxβ,γ is ending with A1A2A3},\mathcal{J}=\{\gamma\in\Gamma:\ \|\frac{\kappa(\gamma)}{n}-x\|\leqslant\beta,~{}~{}\gamma\text{ is ending with }A_{1}A_{2}A_{3}\},

there are at most eCβne^{C\beta n} elements in 𝒥\mathcal{J} which is starting with j.j.

Proof.

For every VΔV\in\Delta, let v=ivieiVv=\sum_{i}v_{i}e_{i}\in V be a unit vector with v1,v2,v30.v_{1},v_{2},v_{3}\geqslant 0. Since all the entries of γei\gamma e_{i} are non-negative, we have

γvmaxi{viγei}12mini{γei}.\|\gamma v\|\geqslant\max_{i}\{v_{i}\|\gamma e_{i}\|\}\geqslant\frac{1}{2}\min_{i}\{\|\gamma e_{i}\|\}.

Take n=2n=2 in 5.2, for any element γ\gamma with last two digits different, we have γeiϵ2σ1(γ)\|\gamma e_{i}\|\geqslant\epsilon_{2}\sigma_{1}(\gamma), where ϵ2\epsilon_{2} is defined in 5.2. Hence

γvϵ22σ1(γ).\|\gamma v\|\geqslant\frac{\epsilon_{2}}{2}\sigma_{1}(\gamma).

Then for any j,jj′′𝒥j,jj^{\prime\prime}\in\mathcal{J} with j′′Γ{id}j^{\prime\prime}\in\Gamma\setminus\{\mathrm{id}\} and any unit vVΔv\in V\in\Delta, since j′′VΔj^{\prime\prime}V\in\Delta and j,j′′j,j^{\prime\prime} end with A1A2A3A_{1}A_{2}A_{3}, we obtain

σ1(jj′′)jj′′v=j(j′′v)j′′vj′′vvϵ224σ1(j)σ1(j′′).\sigma_{1}(jj^{\prime\prime})\geqslant\|jj^{\prime\prime}v\|=\frac{\|j(j^{\prime\prime}v)\|}{\|j^{\prime\prime}v\|}\frac{\|j^{\prime\prime}v\|}{\|v\|}\geqslant\frac{\epsilon_{2}^{2}}{4}\sigma_{1}(j)\sigma_{1}(j^{\prime\prime}). (5.2)

Since 1nκ(jj′′)xβ\|\frac{1}{n}\kappa(jj^{\prime\prime})-x\|\leqslant\beta and 1nκ(j)xβ\|\frac{1}{n}\kappa(j)-x\|\leqslant\beta, we have

logσ1(jj′′)2βn+logσ1(j).\log\sigma_{1}(jj^{\prime\prime})\leqslant 2\beta n+\log\sigma_{1}(j). (5.3)

Combining (5.2),(5.3) we get

σ1(j′′)4e2βn/ϵ22.\sigma_{1}(j^{\prime\prime})\leqslant 4e^{2\beta n}/\epsilon_{2}^{2}.

Since the semigroup Γ\Gamma is discrete in SL3()\mathrm{SL}_{3}(\mathbb{R}), up to a constant the number

#{γΓ,γt}\#\{\gamma\in\Gamma,\ \|\gamma\|\leqslant t\}

is bounded by the volume of the set {gSL3(),gt}\{g\in\mathrm{SL}_{3}(\mathbb{R}),\ \|g\|\leqslant t\} , which grows at most polynomially on tt. Then the possible number of j′′j^{\prime\prime} is bounded by qCβnq^{C\beta n} for some constant C>0.C>0.

In order to estimate the Lyapunov exponents of the constructing measure, we need to estimate the Cartan projection of products. Let us recall some notions.

Definition 5.16.
  1. (1)

    For an element gSL3()g\in\mathrm{SL}_{3}(\mathbb{R}), we call it (r,ϵ)(r,\epsilon)-loxodromic for r,ϵ>0r,\epsilon>0, if σi(g)/σi+1(g)1/ϵ\sigma_{i}(g)/\sigma_{i+1}(g)\geqslant 1/\epsilon and

    d(Vg+,Hg)>r,d(Vg,23+,Hg,23)>r,d(V_{g}^{+},H^{-}_{g})>r,\ d(V_{g,\wedge^{2}\mathbb{R}^{3}}^{+},H^{-}_{g,\wedge^{2}\mathbb{R}^{3}})>r,

    where 23\wedge^{2}\mathbb{R}^{3} is the wedge representation of SL3()\mathrm{SL}_{3}(\mathbb{R}) and Vg,23+=k~g(E1E2)V^{+}_{g,\wedge^{2}\mathbb{R}^{3}}=\widetilde{k}_{g}(E_{1}\wedge E_{2}) and Hg,23=kg1(E2E3E3E1)H^{-}_{g,\wedge^{2}\mathbb{R}^{3}}=k_{g}^{-1}(E_{2}\wedge E_{3}\oplus E_{3}\wedge E_{1}) are the corresponding attracting point and repelling hyperplane in the projective space (23),\mathbb{P}(\wedge^{2}\mathbb{R}^{3}), where g=k~gagkgKA+Kg=\widetilde{k}_{g}a_{g}k_{g}\in KA^{+}K is the Cartan decomposition.

  2. (2)

    Let FF be a subset in SL3()\mathrm{SL}_{3}(\mathbb{R}). We call FF a (r,ϵ)(r,\epsilon)-Schottky family if every element in FF is (r,ϵ)(r,\epsilon)-loxodromic and for any pair (g,h)(g,h) in FF, we have

    d(Vg+,Hh)>6r,d(Vg,23+,Hh,23)>6r.d(V_{g}^{+},H^{-}_{h})>6r,\ d(V_{g,\wedge^{2}\mathbb{R}^{3}}^{+},H^{-}_{h,\wedge^{2}\mathbb{R}^{3}})>6r.
  3. (3)

    A set FF of SL3()\mathrm{SL}_{3}(\mathbb{R}) is called η\eta-narrow if for gg in FF, the attracting points Vg+V_{g}^{+} (resp. Vg,23+V_{g,\wedge^{2}\mathbb{R}^{3}}^{+}) are within η\eta-distance of one another and the repelling hyperplanes HgH_{g}^{-} (resp. Hg,23H_{g,\wedge^{2}\mathbb{R}^{3}}^{-}) are within η\eta Hausdorff distance of one another.

  4. (4)

    A set FF is η\eta-narrow around hh if for gg in FF, the attracting points Vg+V_{g}^{+} (resp. Vg,23+V_{g,\wedge^{2}\mathbb{R}^{3}}^{+}) are within η\eta-distance to Vh+V_{h}^{+}(resp. Vh,23+V_{h,\wedge^{2}\mathbb{R}^{3}}^{+}) and the repelling hyperplanes HgH_{g}^{-} (resp. Hg,23H_{g,\wedge^{2}\mathbb{R}^{3}}^{-}) are within η\eta Hausdorff distance to HhH_{h}^{-} (resp. Hh,23H_{h,\wedge^{2}\mathbb{R}^{3}}^{-}).

These definitions and properties are originally due to Benoist [Ben97]. We borrow them from [MS23, Corollary 2.16, 2.17]

Lemma 5.17.
  1. (1)

    For r>4ϵ>0r>4\epsilon>0, if FF is a (r,ϵ)(r,\epsilon)-Schottky family, then the semigroup generated by FF is a (r/2,2ϵ)(r/2,2\epsilon)-Schottky family.

  2. (2)

    Let EE be a η\eta-narrow collection of (r,ϵ)(r,\epsilon)-loxodromic elements with r>4max{ϵ,η}r>4\max\{\epsilon,\eta\}. Then, EE is a (r/4,ϵ)(r/4,\epsilon)-Schottky family.

  3. (3)

    If FF is a (r,ϵ)(r,\epsilon)-Schottky family, then there exists Cr>0C_{r}>0 only depending on rr such that for any g1,,gg_{1},\cdots,g_{\ell} in FF, we have

    κ(g1g)1iκ(gi)Cr.\|\kappa(g_{1}\cdots g_{\ell})-\sum_{1\leqslant i\leqslant\ell}\kappa(g_{i})\|\leqslant\ell\cdot C_{r}.

Now we state the main construction, which gives a good set to support a desired random walk.

Lemma 5.18.

For every β>0.\beta>0. There exists NN\in\mathbb{N}, x𝔞+,𝒥Γ,x\in\mathfrak{a}^{+},\mathcal{J}\subset\Gamma, such that

  1. (1)

    |ψs(x)1|β|\psi_{s}(x)-1|\leqslant\beta.

  2. (2)

    The semigroup generated by 𝒥\mathcal{J} is Zariski dense.

  3. (3)

    For every kk\in\mathbb{N} and j𝒥k:={j1jk:ji𝒥}j\in\mathcal{J}^{*k}:=\{j_{1}\cdots j_{k}:j_{i}\in\mathcal{J}\}, 1kNκ(j)x10β\|\frac{1}{kN}\kappa(j)-x\|\leqslant 10\beta.

  4. (4)

    The set 𝒥\mathcal{J} contains a subset 𝒥1\mathcal{J}_{1} satisfies #𝒥1e(110Cβ)N\#\mathcal{J}_{1}\geqslant e^{(1-10C\beta)N} and no element in 𝒥1\mathcal{J}_{1} is a prefix of another one, where C>0C>0 is the constant given by Lemma 5.15.

Proof.

Fix an x𝔞+x\in\mathfrak{a}^{+} be given by Lemma 5.13. Let SS be a (n,β,x)(n,\beta,x)-approximate subset for some sufficiently large n.n. Now we construct the set 𝒥\mathcal{J} by modifying SS.

Step 1.

For every γS\gamma\in S, we add A1A2A3A_{1}A_{2}A_{3} at the end and denote the new set by W2W_{2}. Since κ(γA1A2A3)κ(γ)κ(A1A2A3)\|\kappa(\gamma A_{1}A_{2}A_{3})-\kappa(\gamma)\|\leqslant\|\kappa(A_{1}A_{2}A_{3})\|, the set W2W_{2} is (n,2β,x)(n,2\beta,x)-approximate for nn large enough.

Step 2.

We apply a theorem of Abels-Margulis-Soifer [AMS95], see also [MS23, Theorem 3.2].

Theorem 5.19 (Abels-Margulis-Soifer).

Let GG be a Zariski-connected real reductive group and Γ\Gamma be a Zariski-dense subsemigroup. Then there exists 0<r=r(Γ)0<r=r(\Gamma) such that for all 0<ϵr0<\epsilon\leqslant r, there exists a finite subset F=F(r,ϵ,Γ)ΓF=F(r,\epsilon,\Gamma)\subset\Gamma with the property that for every gGg\in G, there exists fFf\in F such that fgfg is (r,ϵ)(r,\epsilon)-loxodromic in GG.

We fix r0,ϵ0>0r_{0},\epsilon_{0}>0 sufficiently small comparing to β\beta with r0>100ϵ0.r_{0}>100\epsilon_{0}. By 5.19, we could find a finite subset F1=F(r0,ϵ0,Γ)F_{1}=F(r_{0},\epsilon_{0},\Gamma) of Γ.\Gamma. Therefore for every element γW2\gamma\in W_{2}, there exists fF1f\in F_{1} such that fγf\gamma is (r0,ϵ0)(r_{0},\epsilon_{0})-loxodromic. By the pigeonhole principle, we can find an fF1f\in F_{1} such that for at least (#F1)1(\#F_{1})^{-1} proportion of γ\gamma in W2W_{2}, the product fγf\gamma is (r0,ϵ0)(r_{0},\epsilon_{0})-loxodromic. Fix this fF1f\in F_{1} and let

W3={fγ:γW2,fγ is (r0,ϵ0)-loxodromic }.W_{3}=\left\{f\gamma:\gamma\in W_{2},f\gamma\text{ is $(r_{0},\epsilon_{0})$-loxodromic }\right\}.

Then W3W_{3} is (n,3β,x)(n,3\beta,x)-approximate assuming nn large enough.

Step 3.

By compactness, we can cover 1i2(Vi)×(Vi)\prod_{1\leqslant i\leqslant 2}\mathbb{P}(V_{i})\times\mathbb{P}(V_{i}^{*}) with O(ϵ08)O(\epsilon_{0}^{-8}) balls of radius ϵ0\epsilon_{0}, where V1=3V_{1}=\mathbb{R}^{3} and V2=23V_{2}=\wedge^{2}\mathbb{R}^{3}. By the pigeonhole principle, there exists a subset W4W3W_{4}\subset W_{3}, such that #W4ϵ08#W3\#W_{4}\gg\epsilon_{0}^{8}\cdot\#W_{3} and W4W_{4} is an ϵ0\epsilon_{0}-narrow set of (r0,ϵ0)(r_{0},\epsilon_{0})-loxodromic elements. For nn sufficiently large comparing to ϵ0,\epsilon_{0}, W4W_{4} is (n,4β,x)(n,4\beta,x)-approximate.

Step 4.

Before making the next modification, we recall the following lemma from [Ben97] and [MS23].

Lemma 5.20 ([MS23, Lemma 3.4]).

There exists r1>0r_{1}>0 depending only on Γ\Gamma such that the following hold. For every loxodromic element gGg\in G and ϵ>0,η>0\epsilon>0,\eta>0 there exists a Zariski dense (r1,ϵ)(r_{1},\epsilon)-Schottky subgroup of Γ\Gamma which is η\eta-narrow around g.g.

Since r1r_{1} is determined by Γ,\Gamma, we can assume at first that r0<r1.r_{0}<r_{1}. Now we fix an element gW4.g\in W_{4}. Then we can find a Zariski dense (r0,ϵ0)(r_{0},\epsilon_{0})-Schottky subgroup Γ1\Gamma_{1} of Γ\Gamma which is ϵ0\epsilon_{0}-narrow around g.g. By the proof of k=0k=0 in Lemma 3.9 (see also [MS23, Lemma 3.6]), we can find a finite subset {θi:i=1,,p}Γ1,\{\theta_{i}:i=1,\cdots,p\}\subset\Gamma_{1}, which generates Zariski dense sub-semigroup. Let W=W4{θi:i=1,,p},W^{\prime}=W_{4}\cup\left\{\theta_{i}:i=1,\cdots,p\right\}, which consists of (r0,ϵ0)(r_{0},\epsilon_{0})-loxodromic elements. Moreover, every element in WW^{\prime} is ϵ0\epsilon_{0}-narrow around g.g. Hence WW^{\prime} is 2ϵ02\epsilon_{0}-narrow. Therefore, WW^{\prime} is a (r0/4,ϵ0)(r_{0}/4,\epsilon_{0})-Schottky family and the semigroup it generates is a (r0/8,2ϵ0)(r_{0}/8,2\epsilon_{0})-Shottky family, by Lemma 5.17.

Take an mm\in{\mathbb{N}} large enough depending on θi\theta_{i}, β\beta and ϵ0\epsilon_{0}. Let

W5W4m{θigm,i=1,,p}W_{5}\coloneqq W_{4}^{*m}\cup\{\theta_{i}g^{m},\ i=1,\cdots,p\}

Now we verify that the set 𝒥W5\mathcal{J}\coloneqq W_{5} satisfies the condition for N=nm.N=nm.

  1. (1)

    This is because xx is given by Lemma 5.13.

  2. (2)

    Let 𝐇{\mathbf{H}} be the Zariski closure of the semigroup generated by W5,W_{5}, which is an algebraic subgroup of SL3().\mathrm{SL}_{3}({\mathbb{R}}). Note that gmW4mW5g^{m}\in W_{4}^{*m}\subset W_{5} and θigmW5\theta_{i}g^{m}\in W_{5} for every i.i. We obtain θi𝐇.\theta_{i}\in{\mathbf{H}}. Since {θi:i=1,,p}\left\{\theta_{i}:i=1,\cdots,p\right\} generates a Zariski dense subgroup, we have 𝐇=SL3().{\mathbf{H}}=\mathrm{SL}_{3}({\mathbb{R}}).

  3. (3)

    Recall that the semigroup generated by WW^{\prime} is a (r,ϵ)(r,\epsilon)-Schottky family, where r=r0/8r=r_{0}/8 and ϵ=2ϵ0.\epsilon=2\epsilon_{0}. Also recall that W4W_{4} is (n,4β,x)(n,4\beta,x)-approximate. By 5.17, for every hW4m,h\in W_{4}^{*m}, we have

    κ(h)Nx=κ(h)nmxCrn+4β.\|\frac{\kappa(h)}{N}-x\|=\|\frac{\kappa(h)}{nm}-x\|\leqslant\frac{C_{r}}{n}+4\beta.

    For h{θigm,i=1,,p}h\in\{\theta_{i}g^{m},i=1,\dots,p\}, we have

    κ(h)Nx=κ(h)nmxCrn+4β+κ(θi)nm.\|\frac{\kappa(h)}{N}-x\|=\|\frac{\kappa(h)}{nm}-x\|\leqslant\frac{C_{r}}{n}+4\beta+\frac{\left\|\kappa(\theta_{i})\right\|}{nm}.

    By taking mm large enough depending on θi,β\theta_{i},\beta and then taking nn large enough, we assume that Cr/n+κ(θi)/(nm)<β.C_{r}/n+\left\|\kappa(\theta_{i})\right\|/(nm)<\beta. Therefore, the set 𝒥=W5\mathcal{J}=W_{5} is (N,5β,x)(N,5\beta,x)-approximate. Since the semigroup generated by W5W_{5} is an (r,ϵ)(r,\epsilon)-Schottky family, for every j𝒥k,j\in\mathcal{J}^{*k}, we have

    κ(j)kNxCrN+5β10β.\|\frac{\kappa(j)}{kN}-x\|\leqslant\frac{C_{r}}{N}+5\beta\leqslant 10\beta.
  4. (4)

    We take minimal elements (W4)min(W_{4})_{\min} in W4W_{4}. Let 𝒥1=((W4)min)m.\mathcal{J}_{1}=((W_{4})_{\min})^{*m}. Then there is no element in 𝒥1\mathcal{J}_{1} which is a prefix of another element. Note that W4W_{4} is (n,4β,x)(n,4\beta,x)-approximate and every element in W4W_{4} is ending with A1A2A3.A_{1}A_{2}A_{3}. By 5.15, we have

    #𝒥1=(#(W4)min)m(#W4/e4Cβn)me(14β4Cβ)nme(110Cβ)N.\#\mathcal{J}_{1}=(\#(W_{4})_{\min})^{m}\geqslant(\#W_{4}/e^{4C\beta n})^{m}\geqslant e^{(1-4\beta-4C\beta)nm}\geqslant e^{(1-10C\beta)N}.\qed

Finally, we will construct the random walk and estimate the dimension of the stationary measure. This part is to complete the proof of Lemma 5.10.

Let 𝒥,𝒥1\mathcal{J},\mathcal{J}_{1} be the sets given by Lemma 5.18. We take ν=(1β)ν1+βν2,\nu=(1-\beta)\nu_{1}+\beta\nu_{2}, where ν1\nu_{1} is the uniform measure on 𝒥1\mathcal{J}_{1} and ν2\nu_{2} is the uniform measure on 𝒥𝒥1\mathcal{J}\setminus\mathcal{J}_{1}. Then by 5.18 the support of ν\nu generates a Zariski dense subgroup in SL3()\mathrm{SL}_{3}(\mathbb{R}) and the associated Lyapunov vector is close to Nx.Nx. Hence, by 5.18(3),

|ψs(λ(ν))N1||ψs(x)1|+5λ(ν)Nx100β.|\frac{\psi_{s}(\lambda(\nu))}{N}-1|\leqslant|\psi_{s}(x)-1|+5\cdot\|\frac{\lambda(\nu)}{N}-x\|\leqslant 100\beta. (5.4)

Now we should estimate the random walk entropy of ν.\nu. Firstly, note that 𝒥1\mathcal{J}_{1} freely generates a free semigroup, we have

hRW(ν1)=H(ν1)(110Cβ)N.h_{\mathrm{RW}}(\nu_{1})=H(\nu_{1})\geqslant(1-10C\beta)N.

Moreover, since the support of ν1\nu_{1} satisfies that no element is a prefix of another one, by certain “continuity" property of the random walk entropy and freeness of Γ\Gamma, we have

Lemma 5.21.

Let ν,ν1,ν2\nu,\nu_{1},\nu_{2} be probability measures which supported on Γ\Gamma such that ν=(1β)ν1+βν2\nu=(1-\beta)\nu_{1}+\beta\nu_{2}. If the support of ν1\nu_{1} is a minimal set (i.e. no element is a prefix of another), then

hRW(ν)(1β)hRW(ν1).h_{\mathrm{RW}}(\nu)\geqslant(1-\beta)h_{\mathrm{RW}}(\nu_{1}). (5.5)
Proof.

By the concavity of the entropy function HH,

H(ν11ν2j1ν12ν2jk)H(ν11δg1ν12δgk)dν2j1(g1)dν2jk(gk).H(\nu_{1}^{*\ell_{1}}*\nu_{2}^{*j_{1}}*\nu_{1}^{*\ell_{2}}*\cdots*\nu_{2}^{*j_{k}})\geqslant\int H(\nu_{1}^{*\ell_{1}}*\delta_{g_{1}}*\nu_{1}^{*\ell_{2}}*\cdots*\delta_{g_{k}})\mathrm{d}\nu_{2}^{*j_{1}}(g_{1})\cdots\mathrm{d}\nu_{2}^{*j_{k}}(g_{k}).

We claim that H(ν11δg1ν12δgk)=H(ν1(1++k))H(\nu_{1}^{*\ell_{1}}*\delta_{g_{1}}*\nu_{1}^{*\ell_{2}}*\cdots*\delta_{g_{k}})=H(\nu_{1}^{*(\ell_{1}+\cdots+\ell_{k})}). It suffices to show that all the elements in the support of LHS are distinct. Otherwise suppose that two elements (h1,,hk)(h_{1},\cdots,h_{k}) and (h1,,hk)(h_{1}^{\prime},\cdots,h_{k}^{\prime}) in suppν11××suppν1k\operatorname{\mathrm{supp}}\nu_{1}^{*\ell_{1}}\times\cdots\times\operatorname{\mathrm{supp}}\nu_{1}^{*\ell_{k}} satisfy h1g1hkgk=h1g1hkgkh_{1}g_{1}\cdots h_{k}g_{k}=h_{1}^{\prime}g_{1}\cdots h_{k}^{\prime}g_{k}. Since no element in suppν1\operatorname{\mathrm{supp}}\nu_{1} is a prefix of another one, so does suppν11\operatorname{\mathrm{supp}}\nu_{1}^{*\ell_{1}}. Therefore by freeness of Γ\Gamma, we obtain that h1=h1h_{1}=h_{1}^{\prime}. By induction we get h2=h2,,hk=hkh_{2}=h_{2}^{\prime},\cdots,h_{k}=h_{k}^{\prime}. Therefore

H(ν11ν2j1ν12ν2jk)H(ν1(1++k))dν2j1(g1)𝑑ν2jk(gk)=H(ν1(1++k)).H(\nu_{1}^{*\ell_{1}}*\nu_{2}^{*j_{1}}*\nu_{1}^{*\ell_{2}}*\cdots*\nu_{2}^{*j_{k}})\geqslant\int H(\nu_{1}^{*(\ell_{1}+\cdots+\ell_{k})})\mathrm{d}\nu_{2}^{*j_{1}}(g_{1})\cdots d\nu_{2}^{*j_{k}}(g_{k})=H(\nu_{1}^{*(\ell_{1}+\cdots+\ell_{k})}).

Combining with the concavity of HH, we get

hRW((1β)ν1+βν2)\displaystyle h_{\mathrm{RW}}((1-\beta)\nu_{1}+\beta\nu_{2}) =\displaystyle= lim1H(((1β)ν1+βν2))\displaystyle\lim_{\ell\rightarrow\infty}\frac{1}{\ell}H(((1-\beta)\nu_{1}+\beta\nu_{2})^{*\ell})
\displaystyle\geqslant lim1j(lj)(1β)jβjH(ν1(j))\displaystyle\lim_{\ell\rightarrow\infty}\frac{1}{\ell}\sum_{j}\binom{l}{j}(1-\beta)^{\ell-j}\beta^{j}H(\nu_{1}^{*(\ell-j)})
=\displaystyle= lim1j(j)(1β)jβj(j)hRW(ν1)\displaystyle\lim_{\ell\rightarrow\infty}\frac{1}{\ell}\sum_{j}\binom{\ell}{j}(1-\beta)^{\ell-j}\beta^{j}(\ell-j)h_{\mathrm{RW}}(\nu_{1})
\displaystyle\geqslant lim(1β)j(1j)(1β)j1βjhRW(ν1)\displaystyle\lim_{\ell\rightarrow\infty}(1-\beta)\sum_{j}\binom{\ell-1}{j}(1-\beta)^{\ell-j-1}\beta^{j}h_{\mathrm{RW}}(\nu_{1})
=\displaystyle= (1β)hRW(ν1).\displaystyle(1-\beta)h_{\mathrm{RW}}(\nu_{1}).

As a consequence hRW(ν)(1β)hRW(ν1)(1β)(110Cβ)Nh_{\mathrm{RW}}(\nu)\geqslant(1-\beta)h_{\mathrm{RW}}(\nu_{1})\geqslant(1-\beta)(1-10C\beta)N. Moreover, the group generated by suppν\operatorname{\mathrm{supp}}\nu is Zariski dense in SL3()\mathrm{SL}_{3}(\mathbb{R}) and satisfies exponential separation property (actually discreteness). Let μ\mu be the unique ν\nu-stationary measure. By the dimension formula, i.e. Theorem 1.2, we have dimμ=dimLYμ.\dim\mu=\dim_{\mathrm{LY}}\mu.

We admit the fact that hF(μ,ν)=hRW(ν)h_{\mathrm{F}}(\mu,\nu)=h_{\mathrm{RW}}(\nu) and postpone the proof to the end of this section. We give an estimate on dimLYμ.\dim_{\mathrm{LY}}\mu. Since ψs(λ(ν))(1100β)N,\psi_{s}(\lambda(\nu))\geqslant(1-100\beta)N, we obtain

s(λ1(ν)λ3(ν))ψs(λ(ν))(1100β)N.s(\lambda_{1}(\nu)-\lambda_{3}(\nu))\geqslant\psi_{s}(\lambda(\nu))\geqslant(1-100\beta)N.

Then λ1(ν)λ3(ν)12sN\lambda_{1}(\nu)-\lambda_{3}(\nu)\geqslant\frac{1}{2s}N assuming β\beta small enough. For a given 0<ϵ<s1,0<\epsilon<s-1, we have

ψsϵ(λ(ν))=ψs(λ(ν))ϵ(λ1(ν)λ3(ν))(1+100β)Nϵ2sN.\psi_{s-\epsilon}(\lambda(\nu))=\psi_{s}(\lambda(\nu))-\epsilon(\lambda_{1}(\nu)-\lambda_{3}(\nu))\leqslant(1+100\beta)N-\frac{\epsilon}{2s}N.

Now we take β>0\beta>0 sufficiently small comparing to ϵ,\epsilon, we obtain

ψsϵ(λ(ν))(1+100β)Nϵ2sN(1β)(110Cβ)NhRW(ν).\psi_{s-\epsilon}(\lambda(\nu))\leqslant(1+100\beta)N-\frac{\epsilon}{2s}N\leqslant(1-\beta)(1-10C\beta)N\leqslant h_{\mathrm{RW}}(\nu).

Therefore, dimμ=dimLYμsϵ.\dim\mu=\dim_{\mathrm{LY}}\mu\geqslant s-\epsilon.

To complete the proof, it remains to show the identity between the Furstenberg entropy and the random walk entropy.

Lemma 5.22.

Let ν\nu be a finitely supported probability measure on Γ\Gamma such that GνG_{\nu} is Zariski dense. Let μ\mu be the unique ν\nu-stationary measure on (3),{\mathbb{P}}({\mathbb{R}}^{3}), then hF(μ,ν)=hRW(ν).h_{\mathrm{F}}(\mu,\nu)=h_{\mathrm{RW}}(\nu).

Proof.

As discussions after Definition 5.8, XXniceX\setminus X_{\mathrm{nice}} is a countable set and hence a μ\mu-null set. We consider the space B=SL3()×B=\mathrm{SL}_{3}({\mathbb{R}})^{\times{\mathbb{N}}} endowing with the probability measure ν.\nu^{{\mathbb{N}}}. For almost every b=(b1,b2,)B,b=(b_{1},b_{2},\cdots)\in B, we consider the Furstenberg boundary μb\mu_{b} given by the weak* limit

limn(b1b2bn1)μ,\lim_{n\to\infty}(b_{1}b_{2}\cdots b_{n-1})_{*}\mu,

which is a Dirac measure [BQ16, Lemma 4.5] and denoted by δξ(b).\delta_{\xi(b)}. Because μ=δξ(b)dν(b)\mu=\int\delta_{\xi(b)}\mathrm{d}\nu^{{\mathbb{N}}}(b) and μ(XXnice)=0,\mu(X\setminus X_{\mathrm{nice}})=0, there exists a conull set ΩB\Omega\subset B such that ξ(b)Xnice\xi(b)\in X_{\mathrm{nice}} for every bΩ.b\in\Omega.

Note that ν\nu also induces a random walk on the flag variety =(3)\mathcal{F}=\mathcal{F}({\mathbb{R}}^{3}) with a unique stationary measure ν.\nu_{\mathcal{F}}. We can also consider the Furstenberg boundary on the flag variety. Then for almost every bB,b\in B, we can associate a Dirac measure (μ)b=δξ(b).(\mu_{\mathcal{F}})_{b}=\delta_{\xi_{\mathcal{F}}(b)}. Denote ξ(b)=(ξ(b),ξ2(b))(3),\xi_{\mathcal{F}}(b)=(\xi(b),\xi_{2}(b))\in\mathcal{F}({\mathbb{R}}^{3}), where ξ2(b)\xi_{2}(b) is a two dimensional subspace in 3.{\mathbb{R}}^{3}. Then for every sequence of positive numbers (χn),(\chi_{n}), the image of every nonzero limit of (χnb0b1bn1)(\chi_{n}b_{0}b_{1}\cdots b_{n-1}) in End(23)\mathrm{End}(\wedge^{2}\mathbb{R}^{3}) is ξ2(b).\xi_{2}(b). We aim to show that ξ(b)\xi_{\mathcal{F}}(b) is uniquely determined by ξ(b)\xi(b) for every bΩ.b\in\Omega.

Let bbΩ,b\neq b^{\prime}\in\Omega, then ξ(b)=ξ(b)Xnice.\xi(b)=\xi(b^{\prime})\in X_{\mathrm{nice}}. Recall that every point in XniceX_{\mathrm{nice}} is uniquely coding by an element in {1,2,3}.\left\{1,2,3\right\}^{\mathbb{N}}. We know that b=(b1,b2,)b^{\prime}=(b_{1}^{\prime},b_{2}^{\prime},\cdots) and b=(b1,b2,)b=(b_{1},b_{2},\cdots) are different partitions of a same infinite word Ai1Ai2AinA_{i_{1}}A_{i_{2}}\cdots A_{i_{n}}\cdots. Since ν\nu is finitely supported, for every n0n\geqslant 0 we can find mn0m_{n}\geqslant 0 such that

|b1b2bn||b1b2bmn||b1b2bn|+Nν,|b_{1}b_{2}\cdots b_{n}|\leqslant|b_{1}^{\prime}b_{2}^{\prime}\cdots b_{m_{n}}^{\prime}|\leqslant|b_{1}b_{2}\cdots b_{n}|+N_{\nu},

where |||\cdot| denotes the word length in the semigroup Γ\Gamma with respect to the free generating set {A1,A2,A3},\left\{A_{1},A_{2},A_{3}\right\}, where NνN_{\nu} is a constant only depending on ν.\nu.

Passing to a subsequence, we can assume that (b1bmn)1b1bn(b_{1}^{\prime}\cdots b_{m_{n}}^{\prime})^{-1}b_{1}\cdots b_{n} is a fixed element in SL3().\mathrm{SL}_{3}({\mathbb{R}}). Taking χn=b1b2bn1\chi_{n}=\|b_{1}b_{2}\cdots b_{n}\|^{-1} and passing to a subsequence if necessary, we also assume that (χnb1b2bn)(\chi_{n}b_{1}b_{2}\cdots b_{n}) admits the nonzero limit in End(23).\mathrm{End}(\wedge^{2}\mathbb{R}^{3}). Meanwhile, (χnb1bmn)(\chi_{n}b_{1}^{\prime}\cdots b_{m_{n}}^{\prime}) converges to this limit composing with an element in SL3().\mathrm{SL}_{3}({\mathbb{R}}). Therefore, these two limits has the same image. By our discussions on ξ2,\xi_{2}, we obtain ξ2(b)=ξ2(b)\xi_{2}(b^{\prime})=\xi_{2}(b) and hence

ξ(b)=ξ(b).\xi_{\mathcal{F}}(b)=\xi_{\mathcal{F}}(b^{\prime}). (5.6)

Now we consider a conditional measure at a μ\mu-full measure set ξ(Ω)(3)\xi(\Omega)\subset{\mathbb{P}}({\mathbb{R}}^{3}). For every xξ(Ω)x\in\xi(\Omega), choose an element bΩb\in\Omega with ξ(b)=x(3)\xi(b)=x\in\mathbb{P}(\mathbb{R}^{3}) and let

μx=δξ(b).\mu^{x}=\delta_{\xi_{\mathcal{F}}(b)}.

Due to Eq. 5.6, we have

μ=Ωδξ(b)dν(b)=(3)μxdμ(x).\mu_{\mathcal{F}}=\int_{\Omega}\delta_{\xi_{\mathcal{F}}(b)}\mathrm{d}\nu^{\mathbb{N}}(b)=\int_{\mathbb{P}(\mathbb{R}^{3})}\mu^{x}\mathrm{d}\mu(x).

So the family of measures μx\mu^{x} is actually the disintegration of μ\mu_{\mathcal{F}} with respect to the natural projection π:(3).\pi:\mathcal{F}\to{\mathbb{P}}({\mathbb{R}}^{3}). Hence we obtain the trivial-fiber property. Now we can apply a same argument as in the end of the proof 2.3 by using relative measure preserving property. We obtain the desired equality between the Furstenberg entropy and the random walk entropy. ∎

Appendix A Proof of 1.5

In the appendix, we prove 1.5, that is

sA(Γ)3/2.s_{\mathrm{A}}(\Gamma_{\mathscr{R}})\geqslant 3/2. (A.1)

The secret sauce is to study the semigroup Γ0=A1,A2\Gamma_{0}=\langle A_{1},A_{2}\rangle. Let I=I(E1,E2)I=I(E_{1},E_{2}) be the arc given by {(se1+te2):s,t0}\{\mathbb{R}(se_{1}+te_{2}):\ s,t\in\mathbb{R}_{\geqslant 0}\}, which is preserved by Γ0\Gamma_{0}. The idea is as follows. We can view the semigroup Γ0\Gamma_{0} in two aspects: Γ0\Gamma_{0}, as a subsemigroup in GL(E1E2)\mathrm{GL}(E_{1}\oplus E_{2}), has critical exponent 11 due to its limit set is the whole II; Γ0\Gamma_{0} is also a subsemigroup of SL3()\mathrm{SL}_{3}(\mathbb{R}). Then, we compare the singular values in these two settings and deduce that its affinity exponent is at least 1.51.5. The argument is similar to the one of the dimension jump of the limit sets of representations in Barbot’s component as in [LPX23]. The difficulty in the proof is due to the non-uniform hyperbolic behaviour of the Rauzy semigroup, and we borrow estimates from Section 5.2 to deal with this issue.

Lemma A.1.

There exists ϵ>0\epsilon>0 such that for any γΓ0\gamma\in\Gamma_{0} with the last two digits different, we have

σ2(γ)ϵ,ϵ|γI|1/σ1(γ)2,\displaystyle\sigma_{2}(\gamma)\geqslant\epsilon,\ \epsilon|\gamma I|\leqslant 1/\sigma_{1}(\gamma)^{2}, (A.2)

where |γI||\gamma I| is the length of the arc γI\gamma I.

Proof.

It is a consequence of 5.2 and 5.6.

Due to γ\gamma preserving the subspace generate by E1,E2E_{1},E_{2}, and its restriction having determinant 1, we obtain from 5.2 that

|γI|=d(γE1,γE2)=γe1γe2γe1γe2=1γe1γe21/(ϵ2σ1(γ))2.|\gamma I|=d(\gamma E_{1},\gamma E_{2})=\frac{\|\gamma e_{1}\wedge\gamma e_{2}\|}{\|\gamma e_{1}\|\|\gamma e_{2}\|}=\frac{1}{\|\gamma e_{1}\|\|\gamma e_{2}\|}\leqslant 1/(\epsilon_{2}\sigma_{1}(\gamma))^{2}. (A.3)

Due to the same computation as in 5.6 and from definition, we have

Area(γΔ)γe1γe2γe3/1i3ω(γei)1/σ1(γ)3.\mathrm{Area}(\gamma\Delta)\gg\|\gamma e_{1}\wedge\gamma e_{2}\wedge\gamma e_{3}\|/\prod_{1\leqslant i\leqslant 3}\omega(\gamma e_{i})\gg 1/\sigma_{1}(\gamma)^{3}.

Taking into accout Eq. A.3,

max{|γI(E2,E3)|,|γI(E1,E3)|}Area(γΔ)/|γI|1/σ1(γ).\max\{|\gamma I(E_{2},E_{3})|,|\gamma I(E_{1},E_{3})|\}\geqslant\mathrm{Area}(\gamma\Delta)/|\gamma I|\gg 1/\sigma_{1}(\gamma).

Combining with 5.6, we have

σ2(γ)σ1(γ)diam(γΔ)σ1(γ)max{|γI(E2,E3)|,|γI(E1,E3)|}1.\sigma_{2}(\gamma)\gg\sigma_{1}(\gamma)\operatorname{\mathrm{diam}}(\gamma\Delta)\geqslant\sigma_{1}(\gamma)\max\{|\gamma I(E_{2},E_{3})|,|\gamma I(E_{1},E_{3})|\}\gg 1.

The proof is complete. ∎

Proof of Eq. A.1.

For elements satisfying A.1, we have

φ3/2(γ)=(σ2σ1)(γ)(σ3σ1)1/2(γ)=σ2(γ)1/2σ1(γ)2ϵ1/2σ1(γ)2ϵ2|γI|.\varphi_{3/2}(\gamma)=\left(\frac{\sigma_{2}}{\sigma_{1}}\right)(\gamma)\left(\frac{\sigma_{3}}{\sigma_{1}}\right)^{1/2}(\gamma)=\frac{\sigma_{2}(\gamma)^{1/2}}{\sigma_{1}(\gamma)^{2}}\geqslant\frac{\epsilon^{1/2}}{\sigma_{1}(\gamma)^{2}}\geqslant\epsilon^{2}|\gamma I|.

Therefore

γΓφ3/2(γ)γΓ0φ3/2(γ)\displaystyle\sum_{\gamma\in\Gamma_{\mathscr{R}}}\varphi_{3/2}(\gamma)\geqslant\sum_{\gamma\in\Gamma_{0}}\varphi_{3/2}(\gamma) γΓ0, last two digits differentφ3/2(γ)\displaystyle\geqslant\sum_{\gamma\in\Gamma_{0},\text{ last two digits different}}\varphi_{3/2}(\gamma)
ϵ2γΓ0, last two digits different|γI|.\displaystyle\geqslant\epsilon^{2}\sum_{\gamma\in\Gamma_{0},\text{ last two digits different}}|\gamma I|.

Due to 5.9, every point in IXniceI\cap X_{\mathrm{nice}} is covered infinity many times by the cover

{γI,γΓ0 whose last two digits are different}.\{\gamma I,\ \gamma\in\Gamma_{0}\text{ whose last two digits are different}\}.

Since I(IXnice)I-(I\cap X_{\mathrm{nice}}) is countable, the righthand side of the last inequality is infinite, which implies sA(Γ)1.5s_{\mathrm{A}}(\Gamma_{\mathscr{R}})\geqslant 1.5. ∎

References

  • [AHS16] Artur Avila, Pascal Hubert, and Alexandra Skripchenko. On the Hausdorff dimension of the Rauzy gasket. Bulletin de la Société Mathématique de France, 144(3):539–568, 2016.
  • [AMS95] Herbert Abels, Grigorij A Margulis, and Grigorij A Soifer. Semigroups containing proximal linear maps. Israel journal of mathematics, 91:1–30, 1995.
  • [AR91] Pierre Arnoux and Gérard Rauzy. Geometric representation of sequences of complexity (2n+1)(2n+1). Bulletin de la Société Mathématique de France, 119(2):199–215, 1991.
  • [AS13] Pierre Arnoux and Štěpán Starosta. The Rauzy Gasket. In Julien Barral and Stéphane Seuret, editors, Further Developments in Fractals and Related Fields: Mathematical Foundations and Connections, Trends in Mathematics, pages 1–23. Birkhäuser Boston, Boston, 2013.
  • [Ave72] Andre Avez. Entropie des groupes de type fini. Comptes Rendus Hebdomadaires des Séances de l’Académie des Sciences, Série A, 275:1363–1366, 1972.
  • [Ben97] Y. Benoist. Propriétés Asymptotiques des Groupes Linéaires:. Geometric and Functional Analysis, 7(1):1–47, March 1997.
  • [BHR19] Balázs Bárány, Michael Hochman, and Ariel Rapaport. Hausdorff dimension of planar self-affine sets and measures. Inventiones Mathematicae, 216(3):601–659, 2019.
  • [Bor91] Armand Borel. Linear algebraic groups, volume 126 of Graduate Texts in Mathematics. Springer-Verlag, New York, second edition, 1991.
  • [BPS19] Jairo Bochi, Rafael Potrie, and Andrés Sambarino. Anosov representations and dominated splittings. Journal of the European Mathematical Society, 21(11):3343–3414, July 2019.
  • [BQ16] Yves Benoist and Jean-François Quint. Random Walks on Reductive Groups, volume 62. Springer, 2016.
  • [Can] Richard Canary. Anosov representations: Informal lecture notes.
  • [DD09] Roberto DeLeo and Ivan A Dynnikov. Geometry of plane sections of the infinite regular skew polyhedron {4,6|4}\{4,6|4\}. Geometriae Dedicata, 138:51–67, 2009.
  • [Fal88] K. J. Falconer. The Hausdorff dimension of self-affine fractals. Mathematical Proceedings of the Cambridge Philosophical Society, 103(2):339–350, March 1988.
  • [FH09] De-Jun Feng and Huyi Hu. Dimension theory of iterated function systems. Communications on Pure and Applied Mathematics: A Journal Issued by the Courant Institute of Mathematical Sciences, 62(11):1435–1500, 2009.
  • [Fou20] Charles Fougeron. Dynamical properties of simplicial systems and continued fraction algorithms. arXiv preprint arXiv:2001.01367, 2020.
  • [Fur63] H. Furstenberg. Noncommuting random products. Transactions of the American Mathematical Society, 108:377–428, 1963.
  • [Fur02] Alex Furman. Random walks on groups and random transformations. In B. Hasselblatt and A. Katok, editors, Handbook of Dynamical Systems, volume 1, pages 931–1014. Elsevier Science, January 2002.
  • [GM89] I Ya Goldsheid and G.A. Margulis. Lyapunov exponents of random matrices product. Usp. Mat. Nauk, 44:13–60, 1989.
  • [Gou22] Sébastien Gouëzel. Exponential bounds for random walks on hyperbolic spaces without moment conditions. Tunis. J. Math., 4(4):635–671, 2022.
  • [GR85] Y. Guivarc’h and A. Raugi. Frontière de Furstenberg, propriétés de contraction et théorèmes de convergence. Zeitschrift für Wahrscheinlichkeitstheorie und Verwandte Gebiete, 69(2):187–242, 1985.
  • [GRM20] Rodolfo Gutierrez-Romo and Carlos Matheus. Lower bounds on the dimension of the Rauzy gasket. Bulletin de la Société Mathématique de France, 148(2):321–327, 2020.
  • [Hem04] John Hempel. 3-manifolds. AMS Chelsea Publishing, Providence, RI, 2004. Reprint of the 1976 original.
  • [HJX23] Weikun He, Yuxiang Jiao, and Disheng Xu. On dimension theory of random walks and group actions by circle diffeomorphisms. arXiv preprint arXiv:2304.08372, 2023.
  • [HS17] Michael Hochman and Boris Solomyak. On the dimension of Furstenberg measure for SL2()\mathrm{SL_{2}}(\mathbb{R})-random matrix products. Inventiones mathematicae, 210(3):815–875, December 2017.
  • [KV83] V. A. Kaĭmanovich and A. M. Vershik. Random walks on discrete groups: boundary and entropy. The Annals of Probability, 11(3):457–490, 1983.
  • [Lab06] François Labourie. Anosov flows, surface groups and curves in projective space. Inventiones Mathematicae, 165(1):51–114, 2006.
  • [Led85] François Ledrappier. Poisson boundaries of discrete groups of matrices. Israel Journal of Mathematics, 50(4):319–336, December 1985.
  • [Lev93] Gilbert Levitt. La dynamique des pseudogroupes de rotations. Inventiones mathematicae, 113:633–670, 1993.
  • [LL23a] François Ledrappier and Pablo Lessa. Dimension gap for the limit sets of anosov representations, 2023.
  • [LL23b] François Ledrappier and Pablo Lessa. Exact dimension of Furstenberg measures. Geometric and Functional Analysis, 33(1):245–298, February 2023.
  • [LPX23] Jialun Li, Wenyu Pan, and Disheng Xu. On the dimension of limit sets on (3)\mathbb{P}(\mathbb{R}^{3}) via stationary measures: the theory and applications. Preprint, 2023.
  • [MS19] Ian D. Morris and Pablo Shmerkin. On equality of Hausdorff and affinity dimensions, via self-affine measures on positive subsystems. Transactions of the American Mathematical Society, 371(3):1547–1582, 2019.
  • [MS23] Ian D. Morris and Cagri Sert. A variational principle relating self-affine measures to self-affine sets, March 2023. arXiv:2303.03437 [math].
  • [PS21] Mark Pollicott and Benedict Sewell. An upper bound on the dimension of the Rauzy gasket, October 2021. arXiv:2110.07264 [math].
  • [PSW22] Beatrice Pozzetti, Andres Sambarino, and Anna Wienhard. Anosov representations with Lipschitz limit set. Geometry and Topology, 2022.
  • [Rap21] Ariel Rapaport. Exact dimensionality and Ledrappier-Young formula for the Furstenberg measure. Transactions of the American Mathematical Society, 374(7):5225–5268, April 2021.
  • [Yan19] Wen-yuan Yang. Statistically convex-cocompact actions of groups with contracting elements. Int. Math. Res. Not. IMRN, (23):7259–7323, 2019.
  • [You82] Lai-Sang Young. Dimension, entropy and Lyapunov exponents. Ergodic Theory and Dynamical Systems, 2(1):109–124, March 1982.

Yuxiang Jiao. Peking University, No.5 Yiheyuan Road, Haidian District, Beijing, China.
email: ajorda@pku.edu.cn

Jialun Li. CNRS-Centre de Mathématiques Laurent Schwartz, École Polytechnique, Palaiseau, France.
email: jialun.li@cnrs.fr


Wenyu Pan. University of Toronto, 40 St. George St., Toronto, ON, M5S 2E4, Canada.
email: wenyup.pan@utoronto.ca


Disheng Xu. Great Bay University, Songshanhu International Community, Dongguan, Guangdong, 523000, China.
email: xudisheng@gbu.edu.cn