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

Dimension conservation of harmonic measures in products of hyperbolic spaces

Ryokichi Tanaka Department of Mathematics, Kyoto University, Kyoto 606-8502 JAPAN rtanaka@math.kyoto-u.ac.jp
(Date: September 3, 2025)
Abstract.

We show that the harmonic measure on a product of boundaries satisfies dimension conservation for a random walk with non-elementary marginals on a countable group acting on a product of hyperbolic spaces under the finite first moment condition.

1. Introduction

Let Γ\Gamma and Γ{\Gamma^{\star}} be non-elementary hyperbolic groups. We study a random walk on the product group 𝚪:=Γ×Γ{\bm{\Gamma}}:=\Gamma\times{\Gamma^{\star}} and establish a dimension formula for the harmonic measure on the product of (Gromov) boundaries. After stating our results in this special case, we consider a countable group of isometries of a product of two hyperbolic metric spaces.

Let π\pi be a probability measure on 𝚪{\bm{\Gamma}} such that marginals μ\mu and μ{\mu^{\star}} are non-elementary, i.e., their supports generate non-elementary subgroups in Γ\Gamma and in Γ{\Gamma^{\star}} as groups respectively. For such a π\pi, a harmonic measure νπ\nu_{\pi} is defined on the product of boundaries Γ×Γ\partial\Gamma\times\partial{\Gamma^{\star}} (cf. Section 2.1). It is the unique probability measure satisfying that

νπ=πνπ,where πνπ=𝒙𝚪π(𝒙)𝒙νπ and 𝒙νπ:=νπ𝒙1.\nu_{\pi}=\pi\ast\nu_{\pi},\quad\text{where $\pi\ast\nu_{\pi}=\sum_{\bm{x}\in{\bm{\Gamma}}}\pi(\bm{x})\bm{x}\nu_{\pi}$ and $\bm{x}\nu_{\pi}:=\nu_{\pi}\circ\bm{x}^{-1}$}.

In the above, we consider the natural action of 𝚪=Γ×Γ{\bm{\Gamma}}=\Gamma\times{\Gamma^{\star}} on Γ×Γ\partial\Gamma\times\partial{\Gamma^{\star}}. The harmonic measure νπ\nu_{\pi} has marginals νμ\nu_{\mu} and νμ\nu_{{\mu^{\star}}} on Γ\partial\Gamma and on Γ\partial{\Gamma^{\star}} respectively, and these are determined by μ\mu and by μ{\mu^{\star}}. The measure νμ\nu_{\mu} (or νμ\nu_{{\mu^{\star}}}) for a single hyperbolic group and its generalization has attracted intensive studies, including the dimension, e.g., [Kai98, Led01], more recently, [BHM11, HS17, Tan19, DY23]. However, the harmonic measure νπ\nu_{\pi} for a product group has been studied only in a few cases (see [Vol21] for a special case of products of hyperbolic free product groups). Dimensional properties of such a measure exhibit new features since it contains different behaviors depending on the factors. This manifests an additional difficulty, which arises in a higher rank setting—–in that case, the boundaries are assembled in further intricate ways such as flag varieties. See [KLP11] for a thorough discussion on this matter of subject, and recent works [Les21, Rap21, LL23]. For related results on ergodic invariant measures for affine iterated function systems, see e.g., [KP96, FH09, Fen23]. We study the harmonic measure νπ\nu_{\pi} itself and the conditional measure νπη\nu_{\pi}^{\eta} of νπ\nu_{\pi} on Γ×Γ\partial\Gamma\times\partial{\Gamma^{\star}} for νμ\nu_{{\mu^{\star}}}-almost every ηΓ\eta\in\partial{\Gamma^{\star}}. It is shown that these are exact dimensional under a finite first moment condition. The dimension formula is a sum of ratios of asymptotic entropy over drift corrected with differential entropy. For the conditional measures, we provide a sufficient condition for the strict positivity of dimension.

Let us consider left invariant hyperbolic metrics dd and dd^{\star} quasi-isometric to word metrics in Γ\Gamma and in Γ{\Gamma^{\star}}, respectively. They induce quasi-metrics qq and q{q^{\star}} in the compactified spaces ΓΓ\Gamma\cup\partial\Gamma and ΓΓ{\Gamma^{\star}}\cup\partial{\Gamma^{\star}}, respectively, and let 𝒒(𝝃1,𝝃2):=max{q(ξ1,ξ2),q(η1,η2)}\bm{q}({\bm{\xi}}_{1},{\bm{\xi}}_{2}):=\max\{q(\xi_{1},\xi_{2}),{q^{\star}}(\eta_{1},\eta_{2})\} for 𝝃i=(ξi,ηi)(ΓΓ)×(ΓΓ){\bm{\xi}}_{i}=(\xi_{i},\eta_{i})\in(\Gamma\cup\partial\Gamma)\times({\Gamma^{\star}}\cup\partial{\Gamma^{\star}}) and i=1,2i=1,2 (cf. Section 2.1). Let us assume that π\pi has a finite first moment, i.e., 𝒙𝚪𝒅(𝐢𝐝,𝒙)π(𝒙)<\sum_{\bm{x}\in{\bm{\Gamma}}}\bm{d}(\operatorname{{\bf id}},\bm{x})\pi(\bm{x})<\infty for the identity element 𝐢𝐝\operatorname{{\bf id}} and 𝒅(𝒙1,𝒙2):=max{d(x1,x2),d(y1,y2)}\bm{d}(\bm{x}_{1},\bm{x}_{2}):=\max\{d(x_{1},x_{2}),d^{\star}(y_{1},y_{2})\} for 𝒙i=(xi,yi)𝚪\bm{x}_{i}=(x_{i},y_{i})\in{\bm{\Gamma}} and i=1,2i=1,2. The asymptotic entropy h(π)h(\pi) is defined as the limit for nn-fold convolutions πn:=πn\pi_{n}:=\pi^{\ast n},

h(π)=limn1n𝒙𝚪πn(𝒙)logπn(𝒙).h(\pi)=\lim_{n\to\infty}\frac{1}{n}\sum_{\bm{x}\in{\bm{\Gamma}}}-\pi_{n}(\bm{x})\log\pi_{n}(\bm{x}).

The drift l(Γ,μ)l(\Gamma,\mu) is defined as the limit for μn:=μn\mu_{n}:=\mu^{\ast n},

l(Γ,μ)=limn1nxΓd(id,x)μn(x).l(\Gamma,\mu)=\lim_{n\to\infty}\frac{1}{n}\sum_{x\in\Gamma}d(\operatorname{{\rm id}},x)\mu_{n}(x).

Similarly for l(Γ,μ)l({\Gamma^{\star}},{\mu^{\star}}). These are positive since μ\mu and μ{\mu^{\star}} are non-elementary (cf. Section 2.3). For every real r>0r>0 and and every 𝝃Γ×Γ{\bm{\xi}}\in\partial\Gamma\times\partial{\Gamma^{\star}}, let 𝑩(𝝃,r)\bm{B}({\bm{\xi}},r) denote the open ball of radius rr centered at 𝝃{\bm{\xi}} in (Γ×Γ,𝒒)(\partial\Gamma\times\partial{\Gamma^{\star}},\bm{q}).

Theorem 1.1.

Let Γ\Gamma and Γ{\Gamma^{\star}} be non-elementary hyperbolic groups, and π\pi be a probability measure on 𝚪=Γ×Γ{\bm{\Gamma}}=\Gamma\times{\Gamma^{\star}} with finite first moment and non-elementary marginals μ\mu and μ{\mu^{\star}} on Γ\Gamma and on Γ{\Gamma^{\star}} respectively. Suppose that l(Γ,μ)l(Γ,μ)l(\Gamma,\mu)\geq l({\Gamma^{\star}},{\mu^{\star}}). Then the harmonic measure νπ\nu_{\pi} on (Γ×Γ,𝐪)\left(\partial\Gamma\times\partial{\Gamma^{\star}},\bm{q}\right) is exact dimensional, i.e., for νπ\nu_{\pi}-almost every 𝛏Γ×Γ{\bm{\xi}}\in\partial\Gamma\times\partial{\Gamma^{\star}},

limr0logνπ(𝑩(𝝃,r))logr=h(π)h(μ)l(Γ,μ)+h(μ)l(Γ,μ).\lim_{r\to 0}\frac{\log\nu_{\pi}\left(\bm{B}({\bm{\xi}},r)\right)}{\log r}=\frac{h(\pi)-h({\mu^{\star}})}{l(\Gamma,\mu)}+\frac{h({\mu^{\star}})}{l({\Gamma^{\star}},{\mu^{\star}})}.

In particular, the Hausdorff dimension of νπ\nu_{\pi} is computed as

dimνπ=h(π)h(μ)l(Γ,μ)+h(μ)l(Γ,μ).\dim\nu_{\pi}=\frac{h(\pi)-h({\mu^{\star}})}{l(\Gamma,\mu)}+\frac{h({\mu^{\star}})}{l({\Gamma^{\star}},{\mu^{\star}})}.

Theorem 1.1 is shown in Theorem 4.5 in a more general setting. The above result is based on the exact dimensionality of disintegrated measures: Let νπη\nu_{\pi}^{\eta} for η𝒳\eta\in\partial{{\mathcal{X}}^{\star}} denote a system of conditional measures of νπ\nu_{\pi} on Γ×Γ\partial\Gamma\times\partial{\Gamma^{\star}} with respect to the σ\sigma-algebra generated by the projection from Γ×Γ\partial\Gamma\times\partial{\Gamma^{\star}} to Γ\partial{\Gamma^{\star}}.

Theorem 1.2.

Let Γ\Gamma and Γ{\Gamma^{\star}} be non-elementary hyperbolic groups, and π\pi be a probability measure on 𝚪=Γ×Γ{\bm{\Gamma}}=\Gamma\times{\Gamma^{\star}} with finite first moment and non-elementary marginals μ\mu and μ{\mu^{\star}} on Γ\Gamma and on Γ{\Gamma^{\star}} respectively. For νμ\nu_{{\mu^{\star}}}-almost every ηΓ\eta\in\partial{\Gamma^{\star}}, the conditional measure νπη\nu_{\pi}^{\eta} is exact dimensional on (Γ×Γ,𝐪)(\partial\Gamma\times\partial{\Gamma^{\star}},\bm{q}), i.e., for νπη\nu_{\pi}^{\eta}-almost every 𝛏Γ×Γ{\bm{\xi}}\in\partial\Gamma\times\partial{\Gamma^{\star}},

limr0logνπη(𝑩(𝝃,r))logr=h(π)h(μ)l(Γ,μ),\lim_{r\to 0}\frac{\log\nu_{\pi}^{\eta}(\bm{B}({\bm{\xi}},r))}{\log r}=\frac{h(\pi)-h({\mu^{\star}})}{l(\Gamma,\mu)},

In particular, the Hausdorff dimension of νπη\nu_{\pi}^{\eta} is computed as for νμ\nu_{{\mu^{\star}}}-almost every ηΓ\eta\in\partial{\Gamma^{\star}},

dimνπη=h(π)h(μ)l(Γ,μ).\dim\nu_{\pi}^{\eta}=\frac{h(\pi)-h({\mu^{\star}})}{l(\Gamma,\mu)}.

Theorem 1.2 is shown in Theorem 1.3 in a more general setting. Following Furstenberg [Fur08, Definition 3.1], we say that a Borel probability measure ν\nu on a product of compact metric spaces ×{\mathcal{M}}\times{\mathcal{M}}^{\star} satisfies dimension conservation if the following holds. Let us consider the pushforward ν\nu^{\star} and a system of conditional measures νη\nu^{\eta} for η\eta\in{\mathcal{M}}^{\star} of ν\nu associated with the projection ×{\mathcal{M}}\times{\mathcal{M}}^{\star}\to{\mathcal{M}}^{\star}: the measures ν\nu and ν\nu^{\star} are exact dimensional with dimension dimν\dim\nu and dimν\dim\nu^{\star} respectively, for ν\nu^{\star}-almost every η\eta\in{\mathcal{M}}^{\star}, conditional measures νη\nu^{\eta} are exact dimensional with dimension dimνη\dim\nu^{\eta}, and

dimν=dimνη+dimν.\dim\nu=\dim\nu^{\eta}+\dim\nu^{\star}.

In the above, we understand that the metric in ×{\mathcal{M}}\times{\mathcal{M}}^{\star} is defined as the maximum of metrics in {\mathcal{M}} and {\mathcal{M}}^{\star} (or an arbitrary one bi-Lipschitz to it). It has been shown that νμ\nu_{\mu^{\star}} on (Γ,q)(\partial{\Gamma^{\star}},{q^{\star}}) is exact dimensional with dimension h(μ)/l(Γ,μ)h({\mu^{\star}})/l({\Gamma^{\star}},{\mu^{\star}}) [Tan19, Theorem 3.8]. Therefore Theorems 1.1 and 1.2 imply that the harmonic measure νπ\nu_{\pi} on Γ×Γ\partial\Gamma\times\partial{\Gamma^{\star}} satisfies dimension conservation. The statement holds in a more general setting; see Section 4. For an extension to a product of more than two hyperbolic groups, see Remark 4.6.

Let (𝒳,d)({\mathcal{X}},d) and (𝒳,d)({{\mathcal{X}}^{\star}},d^{\star}) be proper roughly geodesic hyperbolic metric spaces with bounded growth at some scales (for the definitions, see Section 2). Examples of such spaces include Gromov hyperbolic Riemannian manifolds with sectional curvature bounded from below and from above and Cayley graphs of hyperbolic groups. The space 𝒳×𝒳{\mathcal{X}}\times{{\mathcal{X}}^{\star}} is equipped with a base point 𝒐\bm{o} and the metric 𝒅(𝒙1,𝒙2):=max{d(x1,x2),d(y1,y2)}\bm{d}(\bm{x}_{1},\bm{x}_{2}):=\max\{d(x_{1},x_{2}),d^{\star}(y_{1},y_{2})\} for 𝒙i=(xi,yi)𝒳×𝒳\bm{x}_{i}=(x_{i},y_{i})\in{\mathcal{X}}\times{{\mathcal{X}}^{\star}} and i=1,2i=1,2. Let us consider a countable subgroup 𝚪{\bm{\Gamma}} in the product of isometry groups Isom𝒳×Isom𝒳\operatorname{Isom}{\mathcal{X}}\times\operatorname{Isom}{{\mathcal{X}}^{\star}}. We say that 𝚪{\bm{\Gamma}} has a finite exponential growth relative to (𝒳×𝒳,𝒅)({\mathcal{X}}\times{{\mathcal{X}}^{\star}},\bm{d}) if there exists a constant c>0c>0 such that for all r>0r>0,

#{𝒙𝚪:𝒅(𝒐,𝒙𝒐)<r}cecr.\#\big{\{}\bm{x}\in{\bm{\Gamma}}\ :\ \bm{d}(\bm{o},\bm{x}\cdot\bm{o})<r\big{\}}\leq ce^{cr}.

In the above, #A\#A denotes the cardinality of a set AA. For a probability measure π\pi on 𝚪{\bm{\Gamma}}, let μ\mu and μ{\mu^{\star}} denote the marginal on Isom𝒳\operatorname{Isom}{\mathcal{X}} and on Isom𝒳\operatorname{Isom}{{\mathcal{X}}^{\star}} respectively. Let suppμ{\rm supp}\,{\mu^{\star}} denote the support of μ{\mu^{\star}}. The differential entropy of the pair (𝒳,μ)(\partial{{\mathcal{X}}^{\star}},{\mu^{\star}}) is defined by

h(𝒳,μ):=xsuppμμ(x)𝒳logdxνμdνμ(η)𝑑xνμ(η).h(\partial{{\mathcal{X}}^{\star}},{\mu^{\star}}):=\sum_{x\in{\rm supp}\,{\mu^{\star}}}{\mu^{\star}}(x)\int_{\partial{{\mathcal{X}}^{\star}}}\log\frac{dx\nu_{{\mu^{\star}}}}{d\nu_{{\mu^{\star}}}}(\eta)\,dx\nu_{{\mu^{\star}}}(\eta).

In general, it holds that h(𝒳,μ)h(μ)h(\partial{{\mathcal{X}}^{\star}},{\mu^{\star}})\leq h({\mu^{\star}}), and the equality holds if and only if (𝒳,νμ)(\partial{{\mathcal{X}}^{\star}},\nu_{\mu^{\star}}) is a Poisson boundary for the pair (Isom𝒳,μ)(\operatorname{Isom}{{\mathcal{X}}^{\star}},{\mu^{\star}}) (cf. Section 2.3). Let l(𝒳,μ)l({\mathcal{X}},\mu) be the drift associated with a μ\mu-random walk on 𝒳{\mathcal{X}}. Theorem 1.2 is generalized in this setting.

Theorem 1.3.

Let (𝒳,d)({\mathcal{X}},d) and (𝒳,d)({{\mathcal{X}}^{\star}},d^{\star}) be proper roughly geodesic hyperbolic metric spaces with bounded growth at some scale, and 𝚪{\bm{\Gamma}} be a countable subgroup of Isom𝒳×Isom𝒳\operatorname{Isom}{\mathcal{X}}\times\operatorname{Isom}{{\mathcal{X}}^{\star}} with finite exponential growth relative to (𝒳×𝒳,𝐝)({\mathcal{X}}\times{{\mathcal{X}}^{\star}},\bm{d}). If π\pi is a probability measure on 𝚪{\bm{\Gamma}} with finite first moment and non-elementary marginals μ\mu and μ{\mu^{\star}} respectively, then the conditional measure νπη\nu_{\pi}^{\eta} is exact dimensional for νμ\nu_{{\mu^{\star}}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}}. In fact, for νμ\nu_{{\mu^{\star}}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}} and for νπη\nu_{\pi}^{\eta}-almost every 𝛏𝒳×𝒳{\bm{\xi}}\in\partial{\mathcal{X}}\times\partial{{\mathcal{X}}^{\star}},

limr0logνπη(𝑩(𝝃,r))logr=h(π)h(𝒳,μ)l(𝒳,μ).\lim_{r\to 0}\frac{\log\nu_{\pi}^{\eta}(\bm{B}({\bm{\xi}},r))}{\log r}=\frac{h(\pi)-h(\partial{{\mathcal{X}}^{\star}},{\mu^{\star}})}{l({\mathcal{X}},\mu)}.

In particular, the Hausdorff dimension of νπη\nu_{\pi}^{\eta} is computed as for νμ\nu_{{\mu^{\star}}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}},

dimνπη=h(π)h(𝒳,μ)l(𝒳,μ).\dim\nu_{\pi}^{\eta}=\frac{h(\pi)-h(\partial{{\mathcal{X}}^{\star}},{\mu^{\star}})}{l({\mathcal{X}},\mu)}.

For the differential entropy, it holds that h(𝒳,μ)=0h(\partial{{\mathcal{X}}^{\star}},{\mu^{\star}})=0 if and only if (𝒳,νμ)(\partial{{\mathcal{X}}^{\star}},\nu_{{\mu^{\star}}}) is trivial, i.e., νμ\nu_{{\mu^{\star}}} is invariant under the action of 𝚪{\bm{\Gamma}} on 𝒳\partial{{\mathcal{X}}^{\star}}. If μ{\mu^{\star}} is non-elementary, then (𝒳,νμ)(\partial{{\mathcal{X}}^{\star}},\nu_{{\mu^{\star}}}) is non-trivial and h(𝒳,μ)>0h(\partial{{\mathcal{X}}^{\star}},{\mu^{\star}})>0. In the setting of Theorem 1.3, it can be the case that h(π)=h(𝒳,μ)h(\pi)=h(\partial{{\mathcal{X}}^{\star}},{\mu^{\star}}) (see Example 1.5 below). The following result provides a sufficient condition under which h(π)>h(𝒳,μ)h(\pi)>h(\partial{{\mathcal{X}}^{\star}},{\mu^{\star}}), i.e., (𝒳,νμ)(\partial{{\mathcal{X}}^{\star}},\nu_{\mu^{\star}}) is a proper quotient of the Poisson boundary for the pair (𝚪,π)({\bm{\Gamma}},\pi). If this is the case, then the Hausdorff dimension of conditional measures are strictly positive.

Theorem 1.4.

Let Γ\Gamma and Γ{\Gamma^{\star}} be countable subgroups in Isom𝒳\operatorname{Isom}{\mathcal{X}} and in Isom𝒳\operatorname{Isom}{{\mathcal{X}}^{\star}} respectively, and 𝚪:=Γ×Γ{\bm{\Gamma}}:=\Gamma\times{\Gamma^{\star}}. Further let us consider a probability measure π\pi on 𝚪{\bm{\Gamma}} of the following form: For some α(0,1]\alpha\in(0,1],

π=αλ×λ+(1α)π0\pi=\alpha\lambda\times{\lambda^{\star}}+(1-\alpha)\pi_{0}

with non-elementary probability measures λ\lambda and λ{\lambda^{\star}} on Γ\Gamma and on Γ{\Gamma^{\star}} respectively, and a probability measure π0\pi_{0} on 𝚪{\bm{\Gamma}}. It holds that h(π)h(𝒳,μ)>0h(\pi)-h(\partial{{\mathcal{X}}^{\star}},{\mu^{\star}})>0, where μ{\mu^{\star}} is the marginal of π\pi on Γ{\Gamma^{\star}}.

Theorem 1.4 is shown in Theorem 5.4; moreover, if in addition 𝚪=Γ×Γ{\bm{\Gamma}}=\Gamma\times{\Gamma^{\star}} has a finite exponential growth relative to (𝒳×𝒳,𝒅)({\mathcal{X}}\times{{\mathcal{X}}^{\star}},\bm{d}), then for νμ\nu_{{\mu^{\star}}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}}, the Hausdorff dimension of the conditional measure νπη\nu_{\pi}^{\eta} is positive (cf. Theorem 1.3).

Example 1.5.

Let Γ\Gamma be a hyperbolic group and μ\mu be a non-elementary probability measure on Γ\Gamma with finite first moment relative to a word metric. For ρ[0,1]\rho\in[0,1], let

πρ:=ρμ×μ+(1ρ)μdiag,\pi^{\rho}:=\rho\mu\times\mu+(1-\rho)\mu_{{\rm diag}},

where μdiag((x,x)):=μ(x)\mu_{{\rm diag}}((x,x^{\star})):=\mu(x) if x=xx=x^{\star}, and 0 if otherwise. The πρ\pi^{\rho}-random walk on Γ×Γ\Gamma\times\Gamma appears in the study of noise sensitivity problem on groups [BB23, Tan24]. By Theorems 1.2 and 1.4 applied to the case when Γ=Γ\Gamma={\Gamma^{\star}} and μ=μ\mu={\mu^{\star}}, it holds that for all ρ(0,1]\rho\in(0,1],

dimνπρη=h(πρ)h(μ)l(Γ,μ)>0for νμ-almost every ηΓ.\dim\nu_{\pi^{\rho}}^{\eta}=\frac{h(\pi^{\rho})-h(\mu)}{l(\Gamma,\mu)}>0\quad\text{for $\nu_{\mu}$-almost every $\eta\in\partial\Gamma$}.

For ρ=0\rho=0, since h(πρ)=h(μ)h(\pi^{\rho})=h(\mu), it holds that dimνπρη=0\dim\nu_{\pi^{\rho}}^{\eta}=0 for νμ\nu_{\mu}-almost every ηΓ\eta\in\partial\Gamma. Theorem 1.1 shows that for all ρ[0,1]\rho\in[0,1],

dimνπρ=h(πρ)l(Γ,μ).\dim\nu_{\pi^{\rho}}=\frac{h(\pi^{\rho})}{l(\Gamma,\mu)}.

This reproduces [Tan24, Theorem 3.1].

Example 1.6.

This example is not covered by Theorem 1.4. Suppose that 𝚪=Γ×Γ{\bm{\Gamma}}=\Gamma\times{\Gamma^{\star}} for two hyperbolic groups Γ\Gamma and Γ{\Gamma^{\star}}, and that there exists a (non-injective) surjective homomorphism Π:ΓΓ\Pi:\Gamma\to{\Gamma^{\star}}. Let Δ:ΓΓ×Γ\Delta:\Gamma\to\Gamma\times{\Gamma^{\star}} be the diagonal embedding Δ(x)=(x,Π(x))\Delta(x)=(x,\Pi(x)) for xΓx\in\Gamma. For a non-elementary probability measure μ\mu on Γ\Gamma with finite first moment, let π:=Δμ\pi:=\Delta_{\ast}\mu be the pushforward of μ\mu by Δ\Delta. In this case, marginals of π\pi are μ\mu on Γ\Gamma and Πμ\Pi_{\ast}\mu on Γ{\Gamma^{\star}} respectively, where Πμ\Pi_{\ast}\mu is the pushforward of μ\mu by Π\Pi. Applying to Theorem 1.2 with μ=Πμ{\mu^{\star}}=\Pi_{\ast}\mu shows that

dimνπη=h(μ)h(Πμ)l(Γ,μ)for νΠμ-almost every ηΓ.\dim\nu_{\pi}^{\eta}=\frac{h(\mu)-h(\Pi_{\ast}\mu)}{l(\Gamma,\mu)}\quad\text{for $\nu_{\Pi_{\ast}\mu}$-almost every $\eta\in\partial{\Gamma^{\star}}$}.

This follows since h(π)=h(μ)h(\pi)=h(\mu) and h(Γ,Πμ)=h(Πμ)h(\partial{\Gamma^{\star}},\Pi_{\ast}\mu)=h(\Pi_{\ast}\mu). It holds that h(μ)=h(Πμ)h(\mu)=h(\Pi_{\ast}\mu) if Π\Pi is an isomorphism, and it depends on Π\Pi whether a strict inequality h(μ)>h(Πμ)h(\mu)>h(\Pi_{\ast}\mu) holds or not. As a simple explicit example, let Γ=Fm+1\Gamma=F_{m+1} and Γ=Fm{\Gamma^{\star}}=F_{m} be free groups of rank m+1m+1 and mm respectively for m2m\geq 2, equipped with word metrics associated with free bases. Further let Π:Fm+1Fm\Pi:F_{m+1}\to F_{m} be a homomorphism sending xix_{i} to yiy_{i}, i=1,,mi=1,\dots,m, and sending xm+1x_{m+1} to the identity, where {x1,,xm+1}\{x_{1},\dots,x_{m+1}\} denotes the free basis in Fm+1F_{m+1} and {y1,,ym}\{y_{1},\dots,y_{m}\} denotes the free basis in FmF_{m}. For the uniform distribution μ\mu on the symmetrized free basis in Fm+1F_{m+1}, the induced distribution on FmF_{m} defines a simple random walk on FmF_{m} with holding probability 1/(m+1)1/(m+1). A computation yields l(Fm+1,μ)=m/(m+1)l(F_{m+1},\mu)=m/(m+1), l(Fm,Πμ)=(m1)/(m+1)l(F_{m},\Pi_{\ast}\mu)=(m-1)/(m+1),

h(μ)=mm+1log(2m+1)and h(Πμ)=m1m+1log(2m1).h(\mu)=\frac{m}{m+1}\log(2m+1)\quad\text{and }\quad h(\Pi_{\ast}\mu)=\frac{m-1}{m+1}\log(2m-1).

Therefore

dimνπη=log(2m+1)m1mlog(2m1)for νπη-almost every ηFm.\dim\nu_{\pi}^{\eta}=\log(2m+1)-\frac{m-1}{m}\log(2m-1)\quad\text{for $\nu_{\pi}^{\eta}$-almost every $\eta\in\partial F_{m}$}.

Furthermore, Theorem 1.1 shows that

dimνπ=log(2m+1)+1mlog(2m1).\dim\nu_{\pi}=\log(2m+1)+\frac{1}{m}\log(2m-1).

Outlines of proofs

Let us briefly mention the proof of Theorem 1.1 for a product of two hyperbolic groups. For a single hyperbolic group Γ\Gamma with a non-elementary probability measure μ\mu, the corresponding harmonic measure νμ\nu_{\mu} on Γ\partial\Gamma is exact dimensional [Tan19, Theorem 3.8]. Roughly speaking, it boils down to estimate probabilities that for a μ\mu-random walk wnw_{n}, an independent μ\mu-random walk wnw_{n}^{\prime} is around wnw_{n} within distance o(n)o(n) for n+n\in{\mathbb{Z}}_{+}. This leads an estimate of the harmonic measure νμ\nu_{\mu} on the balls B(w,eln)B(w_{\infty},e^{-ln}) where l:=l(Γ,μ)l:=l(\Gamma,\mu). Here the μ\mu-random walk {wn}n+\{w_{n}\}_{n\in{\mathbb{Z}}_{+}} is for a sampling ww_{\infty} in Γ\partial\Gamma according to νμ\nu_{\mu} and the independent μ\mu-random walk {wn}n+\{w_{n}^{\prime}\}_{n\in{\mathbb{Z}}_{+}} is for the estimate νμ(B(w,eln))\nu_{\mu}(B(w_{\infty},e^{-ln})). Since the probability that wnw_{n}^{\prime} is around wnw_{n} within distance o(n)o(n) is eh(μ)n+o(n)e^{-h(\mu)n+o(n)} by the Shannon theorem for random walks, this explains νμ(B(w,eln))=eh(μ)n+o(n)\nu_{\mu}(B(w_{\infty},e^{-ln}))=e^{-h(\mu)n+o(n)}, which is the exact dimensionality of νμ\nu_{\mu} with the right dimension h(μ)/lh(\mu)/l.

The conditional measure νπη\nu_{\pi}^{\eta} for νμ\nu_{\mu^{\star}}-almost every ηΓ\eta\in\partial{\Gamma^{\star}} is the hitting distribution of a conditional process. This is a Markov chain (although the transition probabilities are not group-invariant) and (one of) the methods developed for a single hyperbolic group in [ibid] applies. The asymptotic entropy of this conditional process equals h(π)h(μ)h(\pi)-h({\mu^{\star}}) by the Shannon theorem for the conditional process [Kai00]. The conditional measure νπη\nu_{\pi}^{\eta} is defined on Γ×Γ\partial\Gamma\times\partial{\Gamma^{\star}} but supported on Γ×{η}\partial\Gamma\times\{\eta\} for νμ\nu_{\mu^{\star}}-almost every ηΓ\eta\in\partial{\Gamma^{\star}}. An analogous discussion to the μ\mu-random walk above works and this leads to estimating the νπη\nu_{\pi}^{\eta}-measures on the balls in the boundary B(w,eln)×{η}B(w_{\infty},e^{-ln})\times\{\eta\} for νπη\nu_{\pi}^{\eta}-almost every (w,η)Γ×Γ(w_{\infty},\eta)\in\partial\Gamma\times\partial{\Gamma^{\star}}. In fact, we obtain νπη(B(w,eln)×{η})=e(h(π)h(μ))n+o(n)\nu_{\pi}^{\eta}(B(w_{\infty},e^{-ln})\times\{\eta\})=e^{-(h(\pi)-h({\mu^{\star}}))n+o(n)}, deducing Theorem 1.2.

The harmonic measure νπ\nu_{\pi} on Γ×Γ\partial\Gamma\times\partial{\Gamma^{\star}} is, however, analyzed in a completely different way. First of all it requires to take into account the difference between ll and l{l^{\star}} where l:=l(Γ,μ){l^{\star}}:=l({\Gamma^{\star}},{\mu^{\star}}). If lll\geq{l^{\star}}, then

h(π)h(μ)l+h(μ)lh(π)h(μ)l+h(μ)l,\frac{h(\pi)-h(\mu)}{{l^{\star}}}+\frac{h(\mu)}{l}\leq\frac{h(\pi)-h({\mu^{\star}})}{l}+\frac{h({\mu^{\star}})}{{l^{\star}}},

since h(π)h(μ)+h(μ)h(\pi)\leq h(\mu)+h({\mu^{\star}}), and the inequality can be strict. Since the right hand side of the above inequality is the correct value, the dimension upper bound should use the inequality lll\geq{l^{\star}} whereas the dimension lower bound would not need it. Concerning the dimension upper bound, the Shannon theorem for the conditional process shows that for νμ\nu_{\mu^{\star}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}}, for the conditional process {𝒘n}n+\{\bm{w}_{n}\}_{n\in{\mathbb{Z}}_{+}},

𝐏η([𝒘0,,𝒘n])eh(μ)n+o(n)𝐏([𝒘0,,𝒘n]).{\bf P}^{\eta}([\bm{w}_{0},\dots,\bm{w}_{n}])\leq e^{h({\mu^{\star}})n+o(n)}{\bf P}([\bm{w}_{0},\dots,\bm{w}_{n}]).

In the above, [𝒘0,,𝒘n][\bm{w}_{0},\dots,\bm{w}_{n}] denotes the cylinder set. At this point, we keep track the whole trajectory up to time nn instead of just looking at the position 𝒘n\bm{w}_{n}. The argument here is inspired by [LL23, Section 8] (where they refer to [Fen23] for the idea). The νπη\nu_{\pi}^{\eta} on the balls 𝑩(𝒘,eln)=B(w,eln)×B(w,eln)\bm{B}(\bm{w}_{\infty},e^{-{l^{\star}}n})=B(w_{\infty},e^{-{l^{\star}}n})\times B(w^{\star}_{\infty},e^{-{l^{\star}}n}) estimates by Theorem 1.2,

νπη(𝑩(𝒘,eln))=exp((h(π)h(μ)l)ln+o(n)).\nu_{\pi}^{\eta}\left(\bm{B}(\bm{w}_{\infty},e^{-{l^{\star}}n})\right)=\exp\left(-\left(\frac{h(\pi)-h({\mu^{\star}})}{l}\right){l^{\star}}n+o(n)\right).

Averaging η\eta over B(w,eln)B(w^{\star}_{\infty},e^{-{l^{\star}}n}) deduces the required lower bound (thus upper bound for the dimension) of νπ(𝑩(𝒘,eln))\nu_{\pi}(\bm{B}(\bm{w}_{\infty},e^{-{l^{\star}}n})). In this discussion, it is crucial to use the balls with radii elne^{-{l^{\star}}n} rather than elne^{-ln} (or other scales) since 𝒒(𝒘,𝒘n)=eln+o(n)\bm{q}(\bm{w}_{\infty},\bm{w}_{n})=e^{-{l^{\star}}n+o(n)}, where lll\geq{l^{\star}},

q(w,wn)=eln+o(n)andq(w,wn)=eln+o(n).q(w_{\infty},w_{n})=e^{-ln+o(n)}\quad\text{and}\quad{q^{\star}}(w^{\star}_{\infty},w^{\star}_{n})=e^{-{l^{\star}}n+o(n)}.

Concerning the dimension lower bound, a slight strengthened version for the lower bound in Theorem 1.2 enables us to exploit the naive disintegration formula. Roughly, estimating along the following heuristic can be justified:

νπ(𝑩(𝒘,eln))νπη(B(w,eln)×{η})νμ(B(w,eln)).\nu_{\pi}\left(\bm{B}(\bm{w}_{\infty},e^{-ln})\right)\approx\nu_{\pi}^{\eta}(B(w_{\infty},e^{-ln})\times\{\eta\})\cdot\nu_{\mu^{\star}}\left(B(w^{\star}_{\infty},e^{-ln})\right).

Since this works only for νπ\nu_{\pi} restricted on a large subset, the argument is merely for the upper bound (thus lower bound for the dimension) of νπ\nu_{\pi} (up to a density lemma which is guaranteed by a weak version of the Lebesgue differentiation theorem Lemma 2.2). Thus Theorem 1.2 and the exact dimensionality of νμ\nu_{\mu^{\star}} with h(μ)/lh({\mu^{\star}})/{l^{\star}} conclude the required dimension lower bound on νπ\nu_{\pi}.

The above sketch for hyperbolic groups can be extended to a countable group of isometries acting on a product of two hyperbolic metric spaces in Theorem 4.5. The positive lower bound for h(π)h(𝒳,μ)h(\pi)-h(\partial{{\mathcal{X}}^{\star}},{\mu^{\star}}) in Theorem 1.4 uses the pivotal time technique developed by Gouëzel in [Gou22]. We mention possible extensions of Theorems 1.1 and 4.5 and questions in Remark 4.6.

Organization

Section 2 recalls basics on hyperbolic metric spaces and random walks. Section 3 concerns dimensions of the conditional measures, showing Theorem 1.3 (and thus Theorem 1.2). Section 4 concerns dimensions of the harmonic measures on products of boundaries, showing Theorem 4.5 (and thus Theorem 1.1). Section 5 is about a sufficient condition on a positivity of the dimension for conditional measures, showing Theorem 1.4 in Theorem 5.4.

Notation

We denote by cc, CC, …, constants whose exact values may vary from line to line, and by CδC_{\delta} a constant which depends on the other constant δ\delta to emphasize its dependency. For a real valued sequence {f(n)}n+\{f(n)\}_{n\in{\mathbb{Z}}_{+}} on non-negative integers +{\mathbb{Z}}_{+}, we write f(n)=o(n)f(n)=o(n) if |f(n)|/n0|f(n)|/n\to 0 as nn\to\infty. For a set AA, we denote by A𝖼A^{\sf c} the complement set, and by #A\#A the cardinality.

2. Preliminaries

2.1. Hyperbolic metric spaces

For background, we refer to the original paper by Gromov [Gro87]. For a metric space (𝒳,d)({\mathcal{X}},d), the Gromov product is defined by

(x|y)z:=12(d(x,z)+d(z,y)d(x,y))for x,y,z𝒳.(x|y)_{z}:=\frac{1}{2}(d(x,z)+d(z,y)-d(x,y))\quad\text{for $x,y,z\in{\mathcal{X}}$}.

A metric space (𝒳,d)({\mathcal{X}},d) is δ\delta-hyperbolic for a non-negative real δ+\delta\in{\mathbb{R}}_{+} if it holds that

(x|y)wmin{(x|z)w,(z|y)w}δfor all x,y,z,w𝒳.(x|y)_{w}\geq\min\big{\{}(x|z)_{w},(z|y)_{w}\big{\}}-\delta\quad\text{for all $x,y,z,w\in{\mathcal{X}}$}. (2.1)

It is called hyperbolic if it is δ\delta-hyperbolic for some δ+\delta\in{\mathbb{R}}_{+}. A map γ:I𝒳\gamma:I\to{\mathcal{X}} from an interval II in {\mathbb{R}} to 𝒳{\mathcal{X}} is called a CC-rough geodesic for C+C\in{\mathbb{R}}_{+} if |d(γ(s),γ(t))|ts||C|d(\gamma(s),\gamma(t))-|t-s||\leq C for all s,tIs,t\in I. Further a map γ:I𝒳\gamma:I\to{\mathcal{X}} is called a CC-rough geodesic ray in the case when I=[0,)I=[0,\infty). A metric space is called CC-roughly geodesic for C+C\in{\mathbb{R}}_{+} if for all pairs of points x,y𝒳x,y\in{\mathcal{X}} there exists a CC-rough geodesic γ:[a,b]𝒳\gamma:[a,b]\to{\mathcal{X}} such that γ(a)=x\gamma(a)=x and γ(b)=y\gamma(b)=y. In this terminology, a metric space is called geodesic if it is 0-roughly geodesic. A graph endowed with a path metric of unit edge length (e.g., a Cayley graph) is also considered as a geodesic metric space by using intervals in the integers {\mathbb{Z}} in the definition. Let us simply call a metric space roughly geodesic if it is CC-roughly geodesic for some C+C\in{\mathbb{R}}_{+}. For a hyperbolic group Γ\Gamma equipped with a left invariant hyperbolic metric dd quasi-isometric to a word metric, (Γ,d)(\Gamma,d) is roughly geodesic (cf. [BS00, Proposition 5.6] and [BHM11, Theorem 2.2]). A metric space (𝒳,d)({\mathcal{X}},d) is proper if for all x𝒳x\in{\mathcal{X}} and all r+r\in{\mathbb{R}}_{+}, the ball B(x,r):={y𝒳:d(x,y)<r}B(x,r):=\big{\{}y\in{\mathcal{X}}\ :\ d(x,y)<r\big{\}} is relatively compact.

For a hyperbolic metric space (𝒳,d)({\mathcal{X}},d), the (Gromov) boundary 𝒳\partial{\mathcal{X}} is defined as the set of equivalence classes of divergent sequences in 𝒳{\mathcal{X}}. Let us fix a point o𝒳o\in{\mathcal{X}}. Further let q(x,y):=exp((x|y)o)q(x,y):=\exp(-(x|y)_{o}) for x,y𝒳x,y\in{\mathcal{X}} and q(x,y):=0q(x,y):=0 if x=yx=y. Since the space is δ\delta-hyperbolic for some δ+\delta\in{\mathbb{R}}_{+}, it holds that

q(x,y)eδmax{q(x,z),q(z,y)}for x,y,z𝒳.q(x,y)\leq e^{\delta}\max\big{\{}q(x,z),q(z,y)\big{\}}\quad\text{for $x,y,z\in{\mathcal{X}}$}. (2.2)

A sequence {xn}n+\{x_{n}\}_{n\in{\mathbb{Z}}_{+}} in 𝒳{\mathcal{X}} is called divergent if it is a Cauchy sequence with respect to qq. Two sequences {xn}n+\{x_{n}\}_{n\in{\mathbb{Z}}_{+}} and {yn}n+\{y_{n}\}_{n\in{\mathbb{Z}}_{+}} are equivalent if q(xn,ym)0q(x_{n},y_{m})\to 0 as n,mn,m\to\infty. It is indeed an equivalence relation in the set of divergence sequences by (2.2).

For ξ𝒳𝒳\xi\in{\mathcal{X}}\cup\partial{\mathcal{X}}, let us write ξ=[{xn}n+]\xi=[\{x_{n}\}_{n\in{\mathbb{Z}}_{+}}] for a divergent sequence {xn}n+\{x_{n}\}_{n\in{\mathbb{Z}}_{+}} which represents ξ\xi if ξ𝒳\xi\in\partial{\mathcal{X}}, or for the constant sequence xn=ξx_{n}=\xi for all n+n\in{\mathbb{Z}}_{+} if ξ𝒳\xi\in{\mathcal{X}}. The Gromov product is extended to 𝒳𝒳{\mathcal{X}}\cup\partial{\mathcal{X}} by

(ξ|η)o:=inf{lim infn,m(xn|ym)o:ξ=[{xn}n+],η=[{ym}m+]}.(\xi|\eta)_{o}:=\inf\Big{\{}\liminf_{n,m\to\infty}(x_{n}|y_{m})_{o}\ :\ \xi=[\{x_{n}\}_{n\in{\mathbb{Z}}_{+}}],\ \eta=[\{y_{m}\}_{m\in{\mathbb{Z}}_{+}}]\Big{\}}.

For a δ\delta-hyperbolic space, the extended Gromov product satisfies (2.1) for x,y,z𝒳𝒳x,y,z\in{\mathcal{X}}\cup\partial{\mathcal{X}} and w=ow=o. Let us extend qq on 𝒳{\mathcal{X}} to 𝒳𝒳{\mathcal{X}}\cup\partial{\mathcal{X}} and call it the quasi-metric:

q(ξ,η):=exp((ξ|η)o)if ξη,andq(ξ,η):=0if ξ=η,for ξ,η𝒳𝒳.q(\xi,\eta):=\exp(-(\xi|\eta)_{o})\quad\text{if $\xi\neq\eta$},\quad\text{and}\quad q(\xi,\eta):=0\quad\text{if $\xi=\eta$},\quad\text{for $\xi,\eta\in{\mathcal{X}}\cup\partial{\mathcal{X}}$}.

It is known that there exists an ε0>0\varepsilon_{0}>0 such that for all 0<ε<ε00<\varepsilon<\varepsilon_{0} the power qεq^{\varepsilon} is bi-Lipschitz equivalent to a genuine metric. However, the quasi-metric qq is used to define balls and other notions related to metrics without introducing an additional parameter ε\varepsilon. The space 𝒳𝒳{\mathcal{X}}\cup\partial{\mathcal{X}} is equipped with the topology defined from the (quasi-)metric. If 𝒳{\mathcal{X}} is proper, then 𝒳𝒳{\mathcal{X}}\cup\partial{\mathcal{X}} is a compact metrizable space. If in addition 𝒳{\mathcal{X}} is CC-roughly geodesic for some C+C\in{\mathbb{R}}_{+}, then for every ξ𝒳\xi\in\partial{\mathcal{X}} and every x𝒳x\in{\mathcal{X}} there exists a CC-rough geodesic ray γ\gamma from xx converging to ξ\xi, i.e., q(γ(t),ξ)0q(\gamma(t),\xi)\to 0 as tt\to\infty in 𝒳𝒳{\mathcal{X}}\cup\partial{\mathcal{X}} (cf. [BS00, Proposition 5.2]). Henceforth it is assumed that 𝒳{\mathcal{X}} is a proper roughly geodesic hyperbolic metric space.

Let us denote the open ball of radius r+r\in{\mathbb{R}}_{+} centered at ξ𝒳𝒳\xi\in{\mathcal{X}}\cup\partial{\mathcal{X}} in 𝒳𝒳{\mathcal{X}}\cup\partial{\mathcal{X}} by

B(ξ,r):={η𝒳𝒳:q(ξ,η)<r}.B(\xi,r):=\Big{\{}\eta\in{\mathcal{X}}\cup\partial{\mathcal{X}}\ :\ q(\xi,\eta)<r\Big{\}}.

The shadow (seen from oo) at x𝒳x\in{\mathcal{X}} with thickness R+R\in{\mathbb{R}}_{+} is defined by

𝒪(x,R):={η𝒳:(o|η)x<R}.{\mathcal{O}}(x,R):=\Big{\{}\eta\in\partial{\mathcal{X}}\ :\ (o|\eta)_{x}<R\Big{\}}.

The following is used to compare shadows with balls. For each T>0T>0, there exist constants R0,C>0R_{0},C>0 such that for all R>R0R>R_{0}, all ξ𝒳\xi\in\partial{\mathcal{X}} and all x𝒳x\in{\mathcal{X}} with (o|ξ)xT(o|\xi)_{x}\leq T,

B(ξ,C1ed(o,x)+R)𝒳𝒪(x,R)B(ξ,Ced(o,x)+R)𝒳.B(\xi,C^{-1}e^{-d(o,x)+R})\cap\partial{\mathcal{X}}\subset{\mathcal{O}}(x,R)\subset B(\xi,Ce^{-d(o,x)+R})\cap\partial{\mathcal{X}}. (2.3)

For another such hyperbolic metric space (𝒳,d)({{\mathcal{X}}^{\star}},d^{\star}) with base point oo^{\star}, let q{q^{\star}} denote the quasi-metric in 𝒳\partial{{\mathcal{X}}^{\star}}. In the product space, for 𝝃i=(ξi,ηi)(𝒳𝒳)×(𝒳𝒳){\bm{\xi}}_{i}=(\xi_{i},\eta_{i})\in\left({\mathcal{X}}\cup\partial{\mathcal{X}}\right)\times\left({{\mathcal{X}}^{\star}}\cup\partial{{\mathcal{X}}^{\star}}\right) and i=1,2i=1,2, let

𝒒(𝝃1,𝝃2):=max{q(ξ1,ξ2),q(η1,η2)}.\bm{q}({\bm{\xi}}_{1},{\bm{\xi}}_{2}):=\max\big{\{}q(\xi_{1},\xi_{2}),{q^{\star}}(\eta_{1},\eta_{2})\big{\}}.

By (2.2) extended to 𝒳𝒳{\mathcal{X}}\cup\partial{\mathcal{X}} and 𝒳𝒳{{\mathcal{X}}^{\star}}\cup\partial{{\mathcal{X}}^{\star}}, there exists a constant C:=C𝒒>0C:=C_{\bm{q}}>0 such that

𝒒(𝒙,𝒚)Cmax{𝒒(𝒙,𝒛),𝒒(𝒛,𝒚)}for all 𝒙,𝒚,𝒛(𝒳𝒳)×(𝒳𝒳).\bm{q}(\bm{x},\bm{y})\leq C\max\{\bm{q}(\bm{x},\bm{z}),\bm{q}(\bm{z},\bm{y})\}\quad\text{for all $\bm{x},\bm{y},\bm{z}\in\left({\mathcal{X}}\cup\partial{\mathcal{X}}\right)\times\left({{\mathcal{X}}^{\star}}\cup\partial{{\mathcal{X}}^{\star}}\right)$}.

Let 𝑩(𝝃,r)\bm{B}({\bm{\xi}},r) denote the ball in (𝒳𝒳)×(𝒳𝒳)\left({\mathcal{X}}\cup\partial{\mathcal{X}}\right)\times\left({{\mathcal{X}}^{\star}}\cup\partial{{\mathcal{X}}^{\star}}\right) with respect to 𝒒\bm{q}.

2.2. Hausdorff dimensions

Let (,q)({\mathcal{M}},q) be a compact metrizable space {\mathcal{M}} with a quasi-metric qq. It is basically intended as (𝒳,q)(\partial{\mathcal{X}},q) or (𝒳×𝒳,𝒒)(\partial{\mathcal{X}}\times\partial{{\mathcal{X}}^{\star}},\bm{q}). For a set EE in {\mathcal{M}}, let dimE\dim E denote the Hausdorff dimension of EE with respect to the quasi-metric qq. The definition is recalled briefly. Let |E|:=sup{q(ξ,η):ξ,ηE}|E|:=\sup\{q(\xi,\eta)\ :\ \xi,\eta\in E\}. For all α,Δ+\alpha,\Delta\in{\mathbb{R}}_{+} with Δ>0\Delta>0, let

Δα(E):=inf{i=0|Ei|α:Ei=0Eiand|Ei|Δ}.{\mathcal{H}}^{\alpha}_{\Delta}(E):=\inf\Big{\{}\sum_{i=0}^{\infty}|E_{i}|^{\alpha}\ :\ E\subset\bigcup_{i=0}^{\infty}E_{i}\ \text{and}\ |E_{i}|\leq\Delta\Big{\}}.

The α\alpha-dimensional Hausdorff measure of a set EE is defined by

α(E):=limΔ0Δα(E)=supΔ>0Δα(E).{\mathcal{H}}^{\alpha}(E):=\lim_{\Delta\to 0}{\mathcal{H}}^{\alpha}_{\Delta}(E)=\sup_{\Delta>0}{\mathcal{H}}^{\alpha}_{\Delta}(E).

Moreover the Hausdorff dimension of a set EE is defined by

dimE:=sup{α0:α(E)>0}=inf{α0:α(E)=0}.\dim E:=\sup\Big{\{}\alpha\geq 0\ :\ {\mathcal{H}}^{\alpha}(E)>0\Big{\}}=\inf\Big{\{}\alpha\geq 0\ :\ {\mathcal{H}}^{\alpha}(E)=0\Big{\}}.

Let ν\nu be a Borel probability measure on 𝒳\partial{\mathcal{X}}. The upper Hausdorff dimension of ν\nu is

dim¯ν:=inf{dimE:E is Borel and ν(E)=0},\overline{\dim}\,\nu:=\inf\Big{\{}\dim E\ :\ \text{$E$ is Borel and $\nu({\mathcal{M}}\setminus E)=0$}\Big{\}},

and the lower Hausdorff dimension of ν\nu is

dim¯ν:=inf{dimE:E is Borel and ν(E)>0}.\underline{\dim}\,\nu:=\inf\Big{\{}\dim E\ :\ \text{$E$ is Borel and $\nu(E)>0$}\Big{\}}.

If the upper and lower Hausdorff dimensions of ν\nu coincide, then the common value is called the Hausdorff dimension of ν\nu and is denoted by dimν\dim\,\nu. The following is a fundamental lemma which relates pointwise behaviors of a measure to Hausdorff dimensions. This is called the Billingsley lemma (in the case of Euclidean spaces).

Lemma 2.1 (cf. Section 8.7 in [Hei01]).

For every Borel probability measure ν\nu on {\mathcal{M}}, if for α1,α2+\alpha_{1},\alpha_{2}\in{\mathbb{R}}_{+},

α1lim infr0logν(B(ξ,r))logrα2for ν-almost every ξ,\alpha_{1}\leq\liminf_{r\to 0}\frac{\log\nu(B(\xi,r))}{\log r}\leq\alpha_{2}\quad\text{for $\nu$-almost every $\xi\in{\mathcal{M}}$},

then α1dim¯να2\alpha_{1}\leq\overline{\dim}\,\nu\leq\alpha_{2}.

It is deduced that

dim¯ν=supν-a.e. ξlim infr0logν(B(ξ,r))logranddim¯ν=infν-a.e. ξlim infr0logν(B(ξ,r))logr.\overline{\dim}\,\nu=\sup_{\text{$\nu$-a.e.\ $\xi$}}\liminf_{r\to 0}\frac{\log\nu(B(\xi,r))}{\log r}\quad\text{and}\quad\underline{\dim}\,\nu=\inf_{\text{$\nu$-a.e.\ $\xi$}}\liminf_{r\to 0}\frac{\log\nu(B(\xi,r))}{\log r}.

In the above, supν-a.e. ξ\sup_{\text{$\nu$-a.e.\ $\xi$}} and infν-a.e. ξ\inf_{\text{$\nu$-a.e.\ $\xi$}} denote the essential supremum and the essential infimum relative to ν\nu respectively. A Borel probability measure ν\nu on {\mathcal{M}} is exact dimensional if the following limit exists and is constant ν\nu-almost everywhere on {\mathcal{M}}:

limr0logν(B(ξ,r))logr.\lim_{r\to 0}\frac{\log\nu(B(\xi,r))}{\log r}.

In that case, the Hausdorff dimension of ν\nu exists and equals the constant.

A metric space (𝒳,d)({\mathcal{X}},d) is called bounded growth at some scale if there exist constants r,R+r,R\in{\mathbb{R}}_{+} with 0<r<R0<r<R and N+N\in{\mathbb{Z}}_{+} such that every open ball of radius RR is covered by at most NN open balls of radius rr. The examples include Gromov hyperbolic Riemannian manifolds whose sectional curvature is uniformly bounded from below and from above and Cayley graphs of hyperbolic groups.

Lemma 2.2.

Let (𝒳,d)({\mathcal{X}},d) and (𝒳,d)({{\mathcal{X}}^{\star}},d^{\star}) be hyperbolic metric spaces with bounded growth at some scale. There exists a constant L1L\geq 1 such that the following holds for every Borel probability measure ν\nu on 𝒳×𝒳\partial{\mathcal{X}}\times\partial{{\mathcal{X}}^{\star}} and for every Borel set FF in 𝒳×𝒳\partial{\mathcal{X}}\times\partial{{\mathcal{X}}^{\star}} with ν(F)>0\nu(F)>0. For ν\nu-almost every 𝛏F{\bm{\xi}}\in F, there exists a constant r(𝛏)>0r({\bm{\xi}})>0 such that for every r(0,r(𝛏))r\in(0,r({\bm{\xi}})),

ν(F𝑩(𝝃,Lr))910ν(𝑩(𝝃,r)).\nu(F\cap\bm{B}({\bm{\xi}},Lr))\geq\frac{9}{10}\nu(\bm{B}({\bm{\xi}},r)).
Proof.

The assumption on (𝒳,d)({\mathcal{X}},d) implies that for every α(0,1)\alpha\in(0,1) there exists a bi-Lipschitz embedding ff from (𝒳,qα)(\partial{\mathcal{X}},q^{\alpha}) to some finite dimensional standard Euclidean space (𝙴𝟶,𝙴𝟶)({\tt E^{0}},\|\cdot\|_{{\tt E^{0}}}) (cf. [BS00, Theorem 9.2] and [Ass83, 2.6. Proposition]). More precisely, there exists a constant L01L^{0}\geq 1 such that for all ξ,η𝒳\xi,\eta\in\partial{\mathcal{X}},

(1/L0)q(ξ,η)αf(ξ)f(η)𝙴0L0q(ξ,η)α.(1/L^{0})q(\xi,\eta)^{\alpha}\leq\|f(\xi)-f(\eta)\|_{{\tt E}_{0}}\leq L^{0}q(\xi,\eta)^{\alpha}.

Similarly, for (𝒳,d)({{\mathcal{X}}^{\star}},d^{\star}) there exists a bi-Lipschitz embedding ff^{\star} from (𝒳,qα)(\partial{{\mathcal{X}}^{\star}},{q^{\star}}^{\alpha}) into some Euclidean space (𝙴,𝙴)({\tt E^{\star}},\|\cdot\|_{{\tt E^{\star}}}) with a Lipschitz constant L1L^{\star}\geq 1. Let

𝒇:𝒳×𝒳𝙴:=𝙴𝟶×𝙴,(ξ,η)(f(ξ),f(η)).{\bm{f}}:\partial{\mathcal{X}}\times\partial{{\mathcal{X}}^{\star}}\to{\tt E}:={\tt E^{0}}\times{\tt E^{\star}},\quad(\xi,\eta)\mapsto(f(\xi),f^{\star}(\eta)).

The map 𝒇{\bm{f}} is a homeomorphism onto its image since 𝒳×𝒳\partial{\mathcal{X}}\times\partial{{\mathcal{X}}^{\star}} is compact. The product space 𝙴{\tt E} is endowed with the maximum norm 𝙴\|\cdot\|_{{\tt E}} of the factors. Let B𝙴(𝒗,r)B_{{\tt E}}(\bm{v},r) denote the ball in 𝙴{\tt E} with respect to the norm. It holds that

(1/L)𝒒(𝝃,𝜼)α𝒇(𝝃)𝒇(𝜼)𝙴L𝒒(𝝃,𝜼)αfor 𝝃,𝜼𝒳×𝒳,(1/L)\bm{q}({\bm{\xi}},{\bm{\eta}})^{\alpha}\leq\|{\bm{f}}({\bm{\xi}})-{\bm{f}}({\bm{\eta}})\|_{{\tt E}}\leq L\bm{q}({\bm{\xi}},{\bm{\eta}})^{\alpha}\quad\text{for ${\bm{\xi}},{\bm{\eta}}\in\partial{\mathcal{X}}\times\partial{{\mathcal{X}}^{\star}}$}, (2.4)

where L:=max{L0,L}L:=\max\{L^{0},L^{\star}\}. The pushforward 𝒇ν{\bm{f}}_{\ast}\nu satisfies that 𝒇ν(B𝙴(𝒇(𝝃),r))>0{\bm{f}}_{\ast}\nu(B_{{\tt E}}({\bm{f}}({\bm{\xi}}),r))>0 for all r>0r>0 and for ν\nu-almost all 𝝃F{\bm{\xi}}\in F. This follows since ν(F)>0\nu(F)>0 and the intersection of FF and the support of ν\nu has a positive ν\nu-measure, it holds that ν(𝑩(𝝃,r))>0\nu(\bm{B}({\bm{\xi}},r))>0 for all r>0r>0 and for ν\nu-almost every 𝝃F{\bm{\xi}}\in F. The Lebesgue differentiation theorem on 𝒇ν{\bm{f}}_{\ast}\nu yields

limr0𝒇ν(𝒇(F)B𝙴(𝒇(𝝃),r))𝒇ν(B𝙴(𝒇(𝝃),r))=1for ν-almost every 𝝃F.\lim_{r\to 0}\frac{{\bm{f}}_{\ast}\nu({\bm{f}}(F)\cap B_{{\tt E}}({\bm{f}}({\bm{\xi}}),r))}{{\bm{f}}_{\ast}\nu(B_{{\tt E}}({\bm{f}}({\bm{\xi}}),r))}=1\quad\text{for $\nu$-almost every ${\bm{\xi}}\in F$}.

By (2.4), it holds that

lim infr0ν(FB(𝝃,(Lr)1/α))ν(B(𝝃,(r/L)1/α))1for ν-almost every 𝝃F.\liminf_{r\to 0}\frac{\nu(F\cap B({\bm{\xi}},(Lr)^{1/{\alpha}}))}{\nu(B({\bm{\xi}},(r/L)^{1/{\alpha}}))}\geq 1\quad\text{for $\nu$-almost every ${\bm{\xi}}\in F$}.

Hence for ν\nu-almost every 𝝃F{\bm{\xi}}\in F there exists some r(𝝃)>0r({\bm{\xi}})>0 such that

ν(FB(𝝃,L2/αr))910ν(B(𝝃,r))for all r(0,r(𝝃)).\nu(F\cap B({\bm{\xi}},L^{2/\alpha}r))\geq\frac{9}{10}\nu(B({\bm{\xi}},r))\quad\text{for all $r\in(0,r({\bm{\xi}}))$}.

Shifting the constant L2/αL^{2/\alpha} to LL deduces the claim. ∎

2.3. Random walks

Let 𝚪{\bm{\Gamma}} be a countable group. Further let Ω:=𝚪+\Omega:={\bm{\Gamma}}^{{\mathbb{Z}}_{+}} be the product space endowed with the σ\sigma-algebra {\mathcal{F}} generated by cylinder sets. For a probability measure π\pi on 𝚪{\bm{\Gamma}}, let π+\pi^{{\mathbb{Z}}_{+}} be the product measure on 𝚪+{\bm{\Gamma}}^{{\mathbb{Z}}_{+}}. Let us define the map w:𝚪+Ωw:{\bm{\Gamma}}^{{\mathbb{Z}}_{+}}\to\Omega, {𝒙n}n+{𝒘n}n+\{\bm{x}_{n}\}_{n\in{\mathbb{Z}}_{+}}\mapsto\{\bm{w}_{n}\}_{n\in{\mathbb{Z}}_{+}} where 𝒘0:=id\bm{w}_{0}:=\operatorname{{\rm id}} (the identity element) and

𝒘n:=𝒙1𝒙nfor n=1,2,.\bm{w}_{n}:=\bm{x}_{1}\cdots\bm{x}_{n}\quad\text{for $n=1,2,\dots$}.

The pushforward of π+\pi^{{\mathbb{Z}}_{+}} by the map ww is denoted by 𝐏{\bf P}. The probability space (Ω,,𝐏)(\Omega,{\mathcal{F}},{\bf P}) is a standard probability space; this is the most basic space in the following discussion. The maps Ω𝚪\Omega\to{\bm{\Gamma}}, {𝒘n}n+𝒘n\{\bm{w}_{n}\}_{n\in{\mathbb{Z}}_{+}}\mapsto\bm{w}_{n} defines a Markov chain {𝒘n}n+\{\bm{w}_{n}\}_{n\in{\mathbb{Z}}_{+}} called a π\pi-random walk starting from id\operatorname{{\rm id}}.

For a hyperbolic metric space (𝒳,d)({\mathcal{X}},d), let Isom𝒳\operatorname{Isom}{\mathcal{X}} denote the isometry group. A probability measure μ\mu (with a countable support) on Isom𝒳\operatorname{Isom}{\mathcal{X}} is called non-elementary if the group generated by the support of μ\mu (as a group) contains a free group of rank 22.

Let (𝒳,d)({\mathcal{X}},d) and (𝒳,d)({{\mathcal{X}}^{\star}},d^{\star}) be hyperbolic metric spaces, and 𝚪{\bm{\Gamma}} be a countable subgroup of Isom𝒳×Isom𝒳\operatorname{Isom}{\mathcal{X}}\times\operatorname{Isom}{{\mathcal{X}}^{\star}}. Further let π\pi be a probability measure on 𝚪{\bm{\Gamma}} such that the pushforwards μ\mu and μ{\mu^{\star}} by the projections from Isom𝒳×Isom𝒳\operatorname{Isom}{\mathcal{X}}\times\operatorname{Isom}{{\mathcal{X}}^{\star}} to Isom𝒳\operatorname{Isom}{\mathcal{X}} and to Isom𝒳\operatorname{Isom}{{\mathcal{X}}^{\star}} respectively are non-elementary. In this setting, a π\pi-random walk {𝒘n}n+\{\bm{w}_{n}\}_{n\in{\mathbb{Z}}_{+}} starting from id\operatorname{{\rm id}} yields by letting 𝒘n=(wn,wn)\bm{w}_{n}=(w_{n},w^{\star}_{n}), a μ\mu-random walk {wn}n+\{w_{n}\}_{n\in{\mathbb{Z}}_{+}} with w0=idw_{0}=\operatorname{{\rm id}} and a μ{\mu^{\star}}-random walk {wn}n+\{w^{\star}_{n}\}_{n\in{\mathbb{Z}}_{+}} with w0=idw^{\star}_{0}=\operatorname{{\rm id}}. For fixed base points o𝒳o\in{\mathcal{X}} and o𝒳o^{\star}\in{{\mathcal{X}}^{\star}}, let

𝒛n:=(zn,zn),where zn:=wno and zn:=wno for 𝒘n=(wn,wn).\bm{z}_{n}:=(z_{n},z^{\star}_{n}),\quad\text{where $z_{n}:=w_{n}\cdot o$ and $z^{\star}_{n}:=w^{\star}_{n}\cdot o^{\star}$ for $\bm{w}_{n}=(w_{n},w^{\star}_{n})$}.

The assumption that μ\mu and μ{\mu^{\star}} are non-elementary implies that 𝐏{\bf P}-almost surely there exist z𝒳z_{\infty}\in\partial{\mathcal{X}} and z𝒳z^{\star}_{\infty}\in\partial{{\mathcal{X}}^{\star}} such that znzz_{n}\to z_{\infty} in 𝒳𝒳{\mathcal{X}}\cup\partial{\mathcal{X}} and znzz^{\star}_{n}\to z^{\star}_{\infty} in 𝒳𝒳{{\mathcal{X}}^{\star}}\cup\partial{{\mathcal{X}}^{\star}} as nn\to\infty respectively. Let νπ\nu_{\pi} be the distribution of (z,z)(z_{\infty},z^{\star}_{\infty}) on 𝒳×𝒳\partial{\mathcal{X}}\times\partial{{\mathcal{X}}^{\star}}, and νμ\nu_{\mu} and νμ\nu_{{\mu^{\star}}} be the distributions of zz_{\infty} and of zz^{\star}_{\infty} respectively. The probability measure νπ\nu_{\pi} is called the harmonic measure for the π\pi-random walk. Similarly, νμ\nu_{\mu} and νμ\nu_{{\mu^{\star}}} are called the harmonic measures for the μ\mu-random walk and for the μ{\mu^{\star}}-random walk respectively. Let us denote the measurable map by

𝐛𝐧𝐝=(bnd,bnd):Ω𝒳×𝒳,𝒘𝒛:=(z,z).{\bf bnd}=({\rm bnd},{\rm bnd}^{\star}):\Omega\to\partial{\mathcal{X}}\times\partial{{\mathcal{X}}^{\star}},\quad\bm{w}\mapsto\bm{z}_{\infty}:=(z_{\infty},z^{\star}_{\infty}).

The harmonic measures νπ\nu_{\pi}, νμ\nu_{\mu} and νμ\nu_{{\mu^{\star}}} are obtained as the pushforwards of 𝐏{\bf P} by 𝐛𝐧𝐝{\bf bnd}, by bnd{\rm bnd} and by bnd{\rm bnd}^{\star} respectively.

The group Isom𝒳×Isom𝒳\operatorname{Isom}{\mathcal{X}}\times\operatorname{Isom}{{\mathcal{X}}^{\star}} acts on 𝒳×𝒳\partial{\mathcal{X}}\times\partial{{\mathcal{X}}^{\star}} coordinatewise and 𝚪{\bm{\Gamma}} acts on 𝒳×𝒳\partial{\mathcal{X}}\times\partial{{\mathcal{X}}^{\star}}. It satisfies that νπ\nu_{\pi} is π\pi-stationary, i.e.,

νπ=𝒙𝚪π(𝒙)𝒙νπwhere 𝒙νπ=νπ𝒙1.\nu_{\pi}=\sum_{\bm{x}\in{\bm{\Gamma}}}\pi(\bm{x})\bm{x}\nu_{\pi}\quad\text{where $\bm{x}\nu_{\pi}=\nu_{\pi}\circ\bm{x}^{-1}$}.

Similarly, 𝚪{\bm{\Gamma}} acts on 𝒳\partial{\mathcal{X}} and on 𝒳\partial{{\mathcal{X}}^{\star}} through the projection from Isom𝒳×Isom𝒳\operatorname{Isom}{\mathcal{X}}\times\operatorname{Isom}{{\mathcal{X}}^{\star}} to each one of factors, and thus νμ\nu_{\mu} and νμ\nu_{{\mu^{\star}}} are also π\pi-stationary, i.e.,

νμ=𝒙𝚪π(𝒙)𝒙νμandνμ=𝒙𝚪π(𝒙)𝒙νμ.\nu_{\mu}=\sum_{\bm{x}\in{\bm{\Gamma}}}\pi(\bm{x})\bm{x}\nu_{\mu}\quad\text{and}\quad\nu_{{\mu^{\star}}}=\sum_{\bm{x}\in{\bm{\Gamma}}}\pi(\bm{x})\bm{x}\nu_{{\mu^{\star}}}. (2.5)

Since μ\mu and μ{\mu^{\star}} are marginals of π\pi, these further lead to

νμ=xsuppμμ(x)xνμandνμ=xsuppμμ(x)xνμ.\nu_{\mu}=\sum_{x\in{\rm supp}\,\mu}\mu(x)x\nu_{\mu}\quad\text{and}\quad\nu_{{\mu^{\star}}}=\sum_{x^{\star}\in{\rm supp}\,{\mu^{\star}}}{\mu^{\star}}(x^{\star})x^{\star}\nu_{{\mu^{\star}}}.

Let us define the metric on 𝒳×𝒳{\mathcal{X}}\times{{\mathcal{X}}^{\star}} by

𝒅(𝒛1,𝒛2):=max{d(z1,z2),d(z1,z2)}for 𝒛i=(zi,zi)𝒳×𝒳 and i=1,2.\bm{d}(\bm{z}_{1},\bm{z}_{2}):=\max\big{\{}d(z_{1},z_{2}),d^{\star}(z^{\star}_{1},z^{\star}_{2})\big{\}}\quad\text{for $\bm{z}_{i}=(z_{i},z^{\star}_{i})\in{\mathcal{X}}\times{{\mathcal{X}}^{\star}}$ and $i=1,2$}.

A probability measure π\pi on 𝚪<Isom𝒳×Isom𝒳{\bm{\Gamma}}<\operatorname{Isom}{\mathcal{X}}\times\operatorname{Isom}{{\mathcal{X}}^{\star}} has a finite first moment if

𝒙𝚪𝒅(𝒐,𝒙𝒐)π(𝒙)<,where 𝒐=(o,o).\sum_{\bm{x}\in{\bm{\Gamma}}}\bm{d}(\bm{o},\bm{x}\cdot\bm{o})\pi(\bm{x})<\infty,\quad\text{where $\bm{o}=(o,o^{\star})$}.

This condition is independent of the choice of the point 𝒐=(o,o)\bm{o}=(o,o^{\star}). Let us assume that π\pi has a finite first moment. The Kingman subadditive ergodic theorem implies that the following limits exist and are constant 𝐏{\bf P}-almost everywhere:

l(𝒳,μ):=limn1nd(o,zn)andl(𝒳,μ):=limn1nd(o,zn).l({\mathcal{X}},\mu):=\lim_{n\to\infty}\frac{1}{n}d(o,z_{n})\quad\text{and}\quad l({{\mathcal{X}}^{\star}},{\mu^{\star}}):=\lim_{n\to\infty}\frac{1}{n}d^{\star}(o^{\star},z^{\star}_{n}).

The limits l(𝒳,μ)l({\mathcal{X}},\mu) and l(𝒳,μ)l({{\mathcal{X}}^{\star}},{\mu^{\star}}) are called the drifts of {zn}n+\{z_{n}\}_{n\in{\mathbb{Z}}_{+}} and {zn}n+\{z^{\star}_{n}\}_{n\in{\mathbb{Z}}_{+}} respectively. In the case when μ\mu and μ{\mu^{\star}} are non-elementary, then l(𝒳,μ)>0l({\mathcal{X}},\mu)>0 and l(𝒳,μ)>0l({{\mathcal{X}}^{\star}},{\mu^{\star}})>0 ([Kai00, Theorem 7.3], and see also [Gou22, Theorem 1.1] for a more recent account).

2.4. Conditional processes and their entropies

If π\pi has a finite first moment, then the entropy H(π):=𝒙𝚪π(𝒙)logπ(𝒙)H(\pi):=-\sum_{\bm{x}\in{\bm{\Gamma}}}\pi(\bm{x})\log\pi(\bm{x}) is finite [Der86, Section VII, B]. The Shannon theorem for random walks says that for such π\pi, the following limit exists and is constant 𝐏{\bf P}-almost everywhere:

h(π):=limn1nlogπn(𝒘n).h(\pi):=\lim_{n\to\infty}-\frac{1}{n}\log\pi_{n}(\bm{w}_{n}). (2.6)

See [KV83, Theorem 2.1] and [Der80, Section IV]. The limit h(π)h(\pi) is called the asymptotic entropy for π\pi-random walk. Let h(μ)h(\mu) and h(μ)h({\mu^{\star}}) be the asymptotic entropies for μ\mu-random walk and μ{\mu^{\star}}-random walk respectively; they exist and are defined in the same way. We will also use a conditional version of the notion. First we introduce a conditional process and then define the conditional entropy.

Recall that (Ω,,𝐏)(\Omega,{\mathcal{F}},{\bf P}) is a standard probability space. Let σ(𝐛𝐧𝐝)\sigma({\bf bnd}) be the σ\sigma-algebra generated by the measurable map 𝐛𝐧𝐝:Ω𝒳×𝒳{\bf bnd}:\Omega\to\partial{\mathcal{X}}\times\partial{{\mathcal{X}}^{\star}}. Disintegrating the measure 𝐏{\bf P} with respect to σ(𝐛𝐧𝐝)\sigma({\bf bnd}) yields the system of conditional probability measures {𝐏𝐛𝐧𝐝(𝒘)}𝒘Ω\{{\bf P}^{{\bf bnd}(\bm{w})}\}_{\bm{w}\in\Omega}. More precisely, for every AA\in{\mathcal{F}}, the map Ω\Omega\to{\mathbb{R}}, 𝒘𝐏𝐛𝐧𝐝(𝒘)(A)\bm{w}\mapsto{\bf P}^{{\bf bnd}(\bm{w})}(A) is σ(𝐛𝐧𝐝)\sigma({\bf bnd})-measurable, and

𝐏=Ω𝐏𝐛𝐧𝐝(𝒘)𝑑𝐏(𝒘).{\bf P}=\int_{\Omega}{\bf P}^{{\bf bnd}(\bm{w})}\,d{\bf P}(\bm{w}).

Noting that νπ=𝐛𝐧𝐝𝐏\nu_{\pi}={\bf bnd}_{\ast}{\bf P}, let us write {𝐏ξ,η}(ξ,η)𝒳×𝒳\{{\bf P}^{\xi,\eta}\}_{(\xi,\eta)\in\partial{\mathcal{X}}\times\partial{{\mathcal{X}}^{\star}}} and

𝐏=𝒳×𝒳𝐏ξ,η𝑑νπ(ξ,η).{\bf P}=\int_{\partial{\mathcal{X}}\times\partial{{\mathcal{X}}^{\star}}}{\bf P}^{\xi,\eta}\,d\nu_{\pi}(\xi,\eta).

Similarly, disintegrating 𝐏{\bf P} with respect to the σ\sigma-algebra σ(bnd)\sigma({\rm bnd}^{\star}) generated by bnd{\rm bnd}^{\star} yields the system of conditional probability measures {𝐏η}η𝒳\{{\bf P}^{\eta}\}_{\eta\in\partial{{\mathcal{X}}^{\star}}} satisfying that

𝐏=𝒳𝐏η𝑑νμ(η).{\bf P}=\int_{\partial{{\mathcal{X}}^{\star}}}{\bf P}^{\eta}\,d\nu_{{\mu^{\star}}}(\eta). (2.7)

Let us disintegrate the harmonic measure νπ\nu_{\pi}. In the present setting, 𝒳×𝒳\partial{\mathcal{X}}\times\partial{{\mathcal{X}}^{\star}} is a compact metrizable space and thus 𝒳×𝒳\partial{\mathcal{X}}\times\partial{{\mathcal{X}}^{\star}} endowed with the Borel σ\sigma-algebra is a standard Borel space. The probability measure νπ\nu_{\pi} is disintegrated with respect to the σ\sigma-algebra generated by the projection 𝒳×𝒳𝒳\partial{\mathcal{X}}\times\partial{{\mathcal{X}}^{\star}}\to\partial{{\mathcal{X}}^{\star}}. This yields the system of conditional probability measures {νπη}η𝒳\{\nu_{\pi}^{\eta}\}_{\eta\in\partial{{\mathcal{X}}^{\star}}} such that

νπ=𝒳νπη𝑑νμ(η).\nu_{\pi}=\int_{\partial{{\mathcal{X}}^{\star}}}\nu_{\pi}^{\eta}\,d\nu_{{\mu^{\star}}}(\eta).

Moreover, it satisfies that νπη(𝒳×𝒳)=νπη(𝒳×{η})=1\nu_{\pi}^{\eta}(\partial{\mathcal{X}}\times\partial{{\mathcal{X}}^{\star}})=\nu_{\pi}^{\eta}(\partial{\mathcal{X}}\times\{\eta\})=1 for νμ\nu_{{\mu^{\star}}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}}. These disintegrations lead to by the Fubini theorem,

𝐏η=𝒳×{η}𝐏ξ,η𝑑νπη(ξ)for νμ-almost every η𝒳.{\bf P}^{\eta}=\int_{\partial{\mathcal{X}}\times\{\eta\}}{\bf P}^{\xi,\eta}\,d\nu_{\pi}^{\eta}(\xi)\quad\text{for $\nu_{{\mu^{\star}}}$-almost every $\eta\in\partial{{\mathcal{X}}^{\star}}$}. (2.8)

For νμ\nu_{{\mu^{\star}}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}}, the conditional probability measure 𝐏η{\bf P}^{\eta} coincides with the distribution of a conditional process on 𝚪{\bm{\Gamma}}. This is a Markov chain whose transition probability is defined for νμ\nu_{{\mu^{\star}}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}},

pη(𝒙,𝒚):=π(𝒙1𝒚)d𝒚νμd𝒙νμ(η)if π(𝒙1𝒚)>0, and 0 if otherwise, for 𝒙,𝒚𝚪.p^{\eta}(\bm{x},\bm{y}):=\pi(\bm{x}^{-1}\bm{y})\frac{d\bm{y}\nu_{{\mu^{\star}}}}{d\bm{x}\nu_{{\mu^{\star}}}}(\eta)\quad\text{if $\pi(\bm{x}^{-1}\bm{y})>0$, and $0$ if otherwise, for $\bm{x},\bm{y}\in{\bm{\Gamma}}$}.

Note that 𝒚νμ\bm{y}\nu_{{\mu^{\star}}} is absolutely continuous with respect to 𝒙νμ\bm{x}\nu_{{\mu^{\star}}} and d𝒚νμ/d𝒙νμd\bm{y}\nu_{{\mu^{\star}}}/d\bm{x}\nu_{{\mu^{\star}}} is well-defined νμ\nu_{{\mu^{\star}}}-almost everywhere if π(𝒙1𝒚)>0\pi(\bm{x}^{-1}\bm{y})>0 by (2.5). Moreover since

d𝒚νμd𝒙νμ(η)=d𝒙1𝒚νμdνμ(𝒙1η)for νμ-almost every η𝒳,\frac{d\bm{y}\nu_{{\mu^{\star}}}}{d\bm{x}\nu_{{\mu^{\star}}}}(\eta)=\frac{d\bm{x}^{-1}\bm{y}\nu_{{\mu^{\star}}}}{d\nu_{{\mu^{\star}}}}(\bm{x}^{-1}\eta)\quad\text{for $\nu_{{\mu^{\star}}}$-almost every $\eta\in\partial{{\mathcal{X}}^{\star}}$},

the above pη(𝒙,𝒚)p^{\eta}(\bm{x},\bm{y}) indeed defines a transition probability by the π\pi-stationarity of νμ\nu_{\mu^{\star}}. Let πnη(𝒙):=𝐏η(𝒘n=𝒙)\pi^{\eta}_{n}(\bm{x}):={\bf P}^{\eta}(\bm{w}_{n}=\bm{x}) for 𝒙𝚪\bm{x}\in{\bm{\Gamma}} and n+n\in{\mathbb{Z}}_{+}. It holds that for νμ\nu_{{\mu^{\star}}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}}, for every cylinder set [𝒘0,,𝒘n][\bm{w}_{0},\dots,\bm{w}_{n}] in (Ω,,𝐏)(\Omega,{\mathcal{F}},{\bf P}),

𝐏η([𝒘0,,𝒘n])=𝐏([𝒘0,,𝒘n])d𝒘nνμdνμ(η).{\bf P}^{\eta}([\bm{w}_{0},\dots,\bm{w}_{n}])={\bf P}([\bm{w}_{0},\dots,\bm{w}_{n}])\frac{d\bm{w}_{n}\nu_{\mu^{\star}}}{d\nu_{\mu^{\star}}}(\eta). (2.9)

In particular, for νμ\nu_{\mu^{\star}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}},

πnη(𝒙)=πn(𝒙)d𝒙νμdνμ(η)for 𝒙𝚪 and n+.\pi^{\eta}_{n}(\bm{x})=\pi_{n}(\bm{x})\frac{d\bm{x}\nu_{{\mu^{\star}}}}{d\nu_{{\mu^{\star}}}}(\eta)\quad\text{for $\bm{x}\in{\bm{\Gamma}}$ and $n\in{\mathbb{Z}}_{+}$}.

For more details, see [Kai00, Sections 3 and 4]. There it is shown (in a more general setting) that the Shannon theorem holds for the conditional process. Namely, for νμ\nu_{{\mu^{\star}}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}}, the following limit exists and is constant 𝐏η{\bf P}^{\eta}-almost everywhere:

h(𝐏η):=limn1nlogπnη(𝒘n).h({\bf P}^{\eta}):=\lim_{n\to\infty}-\frac{1}{n}\log\pi^{\eta}_{n}(\bm{w}_{n}). (2.10)

Furthermore, the limit is obtained as

h(𝐏η)=h(π)𝒙𝚪π(𝒙)𝒳logd𝒙νμdνμ(η)𝑑𝒙νμ(η).h({\bf P}^{\eta})=h(\pi)-\sum_{\bm{x}\in{\bm{\Gamma}}}\pi(\bm{x})\int_{\partial{{\mathcal{X}}^{\star}}}\log\frac{d\bm{x}\nu_{{\mu^{\star}}}}{d\nu_{{\mu^{\star}}}}(\eta)\,d\bm{x}\nu_{{\mu^{\star}}}(\eta).

See [Kai00, Theorem 4.5]. The differential entropy for the pair (𝒳,μ)(\partial{{\mathcal{X}}^{\star}},{\mu^{\star}}) is defined by

h(𝒳,μ):=xsuppμμ(x)𝒳logdxνμdνμ(η)𝑑xνμ(η).h(\partial{{\mathcal{X}}^{\star}},{\mu^{\star}}):=\sum_{x\in{\rm supp}\,{\mu^{\star}}}{\mu^{\star}}(x)\int_{\partial{{\mathcal{X}}^{\star}}}\log\frac{dx\nu_{{\mu^{\star}}}}{d\nu_{{\mu^{\star}}}}(\eta)\,dx\nu_{{\mu^{\star}}}(\eta).

Since μ{\mu^{\star}} is a marginal of π\pi, it holds that

h(𝐏η)=h(π)h(𝒳,μ)for νμ-almost every η𝒳.h({\bf P}^{\eta})=h(\pi)-h(\partial{{\mathcal{X}}^{\star}},{\mu^{\star}})\quad\text{for $\nu_{{\mu^{\star}}}$-almost every $\eta\in\partial{{\mathcal{X}}^{\star}}$}.

Let us mention that the differential entropy arises in the theory of Poisson boundary in the following way: It has been proven that h(𝒳,μ)h(μ)h(\partial{{\mathcal{X}}^{\star}},{\mu^{\star}})\leq h({\mu^{\star}}) and the equality holds if and only if (𝒳,νμ)(\partial{{\mathcal{X}}^{\star}},\nu_{{\mu^{\star}}}) is a Poisson boundary for the μ{\mu^{\star}}-random walk [Kai00, Theorem 4.6]. In the present setting, since (𝒳,νμ)(\partial{{\mathcal{X}}^{\star}},\nu_{{\mu^{\star}}}) is π\pi-stationary (2.5), it holds that h(𝒳,μ)=h(𝒳,π)h(\partial{{\mathcal{X}}^{\star}},{\mu^{\star}})=h(\partial{{\mathcal{X}}^{\star}},\pi), and that h(𝒳,π)h(π)h(\partial{{\mathcal{X}}^{\star}},\pi)\leq h(\pi), where the equality holds if and only if (𝒳,νμ)(\partial{{\mathcal{X}}^{\star}},\nu_{\mu^{\star}}) is a Poisson boundary for (𝚪,π)({\bm{\Gamma}},\pi).

3. Exact dimension of conditional measures

For a proper CC-roughly geodesic hyperbolic metric space (𝒳,d)({\mathcal{X}},d) for some C+C\in{\mathbb{R}}_{+} with a fixed base point o𝒳o\in{\mathcal{X}} and for a μ\mu-random walk {wn}n+\{w_{n}\}_{n\in{\mathbb{Z}}_{+}}, the following ray approximation holds for zn=wnoz_{n}=w_{n}\cdot o: If μ\mu is non-elementary and has a finite first moment, then 𝐏{\bf P}-almost surely there exists a CC-rough geodesic ray γz\gamma_{z_{\infty}} such that for l:=l(𝒳,μ)l:=l({\mathcal{X}},\mu) of {zn}n+\{z_{n}\}_{n\in{\mathbb{Z}}_{+}},

d(zn,γz(ln))=o(n).d(z_{n},\gamma_{z_{\infty}}(ln))=o(n). (3.1)

See [Kai00, Theorem 7.3]. In fact, such an assignment ξγξ\xi\mapsto\gamma_{\xi} from 𝒳\partial{\mathcal{X}} to the space of CC-rough geodesic rays from oo in (𝒳,d)({\mathcal{X}},d) equipped with the topology of convergence on compact sets is chosen to be Borel measurable by the Borel selection theorem (cf. [Tan19, Section 3.2]). In the same way, there is a Borel measurable map ηγη\eta\mapsto\gamma_{\eta} from 𝒳\partial{{\mathcal{X}}^{\star}} to the space of CC-rough geodesic rays from oo^{\star} in (𝒳,d)({{\mathcal{X}}^{\star}},d^{\star}). For the drift l:=l(𝒳,μ){l^{\star}}:=l({{\mathcal{X}}^{\star}},{\mu^{\star}}) of {zn}n+\{z^{\star}_{n}\}_{n\in{\mathbb{Z}}_{+}}, it holds that 𝐏{\bf P}-almost surely,

d(zn,γz(ln))=o(n).d^{\star}(z^{\star}_{n},\gamma_{z^{\star}_{\infty}}({l^{\star}}n))=o(n). (3.2)

In the following subsections, for brevity, let

h¯:=h(π)h(𝒳,μ),l:=l(𝒳,μ)andl:=l(𝒳,μ).\overline{h}:=h(\pi)-h(\partial{{\mathcal{X}}^{\star}},{\mu^{\star}}),\quad l:=l({\mathcal{X}},\mu)\quad\text{and}\quad{l^{\star}}:=l({{\mathcal{X}}^{\star}},{\mu^{\star}}).

3.1. Upper bounds on dimensions of conditional measures

Lemma 3.1.

Let 𝚪{\bm{\Gamma}} be a countable subgroup in Isom𝒳×Isom𝒳\operatorname{Isom}{\mathcal{X}}\times\operatorname{Isom}{{\mathcal{X}}^{\star}}, and π\pi be a probability measure on 𝚪{\bm{\Gamma}} with finite first moment and non-elementary marginals μ\mu and μ{\mu^{\star}}. It holds that for νμ\nu_{{\mu^{\star}}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}} and for νπη\nu_{\pi}^{\eta}-almost every 𝛏𝒳×𝒳{\bm{\xi}}\in\partial{\mathcal{X}}\times\partial{{\mathcal{X}}^{\star}},

lim supr0logνπη(𝑩(𝝃,r))logrh(π)h(𝒳,μ)l(𝒳,μ).\limsup_{r\to 0}\frac{\log\nu_{\pi}^{\eta}(\bm{B}({\bm{\xi}},r))}{\log r}\leq\frac{h(\pi)-h(\partial{{\mathcal{X}}^{\star}},{\mu^{\star}})}{l({\mathcal{X}},\mu)}.
Proof.

Recall that for νμ\nu_{{\mu^{\star}}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}}, the distribution 𝐏η{\bf P}^{\eta} is obtained by a Markov chain whose law at time nn is πnη\pi^{\eta}_{n} for n+n\in{\mathbb{Z}}_{+}. For every ε>0\varepsilon>0 and every interval II in +{\mathbb{Z}}_{+}, let

Aε,I:=nI{𝒘Ω:(zn|zn+1)o(lε)n,πnbnd(𝒘)(𝒘n)exp(n(h¯+ε))}.A_{\varepsilon,I}:=\bigcap_{n\in I}\Big{\{}\bm{w}\in\Omega\ :\ (z_{n}|z_{n+1})_{o}\geq(l-\varepsilon)n,\ \pi^{{\rm bnd}^{\star}(\bm{w})}_{n}(\bm{w}_{n})\geq\exp(-n(\overline{h}+\varepsilon))\Big{\}}.

Note that (zn|zn+1)o/nl(z_{n}|z_{n+1})_{o}/n\to l as nn\to\infty almost surely in 𝐏{\bf P} since μ\mu has a finite first moment and the μ\mu-random walk has the drift l>0l>0. By disintegration (2.7), this together with (2.10) implies that for every ε(0,l)\varepsilon\in(0,l) and for νμ\nu_{{\mu^{\star}}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}},

𝐏η(N+Aε,[N,))=1.{\bf P}^{\eta}\Big{(}\bigcup_{N\in{\mathbb{Z}}_{+}}A_{\varepsilon,[N,\infty)}\Big{)}=1.

Hence for every ε>0\varepsilon>0 and for νμ\nu_{{\mu^{\star}}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}}, there exists an Nε,η+N_{\varepsilon,\eta}\in{\mathbb{Z}}_{+} such that

𝐏η(Aε,[Nε,η,))1ε.{\bf P}^{\eta}\big{(}A_{\varepsilon,[N_{\varepsilon,\eta},\infty)}\big{)}\geq 1-\varepsilon.

Let N:=Nε,ηN:=N_{\varepsilon,\eta} and A:=Aε,Nε,ηA:=A_{\varepsilon,N_{\varepsilon,\eta}}. Further for all n>Nn>N, let

A[N,n):=Aε,[N,n)andA[n,):=Aε,[n,).A_{[N,n)}:=A_{\varepsilon,[N,n)}\quad\text{and}\quad A_{[n,\infty)}:=A_{\varepsilon,[n,\infty)}.

Note that A=A[N,n)A[n,)A=A_{[N,n)}\cap A_{[n,\infty)}. For n+n\in{\mathbb{Z}}_{+} and 𝒘Ω\bm{w}\in\Omega, let

Cn(𝒘):={𝒘Ω:𝒘n=𝒘n}.C_{n}(\bm{w}):=\big{\{}\bm{w}^{\prime}\in\Omega\ :\ \bm{w}_{n}^{\prime}=\bm{w}_{n}\big{\}}.

This is the event where the position of the chain is 𝒘n\bm{w}_{n} at time nn. Since the conditional process is a Markov chain, for νμ\nu_{{\mu^{\star}}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}}, for all 𝒘Ω\bm{w}\in\Omega and all n>Nn>N,

𝐏η(ACn(𝒘))=𝐏η(A[N,n)Cn(𝒘))𝐏η(A[n,)Cn(𝒘)).{\bf P}^{\eta}(A\mid C_{n}(\bm{w}))={\bf P}^{\eta}(A_{[N,n)}\mid C_{n}(\bm{w}))\cdot{\bf P}^{\eta}(A_{[n,\infty)}\mid C_{n}(\bm{w})). (3.3)

Furthermore for νμ\nu_{{\mu^{\star}}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}},

𝐏η(A[N,n)Cn(𝒘))=𝐏η(A[N,n)σ(𝒘n,𝒘n+1,))almost everywhere in 𝐏η,{\bf P}^{\eta}(A_{[N,n)}\mid C_{n}(\bm{w}))={\bf P}^{\eta}(A_{[N,n)}\mid\sigma(\bm{w}_{n},\bm{w}_{n+1},\dots))\quad\text{almost everywhere in ${\bf P}^{\eta}$}, (3.4)

where σ(𝒘n,𝒘n+1,)\sigma(\bm{w}_{n},\bm{w}_{n+1},\dots) is the σ\sigma-algebra generated by 𝒘n,𝒘n+1,\bm{w}_{n},\bm{w}_{n+1},\dots. Similarly, one has

𝐏η(A[n,)Cn(𝒘))=𝐏η(A[n,)σ(𝒘0,,𝒘n))almost everywhere in 𝐏η,{\bf P}^{\eta}(A_{[n,\infty)}\mid C_{n}(\bm{w}))={\bf P}^{\eta}(A_{[n,\infty)}\mid\sigma(\bm{w}_{0},\dots,\bm{w}_{n}))\quad\text{almost everywhere in ${\bf P}^{\eta}$}, (3.5)

where σ(𝒘0,,𝒘n)\sigma(\bm{w}_{0},\dots,\bm{w}_{n}) is the σ\sigma-algebra generated by 𝒘0,,𝒘n\bm{w}_{0},\dots,\bm{w}_{n}. Let us denote the tail σ\sigma-algebra by

𝒯:=n+σ(𝒘n,𝒘n+1,).{\mathcal{T}}:=\bigcap_{n\in{\mathbb{Z}}_{+}}\sigma(\bm{w}_{n},\bm{w}_{n+1},\dots).

Note that 𝐏η{\bf P}^{\eta}-almost everywhere on A=A[N,n)A[n,)A=A_{[N,n)}\cap A_{[n,\infty)},

𝐏η(A[N,n)σ(𝒘n,𝒘n+1,))=𝐏η(Aσ(𝒘n,𝒘n+1,)).{\bf P}^{\eta}(A_{[N,n)}\mid\sigma(\bm{w}_{n},\bm{w}_{n+1},\dots))={\bf P}^{\eta}(A\mid\sigma(\bm{w}_{n},\bm{w}_{n+1},\dots)).

By (3.4), the Lévy downward theorem applied to the right hand side above shows that for νμ\nu_{{\mu^{\star}}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}},

limn𝐏η(A[N,n)Cn(𝒘))=𝐏η(A𝒯)almost everywhere in 𝐏η on A.\lim_{n\to\infty}{\bf P}^{\eta}(A_{[N,n)}\mid C_{n}(\bm{w}))={\bf P}^{\eta}(A\mid{\mathcal{T}})\quad\text{almost everywhere in ${\bf P}^{\eta}$ on $A$}. (3.6)

Analogously, note that 𝐏η{\bf P}^{\eta}-almost everywhere on A=A[N,n)A[n,)A=A_{[N,n)}\cap A_{[n,\infty)},

𝐏η(A[n,)σ(𝒘0,,𝒘n))=𝐏η(Aσ(𝒘0,,𝒘n)).{\bf P}^{\eta}(A_{[n,\infty)}\mid\sigma(\bm{w}_{0},\dots,\bm{w}_{n}))={\bf P}^{\eta}(A\mid\sigma(\bm{w}_{0},\dots,\bm{w}_{n})).

By (3.5), the Lévy upward theorem shows that for νμ\nu_{{\mu^{\star}}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}},

limn𝐏η(A[n,)Cn(𝒘))=𝟏Aalmost everywhere in 𝐏η on A.\lim_{n\to\infty}{\bf P}^{\eta}(A_{[n,\infty)}\mid C_{n}(\bm{w}))={\bf 1}_{A}\quad\text{almost everywhere in ${\bf P}^{\eta}$ on $A$}. (3.7)

For νμ\nu_{{\mu^{\star}}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}}, it holds that

𝐏η(A𝒯)>0almost everywhere in 𝐏η on A.{\bf P}^{\eta}(A\mid{\mathcal{T}})>0\quad\text{almost everywhere in ${\bf P}^{\eta}$ on $A$}. (3.8)

Indeed, if we define 𝒩:={𝒘Ω:𝐏η(A𝒯)=0}{\mathcal{N}}:=\{\bm{w}\in\Omega\ :\ {\bf P}^{\eta}(A\mid{\mathcal{T}})=0\}, then 𝒩{\mathcal{N}} is 𝒯{\mathcal{T}}-measurable and 𝐏η(A𝒩𝒯)=0{\bf P}^{\eta}(A\cap{\mathcal{N}}\mid{\mathcal{T}})=0, implying that 𝐏η(A𝒩)=0{\bf P}^{\eta}(A\cap{\mathcal{N}})=0. Thus we have (3.8). Therefore by (3.3), (3.6), (3.7) and (3.8), for νμ\nu_{{\mu^{\star}}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}},

limn𝐏η(ACn(𝒘))=𝐏η(A𝒯)>0almost everywhere in 𝐏η on A.\lim_{n\to\infty}{\bf P}^{\eta}(A\mid C_{n}(\bm{w}))={\bf P}^{\eta}(A\mid{\mathcal{T}})>0\quad\text{almost everywhere in ${\bf P}^{\eta}$ on $A$}. (3.9)

For every 𝒘A\bm{w}\in A, it holds that

q(zn,z)=exp((zn|z)o)Ce(lε)nfor all n>N,q(z_{n},z_{\infty})=\exp(-(z_{n}|z_{\infty})_{o})\leq Ce^{-(l-\varepsilon)n}\quad\text{for all $n>N$},

where CC is independent of 𝒘\bm{w}. This follows since qq with a power is bi-Lipschitz to a genuine metric (cf. Section 2.1). Further for n>Nn>N and for 𝐏η{\bf P}^{\eta}-almost every 𝒘ACn(𝒘)\bm{w}^{\prime}\in A\cap C_{n}(\bm{w}),

𝒘n=𝒘nandq(zn,z)Ce(lε)n,\bm{w}_{n}^{\prime}=\bm{w}_{n}\quad\text{and}\quad q(z_{n}^{\prime},z_{\infty}^{\prime})\leq Ce^{-(l-\varepsilon)n},

where zn:=wnoz^{\prime}_{n}:=w_{n}^{\prime}\cdot o for 𝒘n=(wn,wn)\bm{w}_{n}=(w_{n}^{\prime},{w^{\star}_{n}}^{{}^{\prime}}) and znzz_{n}^{\prime}\to z_{\infty}^{\prime} as nn\to\infty in 𝒳𝒳{\mathcal{X}}\cup\partial{\mathcal{X}}. Since 𝒘n=𝒘n\bm{w}_{n}^{\prime}=\bm{w}_{n} and thus zn=znz_{n}^{\prime}=z_{n}, by (2.1) extended on 𝒳𝒳{\mathcal{X}}\cup\partial{\mathcal{X}},

q(z,z)Ceδ(lε)n,i.e.,zB(z,Ceδ(lε)n)for all n>N.q(z_{\infty},z_{\infty}^{\prime})\leq Ce^{\delta-(l-\varepsilon)n},\quad\text{i.e.,}\quad z_{\infty}^{\prime}\in B(z_{\infty},Ce^{\delta-(l-\varepsilon)n})\quad\text{for all $n>N$}.

Hence for νμ\nu_{{\mu^{\star}}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}},

𝐏η(ACn(𝒘))νπη(B(z,Ceδ(lε)n)×𝒳)for all n>N.{\bf P}^{\eta}(A\cap C_{n}(\bm{w}))\leq\nu_{\pi}^{\eta}(B(z_{\infty},Ce^{\delta-(l-\varepsilon)n})\times\partial{{\mathcal{X}}^{\star}})\quad\text{for all $n>N$}.

The right hand side coincides with νπη(𝑩(𝒛,Ceδ(lε)n))\nu_{\pi}^{\eta}(\bm{B}(\bm{z}_{\infty},Ce^{\delta-(l-\varepsilon)n})) where 𝒛=(z,z)\bm{z}_{\infty}=(z_{\infty},z^{\star}_{\infty}) since νπη\nu_{\pi}^{\eta} is supported in 𝒳×{η}\partial{\mathcal{X}}\times\{\eta\} and η=z\eta=z^{\star}_{\infty} for νμ\nu_{{\mu^{\star}}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}}. For 𝐏η{\bf P}^{\eta}-almost every 𝒘A\bm{w}\in A, it holds that for all n>Nn>N,

𝐏η(ACn(𝒘))=𝐏η(ACn(𝒘))𝐏η(Cn(𝒘))𝐏η(ACn(𝒘))en(h¯+ε).{\bf P}^{\eta}(A\cap C_{n}(\bm{w}))={\bf P}^{\eta}(A\mid C_{n}(\bm{w}))\cdot{\bf P}^{\eta}(C_{n}(\bm{w}))\geq{\bf P}^{\eta}(A\mid C_{n}(\bm{w}))\cdot e^{-n(\overline{h}+\varepsilon)}.

This shows that for νμ\nu_{{\mu^{\star}}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}}, for 𝐏η{\bf P}^{\eta}-almost every 𝒘A\bm{w}\in A and for all n>Nn>N,

νπη(𝑩(𝒛,Ceδ(lε)n))𝐏η(ACn(𝒘))en(h¯+ε).\nu_{\pi}^{\eta}(\bm{B}(\bm{z}_{\infty},Ce^{\delta-(l-\varepsilon)n}))\geq{\bf P}^{\eta}(A\mid C_{n}(\bm{w}))\cdot e^{-n(\overline{h}+\varepsilon)}.

By (3.9), for νμ\nu_{{\mu^{\star}}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}} and for 𝐏η{\bf P}^{\eta}-almost every 𝒘A\bm{w}\in A,

lim supr0logνπη(𝑩(𝒛,r))logrh¯+εlε.\limsup_{r\to 0}\frac{\log\nu_{\pi}^{\eta}(\bm{B}(\bm{z}_{\infty},r))}{\log r}\leq\frac{\overline{h}+\varepsilon}{l-\varepsilon}. (3.10)

This follows first for the sequence rn0r_{n}\to 0 as nn\to\infty where rn:=Ceδ(lε)nr_{n}:=Ce^{\delta-(l-\varepsilon)n} and then for r>0r>0 and r0r\to 0 by noting that rn+1=e(lε)rnr_{n+1}=e^{-(l-\varepsilon)}r_{n}. Recall that 𝐏η(A)1ε{\bf P}^{\eta}(A)\geq 1-\varepsilon for the event A=Aε,[Nε,η,)A=A_{\varepsilon,[N_{\varepsilon,\eta},\infty)} for νμ\nu_{{\mu^{\star}}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}}. For an arbitrary decreasing sequence εn0\varepsilon_{n}\to 0 as nn\to\infty, one has 𝐏η(m+nmAεn,[Nεn,η,))=1{\bf P}^{\eta}(\bigcap_{m\in{\mathbb{Z}}_{+}}\bigcup_{n\geq m}A_{\varepsilon_{n},[N_{\varepsilon_{n},\eta},\infty)})=1 for νμ\nu_{{\mu^{\star}}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}}. Therefore by (3.10) for νμ\nu_{{\mu^{\star}}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}},

lim supr0logνπη(𝑩(𝒛,r))logrh¯lalmost everywhere in 𝐏η.\limsup_{r\to 0}\frac{\log\nu_{\pi}^{\eta}(\bm{B}(\bm{z}_{\infty},r))}{\log r}\leq\frac{\overline{h}}{l}\quad\text{almost everywhere in ${\bf P}^{\eta}$}.

Noting that for νμ\nu_{{\mu^{\star}}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}}, the distribution of 𝒛\bm{z}_{\infty} is νπη\nu_{\pi}^{\eta}, we obtain for νμ\nu_{{\mu^{\star}}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}},

lim supr0logνπη(𝑩(𝝃,r))logrh¯lfor νπη-almost every 𝝃𝒳×𝒳.\limsup_{r\to 0}\frac{\log\nu_{\pi}^{\eta}(\bm{B}({\bm{\xi}},r))}{\log r}\leq\frac{\overline{h}}{l}\quad\text{for $\nu_{\pi}^{\eta}$-almost every ${\bm{\xi}}\in\partial{\mathcal{X}}\times\partial{{\mathcal{X}}^{\star}}$}.

This concludes the claim. ∎

We use the following version of Lemma 3.1 in Section 4.

Lemma 3.2.

In the same setting as in Lemma 3.1, if for νμ\nu_{\mu^{\star}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}} there exists a Borel set FηF_{\eta} in 𝒳×𝒳\partial{\mathcal{X}}\times\partial{{\mathcal{X}}^{\star}} such that νπη(Fη)>0\nu_{\pi}^{\eta}(F_{\eta})>0, then for νμ\nu_{\mu^{\star}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}} and for νπη\nu_{\pi}^{\eta}-almost every 𝛏Fη{\bm{\xi}}\in F_{\eta},

lim supr0logνπη(𝑩(𝝃,r)Fη)logrh(π)h(𝒳,μ)l(𝒳,μ).\limsup_{r\to 0}\frac{\log\nu_{\pi}^{\eta}(\bm{B}({\bm{\xi}},r)\cap F_{\eta})}{\log r}\leq\frac{h(\pi)-h(\partial{{\mathcal{X}}^{\star}},\mu^{\star})}{l({\mathcal{X}},\mu)}.
Proof.

This follows from Lemmas 2.2 and 3.1. ∎

3.2. Lower bounds on dimensions of conditional measures

Lemma 3.3.

Let 𝚪{\bm{\Gamma}} be a countable subgroup in Isom𝒳×Isom𝒳\operatorname{Isom}{\mathcal{X}}\times\operatorname{Isom}{{\mathcal{X}}^{\star}} with finite exponential growth relative to (𝒳×𝒳,𝐝)({\mathcal{X}}\times{{\mathcal{X}}^{\star}},\bm{d}), and π\pi be a probability measure on 𝚪{\bm{\Gamma}} with finite first moment and non-elementary marginals μ\mu and μ{\mu^{\star}}. For every ε>0\varepsilon>0, there exist

  1. (1)

    an N+N\in{\mathbb{Z}}_{+},

  2. (2)

    a Borel set DD in 𝒳\partial{{\mathcal{X}}^{\star}} with νμ(D)1ε\nu_{\mu^{\star}}(D)\geq 1-\varepsilon, and

  3. (3)

    a Borel set FF in 𝒳×𝒳\partial{\mathcal{X}}\times\partial{{\mathcal{X}}^{\star}} with νπη(F)1ε\nu_{\pi}^{\eta}(F)\geq 1-\varepsilon for νμ\nu_{\mu^{\star}}-almost every ηD\eta\in D and νπ(F)1ε\nu_{\pi}(F)\geq 1-\varepsilon,

such that the following holds: For νμ\nu_{\mu^{\star}}-almost every ηD\eta\in D, for all ξ𝒳\xi\in\partial{\mathcal{X}} and all nNn\geq N,

νπη(B(ξ,eln)×𝒳F)Cεen(h¯ε),\nu_{\pi}^{\eta}\left(B(\xi,e^{-ln})\times\partial{{\mathcal{X}}^{\star}}\cap F\right)\leq C_{\varepsilon}e^{-n(\overline{h}-\varepsilon)},

where CεC_{\varepsilon} is a constant depending only on ε\varepsilon.

Proof.

For every ε>0\varepsilon>0 and every N+N\in{\mathbb{Z}}_{+}, let

Aε,N:=nN{𝒘Ω:𝒅(𝒛n,(γz(ln),γz(ln)))εn,πnbnd(𝒘)(𝒘n)exp(n(h¯ε))}.A_{\varepsilon,N}:=\bigcap_{n\geq N}\Big{\{}\bm{w}\in\Omega\ :\ \bm{d}(\bm{z}_{n},(\gamma_{z_{\infty}}(ln),\gamma_{z^{\star}_{\infty}}({l^{\star}}n)))\leq\varepsilon n,\ \pi^{{\rm bnd}^{\star}(\bm{w})}_{n}(\bm{w}_{n})\leq\exp(-n(\overline{h}-\varepsilon))\Big{\}}.

The disintegration formula (2.7) implies that for all event AΩA\subset\Omega if 𝐏(A)=1{\bf P}(A)=1, then 𝐏η(A)=1{\bf P}^{\eta}(A)=1 for νμ\nu_{{\mu^{\star}}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}}. This together with (3.1) and (3.2), and further (2.10) imply that for every ε>0\varepsilon>0,

𝐏η(N+Aε,N)=1for νμ-almost every η𝒳.{\bf P}^{\eta}\Big{(}\bigcup_{N\in{\mathbb{Z}}_{+}}A_{\varepsilon,N}\Big{)}=1\quad\text{for $\nu_{{\mu^{\star}}}$-almost every $\eta\in\partial{{\mathcal{X}}^{\star}}$}.

For N+N\in{\mathbb{Z}}_{+}, let

Dε,N:={η𝒳:𝐏η(Aε,N)1ε}.D_{\varepsilon,N}:=\Big{\{}\eta\in\partial{{\mathcal{X}}^{\star}}\ :\ {\bf P}^{\eta}(A_{\varepsilon,N})\geq 1-\varepsilon\Big{\}}.

Since Aε,NA_{\varepsilon,N} is increasing and 𝐏η(Aε,N)1{\bf P}^{\eta}(A_{\varepsilon,N})\to 1 monotonically as NN\to\infty for νμ\nu_{\mu^{\star}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}}, there exists an Nε+N_{\varepsilon}\in{\mathbb{Z}}_{+} such that

νμ(Dε,Nε)1ε.\nu_{\mu^{\star}}\left(D_{\varepsilon,N_{\varepsilon}}\right)\geq 1-\varepsilon. (3.11)

Let N:=NεN:=N_{\varepsilon}, D:=Dε,NεD:=D_{\varepsilon,N_{\varepsilon}} and A:=Aε,NεA:=A_{\varepsilon,N_{\varepsilon}}. It holds that

𝐏η(A)1εfor νμ-almost every ηD.{\bf P}^{\eta}(A)\geq 1-\varepsilon\quad\text{for $\nu_{\mu^{\star}}$-almost every $\eta\in D$}. (3.12)

Let us define

Fε:={(ξ,η)𝒳×𝒳:𝐏ξ,η(A)ε},F_{\varepsilon}:=\Big{\{}(\xi,\eta)\in\partial{\mathcal{X}}\times\partial{{\mathcal{X}}^{\star}}\ :\ {\bf P}^{\xi,\eta}(A)\geq\varepsilon\Big{\}},

and F:=FεF:=F_{\varepsilon}. We claim that

νπη(F)12εfor νμ-almost every ηD.\nu_{\pi}^{\eta}(F)\geq 1-2\varepsilon\quad\text{for $\nu_{{\mu^{\star}}}$-almost every $\eta\in D$}. (3.13)

Indeed, by (3.12) and by (2.8), for νμ\nu_{{\mu^{\star}}}-almost every ηD\eta\in D,

1ε𝐏η(A)=𝒳×{η}𝐏ξ,η(A)𝑑νπη(ξ)=F𝐏ξ,η(A)𝑑νπη(ξ)+𝒳×𝒳F𝐏ξ,η(A)𝑑νπη(ξ).1-\varepsilon\leq{\bf P}^{\eta}(A)=\int_{\partial{\mathcal{X}}\times\{\eta\}}{\bf P}^{\xi,\eta}(A)\,d\nu_{\pi}^{\eta}(\xi)=\int_{F}{\bf P}^{\xi,\eta}(A)\,d\nu_{\pi}^{\eta}(\xi)+\int_{\partial{\mathcal{X}}\times\partial{{\mathcal{X}}^{\star}}\setminus F}{\bf P}^{\xi,\eta}(A)\,d\nu_{\pi}^{\eta}(\xi).

This implies that

1ενπη(F)+ενπη(𝒳×𝒳F)νπη(F)+ε,1-\varepsilon\leq\nu_{\pi}^{\eta}(F)+\varepsilon\cdot\nu_{\pi}^{\eta}(\partial{\mathcal{X}}\times\partial{{\mathcal{X}}^{\star}}\setminus F)\leq\nu_{\pi}^{\eta}(F)+\varepsilon,

showing (3.13).

Furthermore it holds that

νπ(F)13ε.\nu_{\pi}(F)\geq 1-3\varepsilon. (3.14)

This follows since νπη(F)(12ε)𝟏D\nu_{\pi}^{\eta}(F)\geq(1-2\varepsilon){\bf 1}_{D} by (3.13), integration with respect to νμ\nu_{\mu^{\star}} yields

νπ(F)=𝒳νπη(F)𝑑νμ(η)(12ε)νμ(D)(12ε)(1ε)13ε,\nu_{\pi}(F)=\int_{\partial{{\mathcal{X}}^{\star}}}\nu_{\pi}^{\eta}(F)\,d\nu_{\mu^{\star}}(\eta)\geq(1-2\varepsilon)\nu_{\mu^{\star}}(D)\geq(1-2\varepsilon)(1-\varepsilon)\geq 1-3\varepsilon,

where the second inequality uses (3.11). It holds that for νμ\nu_{{\mu^{\star}}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}}, for every ξ𝒳\xi\in\partial{\mathcal{X}} and R>0R>0,

νπη(𝒪(γξ(ln),R)×𝒳F)=𝐏η({𝒛𝒪(γξ(ln),R)×𝒳F})\displaystyle\nu_{\pi}^{\eta}({\mathcal{O}}(\gamma_{\xi}(ln),R)\times\partial{{\mathcal{X}}^{\star}}\cap F)={\bf P}^{\eta}(\{\bm{z}_{\infty}\in{\mathcal{O}}(\gamma_{\xi}(ln),R)\times\partial{{\mathcal{X}}^{\star}}\cap F\})
=𝐏η({𝒛𝒪(γξ(ln),R)×𝒳F}A)\displaystyle\qquad\qquad={\bf P}^{\eta}(\{\bm{z}_{\infty}\in{\mathcal{O}}(\gamma_{\xi}(ln),R)\times\partial{{\mathcal{X}}^{\star}}\cap F\}\cap A)
+𝐏η({𝒛𝒪(γξ(ln),R)×𝒳F}A𝖼).\displaystyle\qquad\qquad\qquad\qquad\qquad+{\bf P}^{\eta}(\{\bm{z}_{\infty}\in{\mathcal{O}}(\gamma_{\xi}(ln),R)\times\partial{{\mathcal{X}}^{\star}}\cap F\}\cap A^{\sf c}). (3.15)

In the above, A𝖼A^{\sf c} denote the complement event of AA, and 𝒛=(z,z)\bm{z}_{\infty}=(z_{\infty},z^{\star}_{\infty}). First let us bound the first term in (3.2). On the event AA, if z𝒪(γξ(ln),R)z_{\infty}\in{\mathcal{O}}(\gamma_{\xi}(ln),R) and nNn\geq N, then

d(zn,γξ(ln))d(zn,γz(ln))+d(γz(ln),γξ(ln))εn+CR,d(z_{n},\gamma_{\xi}(ln))\leq d(z_{n},\gamma_{z_{\infty}}(ln))+d(\gamma_{z_{\infty}}(ln),\gamma_{\xi}(ln))\leq\varepsilon n+C_{R},

where CR:=2R+2CC_{R}:=2R+2C since γξ\gamma_{\xi} is a CC-rough geodesic ray. Moreover, since νμ\nu_{{\mu^{\star}}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}} it holds that z=ηz^{\star}_{\infty}=\eta almost surely in 𝐏η{\bf P}^{\eta}, it holds that 𝐏η{\bf P}^{\eta}-almost everywhere on AA for all nNn\geq N,

d(zn,γη(ln))εn.d^{\star}(z^{\star}_{n},\gamma_{\eta}({l^{\star}}n))\leq\varepsilon n.

Hence 𝐏η{\bf P}^{\eta}-almost everywhere on AA for all nNn\geq N,

𝒅(𝒛n,γξ,η,n)εn+CRwhere γξ,η,n:=(γξ(ln),γη(ln)).\bm{d}(\bm{z}_{n},\gamma_{\xi,\eta,n})\leq\varepsilon n+C_{R}\quad\text{where $\gamma_{\xi,\eta,n}:=(\gamma_{\xi}(ln),\gamma_{\eta}({l^{\star}}n))$}.

This shows that letting 𝑩(𝒙,r)\bm{B}(\bm{x},r) denote the ball in (𝒳×𝒳,𝒅)({\mathcal{X}}\times{{\mathcal{X}}^{\star}},\bm{d}), we have for all nNn\geq N,

𝐏η({𝒛𝒪(γξ(ln),R)×𝒳F}A)\displaystyle{\bf P}^{\eta}(\{\bm{z}_{\infty}\in{\mathcal{O}}(\gamma_{\xi}(ln),R)\times\partial{{\mathcal{X}}^{\star}}\cap F\}\cap A)
𝐏η({𝒛n𝑩(γξ,η,n,εn+CR)}{πnη(𝒘n)exp(n(h¯ε))}).\displaystyle\qquad\qquad\qquad\leq{\bf P}^{\eta}(\{\bm{z}_{n}\in\bm{B}(\gamma_{\xi,\eta,n},\varepsilon n+C_{R})\}\cap\{\pi^{\eta}_{n}(\bm{w}_{n})\leq\exp(-n(\overline{h}-\varepsilon))\}).

The right hand side is at most πnη(𝒙)\sum\pi^{\eta}_{n}(\bm{x}) where the summation runs over all 𝒙𝚪\bm{x}\in{\bm{\Gamma}} such that 𝒙𝒐𝑩(γξ,η,n,εn+CR)\bm{x}\cdot\bm{o}\in\bm{B}(\gamma_{\xi,\eta,n},\varepsilon n+C_{R}) and πnη(𝒙)exp(n(h¯ε))\pi^{\eta}_{n}(\bm{x})\leq\exp(-n(\overline{h}-\varepsilon)). This is at most

#{𝒙𝚪:𝒙𝒐𝑩(γξ,η,n,εn+CR)}en(h¯ε)cec(εn+CR)en(h¯ε),\#\big{\{}\bm{x}\in{\bm{\Gamma}}\ :\ \bm{x}\cdot\bm{o}\in\bm{B}(\gamma_{\xi,\eta,n},\varepsilon n+C_{R})\big{\}}\cdot e^{-n(\overline{h}-\varepsilon)}\leq ce^{c(\varepsilon n+C_{R})}\cdot e^{-n(\overline{h}-\varepsilon)},

for a constant c>0c>0 since 𝚪{\bm{\Gamma}} has a finite exponential growth relative to (𝒳×𝒳,𝒅)({\mathcal{X}}\times{{\mathcal{X}}^{\star}},\bm{d}). Thus for νμ\nu_{{\mu^{\star}}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}} and for every ξ𝒳\xi\in\partial{\mathcal{X}}, for all nNn\geq N,

𝐏η({𝒛𝒪(γξ(ln),R)×𝒳F}A)cec(εn+CR)en(h¯ε).{\bf P}^{\eta}(\{\bm{z}_{\infty}\in{\mathcal{O}}(\gamma_{\xi}(ln),R)\times\partial{{\mathcal{X}}^{\star}}\cap F\}\cap A)\leq ce^{c(\varepsilon n+C_{R})}\cdot e^{-n(\overline{h}-\varepsilon)}. (3.16)

Next let us bound the second term in (3.2). By (2.8), it holds that for νμ\nu_{\mu^{\star}}-almost every ηD\eta\in D,

𝐏η({𝒛𝒪(γξ(ln),R)×𝒳F}A𝖼)\displaystyle{\bf P}^{\eta}(\{\bm{z}_{\infty}\in{\mathcal{O}}(\gamma_{\xi}(ln),R)\times\partial{{\mathcal{X}}^{\star}}\cap F\}\cap A^{\sf c})
=𝒪(γξ(ln),R)×𝒳F𝐏ζ,η(A𝖼)𝑑νπη(ζ)(1ε)νπη(𝒪(γξ(ln),R)×𝒳F).\displaystyle=\int_{{\mathcal{O}}(\gamma_{\xi}(ln),R)\times\partial{{\mathcal{X}}^{\star}}\cap F}{\bf P}^{\zeta,\eta}(A^{\sf c})\,d\nu_{\pi}^{\eta}(\zeta)\leq(1-\varepsilon)\cdot\nu_{\pi}^{\eta}({\mathcal{O}}(\gamma_{\xi}(ln),R)\times\partial{{\mathcal{X}}^{\star}}\cap F). (3.17)

In the above, the inequality holds since 𝐏ζ,η(A𝖼)1ε{\bf P}^{\zeta,\eta}(A^{\sf c})\leq 1-\varepsilon for νπη\nu_{\pi}^{\eta}-almost every (ζ,η)F(\zeta,\eta)\in F for νμ\nu_{\mu^{\star}}-almost every ηD\eta\in D by the definition of FF.

Finally, combining (3.2), (3.16) and (3.2) yields for νμ\nu_{\mu^{\star}}-almost every ηD\eta\in D, for every ξ𝒳\xi\in\partial{\mathcal{X}} and for all nNn\geq N,

νπη(𝒪(γξ(ln),R)×𝒳F)cec(εn+CR)en(h¯ε)+(1ε)νπη(𝒪(γξ(ln),R)×𝒳F).\nu_{\pi}^{\eta}({\mathcal{O}}(\gamma_{\xi}(ln),R)\times\partial{{\mathcal{X}}^{\star}}\cap F)\leq ce^{c(\varepsilon n+C_{R})}\cdot e^{-n(\overline{h}-\varepsilon)}+(1-\varepsilon)\cdot\nu_{\pi}^{\eta}({\mathcal{O}}(\gamma_{\xi}(ln),R)\times\partial{{\mathcal{X}}^{\star}}\cap F).

Therefore for νμ\nu_{{\mu^{\star}}}-almost every ηD\eta\in D and for every ξ𝒳\xi\in\partial{\mathcal{X}}, for all nNn\geq N,

ενπη(𝒪(γξ(ln),R)×𝒳F)cec(εn+CR)en(h¯ε).\varepsilon\cdot\nu_{\pi}^{\eta}\left({\mathcal{O}}(\gamma_{\xi}(ln),R)\times\partial{{\mathcal{X}}^{\star}}\cap F\right)\leq ce^{c(\varepsilon n+C_{R})}\cdot e^{-n(\overline{h}-\varepsilon)}. (3.18)

Note that B(ξ,C1eln+R)𝒳𝒪(γξ(ln),R)B(\xi,C^{-1}e^{-ln+R})\cap\partial{\mathcal{X}}\subset{\mathcal{O}}(\gamma_{\xi}(ln),R) by (2.3), where we choose a large enough constant RR so that eR/C1e^{R}/C\geq 1 and CC depends only on the hyperbolicity constant. In (3.18), the constant cc depends only on 𝚪{\bm{\Gamma}} and (𝒳×𝒳,𝒅)({\mathcal{X}}\times{{\mathcal{X}}^{\star}},\bm{d}). For every ε>0\varepsilon>0, we argue with ε=ε/(3+c)\varepsilon^{\prime}=\varepsilon/(3+c). By (3.11), (3.13) and (3.14), we obtain

νμ(D)1ε1ε,νπη(F)12ε1εfor νμ-almost every ηD,\nu_{\mu^{\star}}(D)\geq 1-\varepsilon^{\prime}\geq 1-\varepsilon,\quad\nu_{\pi}^{\eta}(F)\geq 1-2\varepsilon^{\prime}\geq 1-\varepsilon\quad\text{for $\nu_{\mu^{\star}}$-almost every $\eta\in D$},

and νπ(F)13ε1ε\nu_{\pi}(F)\geq 1-3\varepsilon^{\prime}\geq 1-\varepsilon. Further by (3.18), for νμ\nu_{\mu^{\star}}-almost every ηD\eta\in D, for every ξ𝒳\xi\in\partial{\mathcal{X}} and for all nNn\geq N,

νπη(B(ξ,eln)F)(c/ε)ecCRen(h¯ε).\nu_{\pi}^{\eta}\left(B(\xi,e^{-ln})\cap F\right)\leq(c/\varepsilon^{\prime})e^{cC_{R}}\cdot e^{-n(\overline{h}-\varepsilon)}.

Defining the constant Cε:=(c/ε)ecCRC_{\varepsilon}:=(c/\varepsilon^{\prime})e^{cC_{R}} yields the claim. ∎

Lemma 3.4.

In the same setting as in Lemma 3.3, it holds that for νμ\nu_{{\mu^{\star}}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}} and for νπη\nu_{\pi}^{\eta}-almost every 𝛏𝒳×𝒳{\bm{\xi}}\in\partial{\mathcal{X}}\times\partial{{\mathcal{X}}^{\star}},

lim infr0logνπη(𝑩(𝝃,r))logrh(π)h(𝒳,μ)l(𝒳,μ).\liminf_{r\to 0}\frac{\log\nu_{\pi}^{\eta}(\bm{B}({\bm{\xi}},r))}{\log r}\geq\frac{h(\pi)-h(\partial{{\mathcal{X}}^{\star}},{\mu^{\star}})}{l({\mathcal{X}},\mu)}.
Proof.

By Lemma 3.3, for every ε>0\varepsilon>0 there exist a Borel set DD in 𝒳\partial{{\mathcal{X}}^{\star}} with νμ(D)1ε\nu_{\mu^{\star}}(D)\geq 1-\varepsilon and a Borel set FF in 𝒳×𝒳\partial{\mathcal{X}}\times\partial{{\mathcal{X}}^{\star}} with νπ(F)1ε\nu_{\pi}(F)\geq 1-\varepsilon such that the following holds: For νμ\nu_{{\mu^{\star}}}-almost every ηD\eta\in D and for every ξ𝒳\xi\in\partial{\mathcal{X}},

lim infnlogνπη(𝑩((ξ,η),eln)F)lnh¯εl.\liminf_{n\to\infty}\frac{\log\nu_{\pi}^{\eta}(\bm{B}((\xi,\eta),e^{-ln})\cap F)}{-ln}\geq\frac{\overline{h}-\varepsilon}{l}.

In fact, in the above the sequence rn:=elnr_{n}:=e^{-ln} for n+n\in{\mathbb{Z}}_{+} is replaced by positive reals rr tending to 0 since rn+1=elrnr_{n+1}=e^{-l}r_{n} for all n+n\in{\mathbb{Z}}_{+}. Applying Lemma 2.2 to the measures νπη\nu_{\pi}^{\eta} and FF implies the following: There exists a constant L1L\geq 1 such that for νμ\nu_{{\mu^{\star}}}-almost every ηD\eta\in D, for νπη\nu_{\pi}^{\eta}-almost every (ξ,η)F(\xi,\eta)\in F and for a constant r(ξ,η)>0r(\xi,\eta)>0,

νπη(𝑩((ξ,η),Lr)F)910νπη(𝑩((ξ,η),r))for all r(0,r(ξ,η)).\nu_{\pi}^{\eta}(\bm{B}((\xi,\eta),Lr)\cap F)\geq\frac{9}{10}\nu_{\pi}^{\eta}(\bm{B}((\xi,\eta),r))\quad\text{for all $r\in(0,r(\xi,\eta))$}.

Hence for νμ\nu_{{\mu^{\star}}}-almost every ηD\eta\in D and for νπη\nu_{\pi}^{\eta}-almost every 𝝃=(ξ,η)F{\bm{\xi}}=(\xi,\eta)\in F,

lim infr0logνπη(𝑩(𝝃,r))logrh¯εl.\liminf_{r\to 0}\frac{\log\nu_{\pi}^{\eta}(\bm{B}({\bm{\xi}},r))}{\log r}\geq\frac{\overline{h}-\varepsilon}{l}.

Since for every ε>0\varepsilon>0 there exists such an FF denoted by FεF_{\varepsilon} with νπη(Fε)1ε\nu_{\pi}^{\eta}(F_{\varepsilon})\geq 1-\varepsilon, for νμ\nu_{{\mu^{\star}}}-almost every ηD\eta\in D, it holds that νπη(m+nmFεn)=1\nu_{\pi}^{\eta}(\bigcap_{m\in{\mathbb{Z}}_{+}}\bigcup_{n\geq m}F_{\varepsilon_{n}})=1 for an arbitrary decreasing sequence εn0\varepsilon_{n}\to 0 as nn\to\infty. Therefore it follows that for νμ\nu_{{\mu^{\star}}}-almost every ηD\eta\in D and for νπη\nu_{\pi}^{\eta}-almost every 𝝃𝒳×𝒳{\bm{\xi}}\in\partial{\mathcal{X}}\times\partial{{\mathcal{X}}^{\star}},

lim infr0logνπη(𝑩(𝝃,r))logrh¯l.\liminf_{r\to 0}\frac{\log\nu_{\pi}^{\eta}(\bm{B}({\bm{\xi}},r))}{\log r}\geq\frac{\overline{h}}{l}.

For every ε>0\varepsilon>0 there exists such a DD with νμ(D)1ε\nu_{\mu^{\star}}(D)\geq 1-\varepsilon, the above holds for νμ\nu_{\mu^{\star}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}}. This concludes the claim. ∎

3.3. Proofs of Theorems 1.2 and 1.3

Proof of Theorem 1.3.

Since μ\mu and μ{\mu^{\star}} are non-elementary and have finite first moments, the drift l(𝒳,μ)l({\mathcal{X}},\mu) is finite and positive, further the asymptotic entropy h(π)h(\pi) and the differential entropy h(𝒳,μ)h(\partial{{\mathcal{X}}^{\star}},{\mu^{\star}}) are finite. The first claim follows from Lemmas 3.1 and 3.4. The second claim follows from Lemma 2.1. ∎

Proof of Theorem 1.2.

Since a probability measure μ{\mu^{\star}} on Γ{\Gamma^{\star}} is non-elementary with finite first moment, it holds that h(μ)=h(Γ,μ)h({\mu^{\star}})=h(\partial{\Gamma^{\star}},{\mu^{\star}}) since (Γ,νμ)(\partial{\Gamma^{\star}},\nu_{\mu^{\star}}) is a Poisson boundary for (Γ,μ)({\Gamma^{\star}},{\mu^{\star}}) [Kai00, Theorem 7.4]. The claim follows from Theorem 1.3 by applying Γ\Gamma and Γ{\Gamma^{\star}} endowed with left invariant hyperbolic metrics quasi-isometric to word metrics respectively to (𝒳,d)({\mathcal{X}},d) and (𝒳,d)({{\mathcal{X}}^{\star}},d^{\star}). ∎

4. Exact dimension of harmonic measures in products spaces

As in Section 3, for brevity, let

l:=l(𝒳,μ),l:=l(𝒳,μ),h:=h(𝒳,μ)andh¯:=h(π)h.l:=l({\mathcal{X}},\mu),\quad{l^{\star}}:=l({{\mathcal{X}}^{\star}},{\mu^{\star}}),\quad h^{\star}:=h(\partial{{\mathcal{X}}^{\star}},{\mu^{\star}})\quad\text{and}\quad\overline{h}:=h(\pi)-h^{\star}.

4.1. Upper bounds on dimensions of harmonic measures in product spaces

The proof of the following proposition is inspired by [LL23, Section 8].

Proposition 4.1.

Let 𝚪{\bm{\Gamma}} be a countable subgroup in Isom𝒳×Isom𝒳\operatorname{Isom}{\mathcal{X}}\times\operatorname{Isom}{{\mathcal{X}}^{\star}}, and π\pi be a probability measure on 𝚪{\bm{\Gamma}} with finite first moment and non-elementary marginals μ\mu and μ{\mu^{\star}}. If l(𝒳,μ)l(𝒳,μ)l({\mathcal{X}},\mu)\geq l({{\mathcal{X}}^{\star}},{\mu^{\star}}), then it holds that for νπ\nu_{\pi}-almost every 𝛏𝒳×𝒳{\bm{\xi}}\in\partial{\mathcal{X}}\times\partial{{\mathcal{X}}^{\star}},

lim supr0logνπ(𝑩(𝝃,r))logrh(π)h(𝒳,μ)l(𝒳,μ)+h(𝒳,μ)l(𝒳,μ).\limsup_{r\to 0}\frac{\log\nu_{\pi}\left(\bm{B}({\bm{\xi}},r)\right)}{\log r}\leq\frac{h(\pi)-h(\partial{{\mathcal{X}}^{\star}},{\mu^{\star}})}{l({\mathcal{X}},\mu)}+\frac{h(\partial{{\mathcal{X}}^{\star}},{\mu^{\star}})}{l({{\mathcal{X}}^{\star}},{\mu^{\star}})}.

We assume that lll\geq{l^{\star}}: if otherwise we argue after exchanging the notations ll and l{l^{\star}}. Fix an arbitrary ε(0,l)\varepsilon\in(0,{l^{\star}}). Let

rn:=e(lε)nfor n+.r_{n}:=e^{-({l^{\star}}-\varepsilon)n}\quad\text{for $n\in{\mathbb{Z}}_{+}$}.

Let us define (recalling that 𝒛t=𝒘t𝒐\bm{z}_{t}=\bm{w}_{t}\cdot\bm{o})

Aε,n:=tn{𝒘Ω:𝒛=(z,z) exists and 𝒒(𝒛t,𝒛)rn}.A_{\varepsilon,n}:=\bigcap_{t\geq n}\Big{\{}\bm{w}\in\Omega\ :\ \text{$\bm{z}_{\infty}=(z_{\infty},z^{\star}_{\infty})$ exists and $\bm{q}(\bm{z}_{t},\bm{z}_{\infty})\leq r_{n}$}\Big{\}}.
Lemma 4.2.

The events Aε,nA_{\varepsilon,n} are increasing in n+n\in{\mathbb{Z}}_{+}, and it holds that

𝐏(n+Aε,n)=1.{\bf P}\Big{(}\bigcup_{n\in{\mathbb{Z}}_{+}}A_{\varepsilon,n}\Big{)}=1.
Proof.

By definition Aε,nA_{\varepsilon,n} are increasing in n+n\in{\mathbb{Z}}_{+}. For 𝐏{\bf P}-almost every 𝒘Ω\bm{w}\in\Omega, for all large enough t+t\in{\mathbb{Z}}_{+} (recalling that zt=wtoz_{t}=w_{t}\cdot o and zt=wtoz^{\star}_{t}=w^{\star}_{t}\cdot o^{\star}),

(zt|z)o(lε)tand(zt|z)o(lε)t.(z_{t}|z_{\infty})_{o}\geq(l-\varepsilon)t\quad\text{and}\quad(z^{\star}_{t}|z^{\star}_{\infty})_{o^{\star}}\geq({l^{\star}}-\varepsilon)t.

In which case, max{e(zt|z)o,e(zt|z)o}e(lε)t\max\{e^{-(z_{t}|z_{\infty})_{o}},e^{-(z^{\star}_{t}|z^{\star}_{\infty})_{o^{\star}}}\}\leq e^{-({l^{\star}}-\varepsilon)t} since lll\geq{l^{\star}}, showing the claim. ∎

For each η𝒳\eta\in\partial{{\mathcal{X}}^{\star}} and n+n\in{\mathbb{Z}}_{+}, let

Eε,n(η):={𝒘Ω:𝐏η([𝒘0,,𝒘n]Aε,n)en(h+ε)𝐏([𝒘0,,𝒘n]Aε,n)},E_{\varepsilon,n}(\eta):=\Big{\{}\bm{w}\in\Omega\ :\ {\bf P}^{\eta}([\bm{w}_{0},\dots,\bm{w}_{n}]\cap A_{\varepsilon,n})\leq e^{n(h^{\star}+\varepsilon)}{\bf P}([\bm{w}_{0},\dots,\bm{w}_{n}]\cap A_{\varepsilon,n})\Big{\}},

and Eε,[n,)(η):=tnEε,t(η)E_{\varepsilon,[n,\infty)}(\eta):=\bigcap_{t\geq n}E_{\varepsilon,t}(\eta).

Lemma 4.3.

For each η𝒳\eta\in\partial{{\mathcal{X}}^{\star}}, the events Eε,[n,)(η)E_{\varepsilon,[n,\infty)}(\eta) are increasing in n+n\in{\mathbb{Z}}_{+}, and for νμ\nu_{\mu^{\star}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}},

𝐏η(n+Eε,[n,)(η))=1.{\bf P}^{\eta}\Big{(}\bigcup_{n\in{\mathbb{Z}}_{+}}E_{\varepsilon,[n,\infty)}(\eta)\Big{)}=1.
Proof.

In the following, let An:=Aε,nA_{n}:=A_{\varepsilon,n} for n+n\in{\mathbb{Z}}_{+}. By (2.9), for νμ\nu_{{\mu^{\star}}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}}, for every cylinder set [𝒘0,,𝒘n][\bm{w}_{0},\dots,\bm{w}_{n}] in (Ω,,𝐏)(\Omega,{\mathcal{F}},{\bf P}),

𝐏η([𝒘0,,𝒘n])=𝐏([𝒘0,,𝒘n])d𝒘nνμdνμ(η).{\bf P}^{\eta}([\bm{w}_{0},\dots,\bm{w}_{n}])={\bf P}([\bm{w}_{0},\dots,\bm{w}_{n}])\frac{d\bm{w}_{n}\nu_{{\mu^{\star}}}}{d\nu_{{\mu^{\star}}}}(\eta).

Note that for 𝐏{\bf P}-almost every 𝒘Ω\bm{w}\in\Omega, for all n+n\in{\mathbb{Z}}_{+},

𝐏([𝒘0,,𝒘n])>0and𝐏bnd(𝒘)([𝒘0,,𝒘n])>0.{\bf P}([\bm{w}_{0},\dots,\bm{w}_{n}])>0\quad\text{and}\quad{\bf P}^{{\rm bnd}^{\star}(\bm{w})}([\bm{w}_{0},\dots,\bm{w}_{n}])>0.

The Birkhoff ergodic theorem implies that 𝐏{\bf P}-almost every 𝒘Ω\bm{w}\in\Omega,

limn1nlogd𝒘nνμdνμ(bnd(𝒘))=h.\lim_{n\to\infty}\frac{1}{n}\log\frac{d\bm{w}_{n}\nu_{{\mu^{\star}}}}{d\nu_{{\mu^{\star}}}}({\rm bnd}^{\star}(\bm{w}))=h^{\star}.

See [Kai00, the proof of Theorem 4.5]. This implies that by disintegration of 𝐏{\bf P} into 𝐏η{\bf P}^{\eta} for η𝒳\eta\in\partial{{\mathcal{X}}^{\star}}, for νμ\nu_{\mu^{\star}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}}, for 𝐏η{\bf P}^{\eta}-almost every 𝒘Ω\bm{w}\in\Omega,

limn1nlogd𝒘nνμdνμ(η)=h.\lim_{n\to\infty}\frac{1}{n}\log\frac{d\bm{w}_{n}\nu_{\mu^{\star}}}{d\nu_{\mu^{\star}}}(\eta)=h^{\star}.

Therefore for νμ\nu_{\mu^{\star}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}}, for 𝐏η{\bf P}^{\eta}-almost every 𝒘Ω\bm{w}\in\Omega,

limn1nlog𝐏η([𝒘0,,𝒘n])𝐏([𝒘0,,𝒘n])=h.\lim_{n\to\infty}\frac{1}{n}\log\frac{{\bf P}^{\eta}([\bm{w}_{0},\dots,\bm{w}_{n}])}{{\bf P}([\bm{w}_{0},\dots,\bm{w}_{n}])}=h^{\star}. (4.1)

Fix an arbitrary N+N\in{\mathbb{Z}}_{+}. For all n[N,)+n\in[N,\infty)\cap{\mathbb{Z}}_{+}, for all cylinder set [𝒘0,,𝒘n][\bm{w}_{0},\dots,\bm{w}_{n}] in (Ω,,𝐏)(\Omega,{\mathcal{F}},{\bf P}) of positive 𝐏{\bf P}-measure, it holds that

𝐏([𝒘0,,𝒘n]AN)𝐏([𝒘0,,𝒘n])𝐏([𝒘0,,𝒘n]An)𝐏([𝒘0,,𝒘n])1.\frac{{\bf P}([\bm{w}_{0},\dots,\bm{w}_{n}]\cap A_{N})}{{\bf P}([\bm{w}_{0},\dots,\bm{w}_{n}])}\leq\frac{{\bf P}([\bm{w}_{0},\dots,\bm{w}_{n}]\cap A_{n})}{{\bf P}([\bm{w}_{0},\dots,\bm{w}_{n}])}\leq 1.

The left most side equals 𝐏(ANσ(𝒘0,,𝒘n)){\bf P}(A_{N}\mid\sigma(\bm{w}_{0},\dots,\bm{w}_{n})) almost everywhere in 𝐏{\bf P}. The martingale convergence theorem yields

limn𝐏(ANσ(𝒘0,,𝒘n))=𝟏ANfor 𝐏-almost every 𝒘Ω.\lim_{n\to\infty}{\bf P}(A_{N}\mid\sigma(\bm{w}_{0},\dots,\bm{w}_{n}))={\bf 1}_{A_{N}}\quad\text{for ${\bf P}$-almost every $\bm{w}\in\Omega$}.

Since NN is arbitrary and 𝐏(N+AN)=1{\bf P}(\bigcup_{N\in{\mathbb{Z}}_{+}}A_{N})=1, it holds that

limn𝐏([𝒘0,,𝒘n]An)𝐏([𝒘0,,𝒘n])=𝟏for 𝐏-almost every 𝒘Ω.\lim_{n\to\infty}\frac{{\bf P}([\bm{w}_{0},\dots,\bm{w}_{n}]\cap A_{n})}{{\bf P}([\bm{w}_{0},\dots,\bm{w}_{n}])}={\bf 1}\quad\text{for ${\bf P}$-almost every $\bm{w}\in\Omega$}.

By disintegration of 𝐏{\bf P} into 𝐏η{\bf P}^{\eta} for η𝒳\eta\in\partial{{\mathcal{X}}^{\star}}, for νμ\nu_{\mu^{\star}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}},

limn𝐏([𝒘0,,𝒘n]An)𝐏([𝒘0,,𝒘n])=𝟏for 𝐏η-almost every 𝒘Ω.\lim_{n\to\infty}\frac{{\bf P}([\bm{w}_{0},\dots,\bm{w}_{n}]\cap A_{n})}{{\bf P}([\bm{w}_{0},\dots,\bm{w}_{n}])}={\bf 1}\quad\text{for ${\bf P}^{\eta}$-almost every $\bm{w}\in\Omega$}. (4.2)

Applying the same discussion to 𝐏η{\bf P}^{\eta} as for 𝐏{\bf P}, we obtain for νμ\nu_{\mu^{\star}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}},

limn𝐏η([𝒘0,,𝒘n]An)𝐏η([𝒘0,,𝒘n])=𝟏for 𝐏η-almost every 𝒘Ω.\lim_{n\to\infty}\frac{{\bf P}^{\eta}([\bm{w}_{0},\dots,\bm{w}_{n}]\cap A_{n})}{{\bf P}^{\eta}([\bm{w}_{0},\dots,\bm{w}_{n}])}={\bf 1}\quad\text{for ${\bf P}^{\eta}$-almost every $\bm{w}\in\Omega$}. (4.3)

Combining (4.1), (4.2) and (4.3) yields for νμ\nu_{\mu^{\star}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}},

limn1nlog𝐏η([𝒘0,,𝒘n]An)𝐏([𝒘0,,𝒘n]An)=hfor 𝐏η-almost every 𝒘Ω.\lim_{n\to\infty}\frac{1}{n}\log\frac{{\bf P}^{\eta}([\bm{w}_{0},\dots,\bm{w}_{n}]\cap A_{n})}{{\bf P}([\bm{w}_{0},\dots,\bm{w}_{n}]\cap A_{n})}=h^{\star}\quad\text{for ${\bf P}^{\eta}$-almost every $\bm{w}\in\Omega$}. (4.4)

By definition for each η𝒳\eta\in\partial{{\mathcal{X}}^{\star}} the events Eε,[n,)(η)E_{\varepsilon,[n,\infty)}(\eta) are increasing in n+n\in{\mathbb{Z}}_{+} respectively, and by (4.4) for νμ\nu_{\mu^{\star}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}},

𝐏η(n+Eε,[n,)(η))=1,{\bf P}^{\eta}\Big{(}\bigcup_{n\in{\mathbb{Z}}_{+}}E_{\varepsilon,[n,\infty)}(\eta)\Big{)}=1,

as claimed. ∎

Proof of Proposition 4.1.

Fix an arbitrary ε(0,l)\varepsilon\in(0,{l^{\star}}) and recall that rn:=e(lε)nr_{n}:=e^{-({l^{\star}}-\varepsilon)n} for n+n\in{\mathbb{Z}}_{+}. Let AnA_{n}, En(η)E_{n}(\eta) and E[n,)(η)E_{[n,\infty)}(\eta) denote Aε,nA_{\varepsilon,n}, Eε,n(η)E_{\varepsilon,n}(\eta) and Eε,[n,)(η)E_{\varepsilon,[n,\infty)}(\eta) respectively for brevity. Lemma 4.2 implies that by disintegration of 𝐏{\bf P} into 𝐏η{\bf P}^{\eta} for η𝒳\eta\in\partial{{\mathcal{X}}^{\star}}, for νμ\nu_{\mu^{\star}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}},

𝐏η(n+An)=1.{\bf P}^{\eta}\Big{(}\bigcup_{n\in{\mathbb{Z}}_{+}}A_{n}\Big{)}=1.

By this together with Lemma 4.3, for νμ\nu_{\mu^{\star}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}}, there exists an Nε,η+N_{\varepsilon,\eta}\in{\mathbb{Z}}_{+} such that

𝐏η(E[Nε,η,)(η)ANε,η)1ε.{\bf P}^{\eta}\left(E_{[N_{\varepsilon,\eta},\infty)}(\eta)\cap A_{N_{\varepsilon,\eta}}\right)\geq 1-\varepsilon. (4.5)

Let N:=Nε,ηN:=N_{\varepsilon,\eta}, E[N,)(η):=E[Nε,η,)(η)E_{[N,\infty)}(\eta):=E_{[N_{\varepsilon,\eta},\infty)}(\eta) and AN:=ANε,ηA_{N}:=A_{N_{\varepsilon,\eta}}. Further let

Fη:={𝝃𝒳×𝒳:𝐏𝝃(E[N,)(η)AN)ε}.F_{\eta}:=\Big{\{}{\bm{\xi}}\in\partial{\mathcal{X}}\times\partial{{\mathcal{X}}^{\star}}\ :\ {\bf P}^{{\bm{\xi}}}\left(E_{[N,\infty)}(\eta)\cap A_{N}\right)\geq\varepsilon\Big{\}}.

Note that FηF_{\eta} is a Borel measurable set in 𝒳×𝒳\partial{\mathcal{X}}\times\partial{{\mathcal{X}}^{\star}} since for each BB\in{\mathcal{F}} in (Ω,,𝐏)(\Omega,{\mathcal{F}},{\bf P}), the map 𝝃𝐏𝝃(B){\bm{\xi}}\mapsto{\bf P}^{{\bm{\xi}}}(B) is Borel measurable. By (4.5) and by disintegration of 𝐏η{\bf P}^{\eta} into 𝐏𝝃{\bf P}^{{\bm{\xi}}} for 𝝃𝒳×𝒳{\bm{\xi}}\in\partial{\mathcal{X}}\times\partial{{\mathcal{X}}^{\star}}, for νμ\nu_{\mu^{\star}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}},

1ε𝐏η(E[N,)(η)AN)=𝒳×𝒳𝐏𝝃(E[N,)(η)AN)𝑑νπη(𝝃)νπη(Fη)+ενπη(Fη𝖼).1-\varepsilon\leq{\bf P}^{\eta}\left(E_{[N,\infty)}(\eta)\cap A_{N}\right)=\int_{\partial{\mathcal{X}}\times\partial{{\mathcal{X}}^{\star}}}{\bf P}^{{\bm{\xi}}}\left(E_{[N,\infty)}(\eta)\cap A_{N}\right)\,d\nu_{\pi}^{\eta}({\bm{\xi}})\leq\nu_{\pi}^{\eta}(F_{\eta})+\varepsilon\nu_{\pi}^{\eta}(F_{\eta}^{\sf c}).

Therefore for νμ\nu_{\mu^{\star}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}},

νπη(Fη)12ε.\nu_{\pi}^{\eta}(F_{\eta})\geq 1-2\varepsilon. (4.6)

Furthermore, for νμ\nu_{\mu^{\star}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}} and for every 𝝃0𝒳×𝒳{\bm{\xi}}_{0}\in\partial{\mathcal{X}}\times\partial{{\mathcal{X}}^{\star}},

νπη(𝑩(𝝃0,rn)Fη)\displaystyle\nu_{\pi}^{\eta}\left(\bm{B}({\bm{\xi}}_{0},r_{n})\cap F_{\eta}\right) =𝐏η({𝒛𝑩(𝝃0,rn)Fη}E[N,)(η)AN)\displaystyle={\bf P}^{\eta}\left(\{\bm{z}_{\infty}\in\bm{B}({\bm{\xi}}_{0},r_{n})\cap F_{\eta}\}\cap E_{[N,\infty)}(\eta)\cap A_{N}\right)
+𝐏η({𝒛𝑩(𝝃0,rn)Fη}(E[N,)(η)AN)𝖼).\displaystyle\qquad+{\bf P}^{\eta}\left(\{\bm{z}_{\infty}\in\bm{B}({\bm{\xi}}_{0},r_{n})\cap F_{\eta}\}\cap(E_{[N,\infty)}(\eta)\cap A_{N})^{\sf c}\right).

By definition of FηF_{\eta}, it holds that

𝐏η({𝒛𝑩(𝝃0,rn)Fη}(E[N,)(η)AN)𝖼)\displaystyle{\bf P}^{\eta}\left(\{\bm{z}_{\infty}\in\bm{B}({\bm{\xi}}_{0},r_{n})\cap F_{\eta}\}\cap(E_{[N,\infty)}(\eta)\cap A_{N})^{\sf c}\right)
=𝑩(𝝃0,rn)Fη𝐏𝝃((E[N,)(η)AN)𝖼)𝑑νπη(𝝃)(1ε)νπη(𝑩(𝝃0,rn)Fη).\displaystyle=\int_{\bm{B}({\bm{\xi}}_{0},r_{n})\cap F_{\eta}}{\bf P}^{{\bm{\xi}}}\left((E_{[N,\infty)}(\eta)\cap A_{N})^{\sf c}\right)\,d\nu_{\pi}^{\eta}({\bm{\xi}})\leq(1-\varepsilon)\nu_{\pi}^{\eta}\left(\bm{B}({\bm{\xi}}_{0},r_{n})\cap F_{\eta}\right).

In summary, for all nNn\geq N,

ενπη(𝑩(𝝃0,rn)Fη)𝐏η({𝒛𝑩(𝝃0,rn)Fη}E[N,)(η)AN).\displaystyle\varepsilon\nu_{\pi}^{\eta}\left(\bm{B}({\bm{\xi}}_{0},r_{n})\cap F_{\eta}\right)\leq{\bf P}^{\eta}\left(\{\bm{z}_{\infty}\in\bm{B}({\bm{\xi}}_{0},r_{n})\cap F_{\eta}\}\cap E_{[N,\infty)}(\eta)\cap A_{N}\right).

By Lemmas 4.2 and 4.3, the events AnA_{n} and E[n,)(η)E_{[n,\infty)}(\eta) are increasing in n+n\in{\mathbb{Z}}_{+}, and E[n,)(η)En(η)E_{[n,\infty)}(\eta)\subset E_{n}(\eta) by the definition. Thus, for νμ\nu_{\mu^{\star}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}}, for every 𝝃0𝒳×𝒳{\bm{\xi}}_{0}\in\partial{\mathcal{X}}\times\partial{{\mathcal{X}}^{\star}}, and for all nN=Nε,ηn\geq N=N_{\varepsilon,\eta},

ενπη(𝑩(𝝃0,rn)Fη)𝐏η({𝒛𝑩(𝝃0,rn)}En(η)An).\varepsilon\nu_{\pi}^{\eta}\left(\bm{B}({\bm{\xi}}_{0},r_{n})\cap F_{\eta}\right)\leq{\bf P}^{\eta}\left(\{\bm{z}_{\infty}\in\bm{B}({\bm{\xi}}_{0},r_{n})\}\cap E_{n}(\eta)\cap A_{n}\right). (4.7)

By definition of AnA_{n}, if AnA_{n} holds, then 𝒛n𝑩(𝒛,rn)\bm{z}_{n}\in\bm{B}(\bm{z}_{\infty},r_{n}), whence for C:=C𝒒>0C:=C_{\bm{q}}>0,

𝐏η({𝒛𝑩(𝝃0,rn)}En(η)An)𝐏η({𝒛n𝑩(𝝃0,Crn)}En(η)An).{\bf P}^{\eta}\left(\{\bm{z}_{\infty}\in\bm{B}({\bm{\xi}}_{0},r_{n})\}\cap E_{n}(\eta)\cap A_{n}\right)\leq{\bf P}^{\eta}\left(\{\bm{z}_{n}\in\bm{B}({\bm{\xi}}_{0},Cr_{n})\}\cap E_{n}(\eta)\cap A_{n}\right). (4.8)

For every 𝝃0𝒳×𝒳{\bm{\xi}}_{0}\in\partial{\mathcal{X}}\times\partial{{\mathcal{X}}^{\star}} and for every n+n\in{\mathbb{Z}}_{+}, let

Bn(𝝃0):={𝒛n𝑩(𝝃0,Crn)}.B_{n}({\bm{\xi}}_{0}):=\big{\{}\bm{z}_{n}\in\bm{B}({\bm{\xi}}_{0},Cr_{n})\big{\}}.

For each such fixed 𝝃0{\bm{\xi}}_{0} and nn, the event Bn(𝝃0)B_{n}({\bm{\xi}}_{0}) is σ(𝒘0,,𝒘n)\sigma(\bm{w}_{0},\dots,\bm{w}_{n})-measurable. Further each fixed η\eta and nn, the event En(η)E_{n}(\eta) is σ(𝒘0,,𝒘n)\sigma(\bm{w}_{0},\dots,\bm{w}_{n})-measurable. Hence for each fixed 𝝃0{\bm{\xi}}_{0}, η\eta and nn, the event Bn(𝝃0)En(η)B_{n}({\bm{\xi}}_{0})\cap E_{n}(\eta) is σ(𝒘0,,𝒘n)\sigma(\bm{w}_{0},\dots,\bm{w}_{n})-measurable and is obtained as a (countable) sum of cylinder sets [𝒘0,,𝒘n][\bm{w}_{0},\dots,\bm{w}_{n}] in (Ω,,𝐏)(\Omega,{\mathcal{F}},{\bf P}). Decomposing the event into a sum of cylinder sets yields

𝐏η(Bn(𝝃0)En(η)An)=[𝒘0,,𝒘n]Bn(𝝃0)En(η)𝐏η([𝒘0,,𝒘n]An).\displaystyle{\bf P}^{\eta}\left(B_{n}({\bm{\xi}}_{0})\cap E_{n}(\eta)\cap A_{n}\right)=\sum_{[\bm{w}_{0},\dots,\bm{w}_{n}]\subset B_{n}({\bm{\xi}}_{0})\cap E_{n}(\eta)}{\bf P}^{\eta}\left([\bm{w}_{0},\dots,\bm{w}_{n}]\cap A_{n}\right).

In the right hand side, each summand is at most en(h+ε)𝐏([𝒘0,,𝒘n]An)e^{n(h^{\star}+\varepsilon)}{\bf P}\left([\bm{w}_{0},\dots,\bm{w}_{n}]\cap A_{n}\right) since [𝒘0,,𝒘n]En(η)[\bm{w}_{0},\dots,\bm{w}_{n}]\subset E_{n}(\eta). Furthermore since 𝐏([𝒘0,,𝒘n]An){\bf P}\left([\bm{w}_{0},\dots,\bm{w}_{n}]\cap A_{n}\right) over those cylinder sets add up to 𝐏(Bn(𝝃0)En(η)An){\bf P}(B_{n}({\bm{\xi}}_{0})\cap E_{n}(\eta)\cap A_{n}), it holds that

𝐏η(Bn(𝝃0)En(η)An)en(h+ε)𝐏(Bn(𝝃0)En(η)An).{\bf P}^{\eta}\left(B_{n}({\bm{\xi}}_{0})\cap E_{n}(\eta)\cap A_{n}\right)\leq e^{n(h^{\star}+\varepsilon)}{\bf P}\left(B_{n}({\bm{\xi}}_{0})\cap E_{n}(\eta)\cap A_{n}\right). (4.9)

By the definitions of Bn(𝝃0)B_{n}({\bm{\xi}}_{0}) and AnA_{n}, if Bn(𝝃0)AnB_{n}({\bm{\xi}}_{0})\cap A_{n} holds, then 𝒛𝑩(𝝃0,C2rn)\bm{z}_{\infty}\in\bm{B}({\bm{\xi}}_{0},C^{2}r_{n}). This in particular implies that

𝐏(Bn(𝝃0)En(η)An)𝐏({𝒛𝑩(𝝃0,C2rn)})=νπ(𝑩(𝝃0,C2rn)){\bf P}\left(B_{n}({\bm{\xi}}_{0})\cap E_{n}(\eta)\cap A_{n}\right)\leq{\bf P}\left(\{\bm{z}_{\infty}\in\bm{B}({\bm{\xi}}_{0},C^{2}r_{n})\}\right)=\nu_{\pi}\left(\bm{B}({\bm{\xi}}_{0},C^{2}r_{n})\right) (4.10)

Combining (4.7), (4.8), (4.9) and (4.10) implies that for νμ\nu_{\mu^{\star}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}}, for every 𝝃0𝒳×𝒳{\bm{\xi}}_{0}\in\partial{\mathcal{X}}\times\partial{{\mathcal{X}}^{\star}}, and for all nN=Nε,ηn\geq N=N_{\varepsilon,\eta},

ενπη(𝑩(𝝃0,rn)Fη)en(h+ε)νπ(𝑩(𝝃0,C2rn)).\varepsilon\nu_{\pi}^{\eta}\left(\bm{B}({\bm{\xi}}_{0},r_{n})\cap F_{\eta}\right)\leq e^{n(h^{\star}+\varepsilon)}\nu_{\pi}\left(\bm{B}({\bm{\xi}}_{0},C^{2}r_{n})\right). (4.11)

Recall that h¯=h(π)h\overline{h}=h(\pi)-h^{\star}. By Lemma 3.2, for νμ\nu_{\mu^{\star}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}}, for νπη\nu_{\pi}^{\eta}-almost every 𝝃0Fη{\bm{\xi}}_{0}\in F_{\eta}, there exists an Nη,𝝃0+N_{\eta,{\bm{\xi}}_{0}}\in{\mathbb{Z}}_{+} such that for all nNη,𝝃0n\geq N_{\eta,{\bm{\xi}}_{0}},

(h¯l+ε)logrnlogνπη(𝑩(𝝃0,rn)Fη).\left(\frac{\overline{h}}{l}+\varepsilon\right)\log r_{n}\leq\log\nu_{\pi}^{\eta}\left(\bm{B}({\bm{\xi}}_{0},r_{n})\cap F_{\eta}\right).

This together with (4.11) shows that for νμ\nu_{\mu^{\star}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}}, for νπη\nu_{\pi}^{\eta}-almost every 𝝃0Fη{\bm{\xi}}_{0}\in F_{\eta} (recalling that rn=e(lε)nr_{n}=e^{-({l^{\star}}-\varepsilon)n}),

lim supnlogνπ(𝑩(𝝃0,C2rn))logrnh¯l+ε+h+εlε.\limsup_{n\to\infty}\frac{\log\nu_{\pi}\left(\bm{B}({\bm{\xi}}_{0},C^{2}r_{n})\right)}{\log r_{n}}\leq\frac{\overline{h}}{l}+\varepsilon+\frac{h^{\star}+\varepsilon}{{l^{\star}}-\varepsilon}.

Recall that νπη(Fη)12ε\nu_{\pi}^{\eta}(F_{\eta})\geq 1-2\varepsilon by (4.6) and that this holds for arbitrary ε(0,l)\varepsilon\in(0,{l^{\star}}). Therefore after replacing the sequence rnr_{n} by reals rr tending to 0, for νμ\nu_{\mu^{\star}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}}, for νπη\nu_{\pi}^{\eta}-almost every 𝝃𝒳×𝒳{\bm{\xi}}\in\partial{\mathcal{X}}\times\partial{{\mathcal{X}}^{\star}},

lim supr0logνπ(𝑩(𝝃,r))logrh¯l+hl.\limsup_{r\to 0}\frac{\log\nu_{\pi}\left(\bm{B}\left({\bm{\xi}},r\right)\right)}{\log r}\leq\frac{\overline{h}}{l}+\frac{h^{\star}}{{l^{\star}}}.

The disintegration of νπ\nu_{\pi} into νπη\nu_{\pi}^{\eta} shows that the above holds for νπ\nu_{\pi}-almost every 𝝃𝒳×𝒳{\bm{\xi}}\in\partial{\mathcal{X}}\times\partial{{\mathcal{X}}^{\star}}, concluding the claim. ∎

4.2. Lower bounds on dimensions of harmonic measures in product spaces

Proposition 4.4.

Let 𝚪=Γ×Γ{\bm{\Gamma}}=\Gamma\times{\Gamma^{\star}} where Γ\Gamma and Γ{\Gamma^{\star}} are countable subgroups in Isom𝒳\operatorname{Isom}{\mathcal{X}} and in Isom𝒳\operatorname{Isom}{{\mathcal{X}}^{\star}} with finite exponential growth relative to (𝒳,d)({\mathcal{X}},d) and to (𝒳,d)({{\mathcal{X}}^{\star}},d^{\star}) respectively. For every probability measure π\pi on 𝚪{\bm{\Gamma}} with finite first moment and non-elementary marginals μ\mu and μ{\mu^{\star}}, it holds that for νπ\nu_{\pi}-almost every 𝛏𝒳×𝒳{\bm{\xi}}\in\partial{\mathcal{X}}\times{{\mathcal{X}}^{\star}},

lim infr0logνπ(𝑩(𝝃,r))logrh(π)h(μ)l(𝒳,μ)+h(μ)l(𝒳,μ).\liminf_{r\to 0}\frac{\log\nu_{\pi}\left(\bm{B}({\bm{\xi}},r)\right)}{\log r}\geq\frac{h(\pi)-h({\mu^{\star}})}{l({\mathcal{X}},\mu)}+\frac{h({\mu^{\star}})}{l({{\mathcal{X}}^{\star}},{\mu^{\star}})}.
Proof.

Note that 𝚪{\bm{\Gamma}} has a finite exponential growth relative to (𝒳×𝒳,𝒅)({\mathcal{X}}\times{{\mathcal{X}}^{\star}},\bm{d}) by the assumption. By Lemma 3.3, for every ε>0\varepsilon>0 there exist an N+N\in{\mathbb{Z}}_{+}, a Borel set DD in 𝒳\partial{{\mathcal{X}}^{\star}} with νμ(D)1ε\nu_{\mu^{\star}}(D)\geq 1-\varepsilon and a Borel set FF in 𝒳×𝒳\partial{\mathcal{X}}\times\partial{{\mathcal{X}}^{\star}} with νπ(F)1ε\nu_{\pi}(F)\geq 1-\varepsilon as stated. If we define F^:=(𝒳×D)F\hat{F}:=(\partial{\mathcal{X}}\times D)\cap F, then

νπ(F^)12ε.\nu_{\pi}(\hat{F})\geq 1-2\varepsilon. (4.12)

This follows since νπ(𝒳×D)=νμ(D)1ε\nu_{\pi}(\partial{\mathcal{X}}\times D)=\nu_{\mu^{\star}}(D)\geq 1-\varepsilon and νπ(F)1ε\nu_{\pi}(F)\geq 1-\varepsilon.

Let rn:=elnr_{n}:=e^{-ln} for n+n\in{\mathbb{Z}}_{+}. For every 𝝃=(ξ,ξ)𝒳×𝒳{\bm{\xi}}=(\xi,\xi^{\star})\in\partial{\mathcal{X}}\times\partial{{\mathcal{X}}^{\star}}, by disintegration of νπ\nu_{\pi} into νπη\nu_{\pi}^{\eta} for η𝒳\eta\in\partial{{\mathcal{X}}^{\star}},

νπ(B(ξ,rn)×B(ξ,rn)F^)=B(ξ,rn)Dνπη(B(ξ,rn)×𝒳F)𝑑νμ(η).\nu_{\pi}\big{(}B(\xi,r_{n})\times B(\xi^{\star},r_{n})\cap\hat{F}\big{)}=\int_{B(\xi^{\star},r_{n})\cap D}\nu_{\pi}^{\eta}\left(B(\xi,r_{n})\times\partial{{\mathcal{X}}^{\star}}\cap F\right)\,d\nu_{\mu^{\star}}(\eta).

By Lemma 3.3, for νμ\nu_{\mu^{\star}}-almost every ηD\eta\in D, for every ξ𝒳\xi\in\partial{\mathcal{X}} and for all nNn\geq N,

νπη(B(ξ,rn)×𝒳F)Cεen(h¯ε).\nu_{\pi}^{\eta}\left(B(\xi,r_{n})\times\partial{{\mathcal{X}}^{\star}}\cap F\right)\leq C_{\varepsilon}e^{-n(\overline{h}-\varepsilon)}.

Therefore for every 𝝃=(ξ,ξ)𝒳×𝒳{\bm{\xi}}=(\xi,\xi^{\star})\in\partial{\mathcal{X}}\times\partial{{\mathcal{X}}^{\star}} and for all nNn\geq N,

νπ(B(ξ,rn)×B(ξ,rn)F^)Cεen(h¯ε)νμ(B(ξ,rn)D).\nu_{\pi}\big{(}B(\xi,r_{n})\times B(\xi^{\star},r_{n})\cap\hat{F}\big{)}\leq C_{\varepsilon}e^{-n(\overline{h}-\varepsilon)}\nu_{\mu^{\star}}\left(B(\xi^{\star},r_{n})\cap D\right). (4.13)

By the dimension formula for (Γ,μ)({\Gamma^{\star}},{\mu^{\star}}) in [Tan19, Theorem 1.2], for νμ\nu_{\mu^{\star}}-almost every ξ𝒳\xi^{\star}\in\partial{{\mathcal{X}}^{\star}},

limnlogνπ(B(ξ,rn))logrn=h(μ)l.\lim_{n\to\infty}\frac{\log\nu_{\pi}\left(B(\xi^{\star},r_{n})\right)}{\log r_{n}}=\frac{h({\mu^{\star}})}{{l^{\star}}}.

(The proof presented there is for geodesic spaces, but it is adapted to roughly geodesic spaces 𝒳{\mathcal{X}}.) Since νπ(B(ξ,rn)D)νπ(B(ξ,rn))\nu_{\pi}\left(B(\xi^{\star},r_{n})\cap D\right)\leq\nu_{\pi}\left(B(\xi^{\star},r_{n})\right) and logrn<0\log r_{n}<0, for νμ\nu_{\mu^{\star}}-almost every ξ𝒳\xi^{\star}\in\partial{{\mathcal{X}}^{\star}},

lim infnlogνπ(B(ξ,rn)D)logrnh(μ)l.\liminf_{n\to\infty}\frac{\log\nu_{\pi}\left(B(\xi^{\star},r_{n})\cap D\right)}{\log r_{n}}\geq\frac{h({\mu^{\star}})}{{l^{\star}}}.

This together with (4.13) implies that since νπ(𝒳×D)=νμ(D)\nu_{\pi}(\partial{\mathcal{X}}\times D)=\nu_{\mu^{\star}}(D) and 𝑩(𝝃,rn)=B(ξ,rn)×B(ξ,rn)\bm{B}({\bm{\xi}},r_{n})=B(\xi,r_{n})\times B(\xi^{\star},r_{n}), for νπ\nu_{\pi}-almost every 𝝃=(ξ,ξ)𝒳×D{\bm{\xi}}=(\xi,\xi^{\star})\in\partial{\mathcal{X}}\times D (recalling that rn=elnr_{n}=e^{-ln}),

lim infnlogνπ(𝑩(𝝃,rn)F^)logrnh¯εl+h(μ)l.\liminf_{n\to\infty}\frac{\log\nu_{\pi}\big{(}\bm{B}({\bm{\xi}},r_{n})\cap\hat{F}\big{)}}{\log r_{n}}\geq\frac{\overline{h}-\varepsilon}{l}+\frac{h({\mu^{\star}})}{{l^{\star}}}. (4.14)

By Lemma 2.2, for νπ\nu_{\pi}-almost every 𝝃F^{\bm{\xi}}\in\hat{F} (where F^𝒳×D\hat{F}\subset\partial{\mathcal{X}}\times D),

lim infnlogνπ(𝑩(𝝃,rn))logrnh¯εl+h(μ)l.\liminf_{n\to\infty}\frac{\log\nu_{\pi}\big{(}\bm{B}({\bm{\xi}},r_{n})\big{)}}{\log r_{n}}\geq\frac{\overline{h}-\varepsilon}{l}+\frac{h({\mu^{\star}})}{{l^{\star}}}. (4.15)

By (4.12), one has νπ(F^)12ε\nu_{\pi}(\hat{F})\geq 1-2\varepsilon, and for every ε>0\varepsilon>0 there exists such an F^\hat{F} in 𝒳×𝒳\partial{\mathcal{X}}\times\partial{{\mathcal{X}}^{\star}}. Thus after replacing the sequence rnr_{n} for n+n\in{\mathbb{Z}}_{+} by positive reals rr tending to 0, we obtain for νπ\nu_{\pi}-almost every 𝝃𝒳×𝒳{\bm{\xi}}\in\partial{\mathcal{X}}\times\partial{{\mathcal{X}}^{\star}},

lim infr0logνπ(𝑩(𝝃,r))logrh¯l+h(μ)l.\liminf_{r\to 0}\frac{\log\nu_{\pi}\big{(}\bm{B}({\bm{\xi}},r)\big{)}}{\log r}\geq\frac{\overline{h}}{l}+\frac{h({\mu^{\star}})}{{l^{\star}}}.

This concludes the claim. ∎

4.3. Exact dimension and the proof of Theorem 1.1

Theorem 4.5.

Let (𝒳,d)({\mathcal{X}},d) and (𝒳,d)({{\mathcal{X}}^{\star}},d^{\star}) be roughly geodesic hyperbolic metric spaces with bounded growth at some scale. Let 𝚪=Γ×Γ{\bm{\Gamma}}=\Gamma\times{\Gamma^{\star}} where Γ\Gamma and Γ{\Gamma^{\star}} are countable subgroups in Isom𝒳\operatorname{Isom}{\mathcal{X}} and in Isom𝒳\operatorname{Isom}{{\mathcal{X}}^{\star}} with finite exponential growth relative to (𝒳,d)({\mathcal{X}},d) and to (𝒳,d)({{\mathcal{X}}^{\star}},d^{\star}) respectively. For every probability measure π\pi on 𝚪{\bm{\Gamma}} with finite first moment and non-elementary marginals μ\mu and μ{\mu^{\star}}, the harmonic measure νπ\nu_{\pi} on 𝒳×𝒳\partial{\mathcal{X}}\times\partial{{\mathcal{X}}^{\star}} is exact dimensional. Moreover, if l(𝒳,μ)l(𝒳,μ)l({\mathcal{X}},\mu)\geq l({{\mathcal{X}}^{\star}},{\mu^{\star}}), then it holds that

dimνπ=h(π)h(μ)l(𝒳,μ)+h(μ)l(𝒳,μ).\dim\nu_{\pi}=\frac{h(\pi)-h({\mu^{\star}})}{l({\mathcal{X}},\mu)}+\frac{h({\mu^{\star}})}{l({{\mathcal{X}}^{\star}},{\mu^{\star}})}.
Proof.

If Γ{\Gamma^{\star}} has finite exponential growth relative to (𝒳,d)({{\mathcal{X}}^{\star}},d^{\star}) and μ{\mu^{\star}} is non-elementary and of finite first moment, then (𝒳,νμ)(\partial{{\mathcal{X}}^{\star}},\nu_{\mu^{\star}}) is a Poisson boundary for (Γ,μ)({\Gamma^{\star}},{\mu^{\star}}) [Kai00, Theorem 7.4]. In particular, h(μ)=h(𝒳,μ)h({\mu^{\star}})=h(\partial{{\mathcal{X}}^{\star}},{\mu^{\star}}). Therefore by Propositions 4.1 and 4.4, if l(𝒳,μ)l(𝒳,μ)l({\mathcal{X}},\mu)\geq l({{\mathcal{X}}^{\star}},{\mu^{\star}}), then for νπ\nu_{\pi}-almost every 𝝃𝒳×𝒳{\bm{\xi}}\in\partial{\mathcal{X}}\times\partial{{\mathcal{X}}^{\star}},

limr0logνπ(𝑩(𝝃,r))logr=h(π)h(μ)l(𝒳,μ)+h(μ)l(𝒳,μ).\lim_{r\to 0}\frac{\log\nu_{\pi}(\bm{B}({\bm{\xi}},r))}{\log r}=\frac{h(\pi)-h({\mu^{\star}})}{l({\mathcal{X}},\mu)}+\frac{h({\mu^{\star}})}{l({{\mathcal{X}}^{\star}},{\mu^{\star}})}.

This shows that νπ\nu_{\pi} is exact dimensional. The second claim follows from Lemma 2.1. ∎

Proof of Theorem 1.1.

The claim follows from Theorem 4.5 as a special case. ∎

Remark 4.6.

Let us mention possible extensions and related questions.

  1. (1)

    The proof of Theorem 4.5 can be extended to a product of more than two hyperbolic metric spaces. For a positive N+N\in{\mathbb{Z}}_{+}, let (𝒳(i),d(i))({\mathcal{X}}^{(i)},d^{(i)}) for i=1,,Ni=1,\dots,N be proper roughly geodesic hyperbolic metric spaces with bounded growth at some scale. Further Γ(i)\Gamma^{(i)} are countable subgroups in Isom𝒳(i)\operatorname{Isom}{\mathcal{X}}^{(i)} with finite exponential growth for each i=1,,Ni=1,\dots,N. Let 𝚪:=Γ(1)××Γ(N){\bm{\Gamma}}:=\Gamma^{(1)}\times\cdots\times\Gamma^{(N)}. For a probability measure π\pi on 𝚪{\bm{\Gamma}} with non-elementary marginals μ(i)\mu^{(i)} of π\pi in Isom𝒳(i)\operatorname{Isom}{\mathcal{X}}^{(i)} of finite first moment, the harmonic measure νπ\nu_{\pi} on 𝒳(1)××𝒳(N)\partial{\mathcal{X}}^{(1)}\times\cdots\times\partial{\mathcal{X}}^{(N)} is exact dimensional: Let π(i)\pi^{(i)} be the pushforward of π\pi to Isom𝒳(i)××Isom𝒳(N)\operatorname{Isom}{\mathcal{X}}^{(i)}\times\cdots\times\operatorname{Isom}{\mathcal{X}}^{(N)} for i=1,,Ni=1,\dots,N. If l(𝒳(1),μ(1))l(𝒳(N),μ(N))l({\mathcal{X}}^{(1)},\mu^{(1)})\geq\cdots\geq l({\mathcal{X}}^{(N)},\mu^{(N)}), then for νπ\nu_{\pi}-almost every 𝝃𝒳(1)××𝒳(N){\bm{\xi}}\in\partial{\mathcal{X}}^{(1)}\times\cdots\times\partial{\mathcal{X}}^{(N)},

    dimνπ=i=1N1h(π(i))h(π(i+1))l(𝒳(i),μ(i))+h(π(N))l(𝒳(N),μ(N)).\dim\nu_{\pi}=\sum_{i=1}^{N-1}\frac{h(\pi^{(i)})-h(\pi^{(i+1)})}{l({\mathcal{X}}^{(i)},\mu^{(i)})}+\frac{h(\pi^{(N)})}{l({\mathcal{X}}^{(N)},\mu^{(N)})}.

    In the above, h(π(i))h(\pi^{(i)}) denotes the asymptotic entropy for a π(i)\pi^{(i)}-random walk and l(𝒳(i),μ(i))l({\mathcal{X}}^{(i)},\mu^{(i)}) denotes the drift associated with a μ(i)\mu^{(i)}-random walk for each i=1,,Ni=1,\dots,N. Further the Hausdorff dimension is computed by the quasi-metric defined as maximum of the ones in 𝒳(i)\partial{\mathcal{X}}^{(i)}. The proof proceeds by the reverse induction in ii from NN to 11 upon extending Propositions 4.1 and 4.4 and Theorem 4.5 to the spaces 𝒳(i)××𝒳(N){\mathcal{X}}^{(i)}\times\cdots\times{\mathcal{X}}^{(N)} for i=1,,Ni=1,\dots,N. Since writing out all the details in this generality would hurt readability, we refrain from producing the whole argument.

  2. (2)

    In [Tan19], the exact dimensionality of the harmonic measures for a single hyperbolic metric space 𝒳{\mathcal{X}} has been extended to several directions. For example, 𝒳{\mathcal{X}} can be replaced by a proper hyperbolic 𝒳{\mathcal{X}} without assuming bounded growth at some scale, and by a non-proper, separable and geodesic hyperbolic 𝒳{\mathcal{X}} with acylindrical action of a group. In those cases, probability measures μ\mu for μ\mu-random walks are assumed to satisfy that the support generates a non-elementary subgroup of isometries as a semigroup rather than a group. It is expected that results in the present paper are extended to products of such hyperbolic spaces (with right assumption on random walks). However, we need that the boundary be Polish (at least the space endowed with Borel structure be a standard Borel space) so that the conditional measures are well-defined. Thus it is not clear as to whether the separability of 𝒳{\mathcal{X}} could be dropped.

  3. (3)

    It is not clear as to whether one can remove the condition on finite exponential growth relative to each factor in Theorem 4.5. It is expected that the harmonic measure is exact dimensional without the assumption in regards of results in [HS17] and [LL23]. The issue in the present setting lies in a lack of Lebesgue differentiation theorem on boundaries. This is available, for example, if the boundaries are Euclidean spaces, more generally, Riemannian manifolds, or if the harmonic measures are doubling (which is stringent). In this paper, we have used a weaker version of Lebesgue differentiation theorem (Lemma 2.2). However, we do not know if this would suffice to remove the condition on growth. See a related question in [Tan19, Quesion 4.3].

5. A positive lower bound for dimension

5.1. Pivotal times

Let us recall the terminology and methods from [Gou22]. For δ+\delta\in{\mathbb{R}}_{+}, let (𝒳,d)({\mathcal{X}},d) be a δ\delta-hyperbolic space with a base point oo. A sequence of points x0,,xnx_{0},\dots,x_{n} is called a (C,D)(C,D)-chain for some C,D+C,D\in{\mathbb{R}}_{+} if

(xi1|xi+1)xiCfor all i=1,,n1,andd(xi1,xi)Dfor all i=1,,n.(x_{i-1}|x_{i+1})_{x_{i}}\leq C\quad\text{for all $i=1,\dots,n-1$},\quad\text{and}\quad d(x_{i-1},x_{i})\geq D\quad\text{for all $i=1,\dots,n$}.

If a sequence x0,,xnx_{0},\dots,x_{n} is a (C,D)(C,D)-chain with C+C\in{\mathbb{R}}_{+} and D2C+2δ+1D\geq 2C+2\delta+1, then

(x0|xn)x1C+δandd(x0,xn)i=0n1(d(xi,xi+1)(2C+2δ))n.(x_{0}|x_{n})_{x_{1}}\leq C+\delta\quad\text{and}\quad d(x_{0},x_{n})\geq\sum_{i=0}^{n-1}(d(x_{i},x_{i+1})-(2C+2\delta))\geq n. (5.1)

See [Gou22, Lemma 3.7]. For C+C\in{\mathbb{R}}_{+}, D=2C+2δ+1D=2C+2\delta+1 and for x,y𝒳x,y\in{\mathcal{X}}, the chain shadow 𝒞𝒮x(y,C){\mathcal{C}}{\mathcal{S}}_{x}(y,C) of yy seen from xx is the set

{z𝒳:there exists a (C,D)-chain x0,,xnx0=xxn=z and (x0|x1)yC}.\Big{\{}z\in{\mathcal{X}}\ :\ \text{there exists a $(C,D)$-chain $x_{0},\dots,x_{n}$; $x_{0}=x$, $x_{n}=z$ and $(x_{0}|x_{1})_{y}\leq C$}\Big{\}}.
Definition 5.1.

For ε,C,D+\varepsilon,C,D\in{\mathbb{R}}_{+}, a set of isometries 𝒮{\mathcal{S}} is called an (ε,C,D)(\varepsilon,C,D)-Schottky set if the following three conditions are satisfied:

  • (1)

    #{s𝒮:(x|sy)oC}(1ε)#𝒮\#\{s\in{\mathcal{S}}\ :\ (x|s\cdot y)_{o}\leq C\}\geq(1-\varepsilon)\#{\mathcal{S}} for all x,y𝒳x,y\in{\mathcal{X}},

  • (2)

    #{s𝒮:(x|s1y)oC}(1ε)#𝒮\#\{s\in{\mathcal{S}}\ :\ (x|s^{-1}\cdot y)_{o}\leq C\}\geq(1-\varepsilon)\#{\mathcal{S}} for all x,y𝒳x,y\in{\mathcal{X}}, and

  • (3)

    d(o,so)Dd(o,s\cdot o)\geq D for all s𝒮s\in{\mathcal{S}}.

Let μ\mu be a non-elementary probability measure on Isom𝒳\operatorname{Isom}{\mathcal{X}} with a countable support. It is shown basically by a classical ping-pong argument that for every ε>0\varepsilon>0 there exists a C0+C_{0}\in{\mathbb{R}}_{+} satisfying the following: For all D+D\in{\mathbb{R}}_{+} there exists an M+M\in{\mathbb{Z}}_{+} such that the support of the MM-fold convolution μM\mu^{\ast M} of μ\mu contains an (ε,C0,D)(\varepsilon,C_{0},D)-Schottky set 𝒮{\mathcal{S}} [Gou22, Corollary 3.13]. Let us fix the constants

ε=1/100,C0+andD20C0+100δ+1.\varepsilon=1/100,\quad C_{0}\in{\mathbb{R}}_{+}\quad\text{and}\quad D\geq 20C_{0}+100\delta+1.

Let λ𝒮\lambda_{\mathcal{S}} denote the uniform distribution on 𝒮{\mathcal{S}}.

Given a sequence of isometries u0,u1,u_{0},u_{1},\dots on 𝒳{\mathcal{X}} and a sequence of independent random isometries s1,s2,s_{1},s_{2},\dots with the identical distribution λ𝒮2\lambda_{\mathcal{S}}^{\ast 2}, let

yn:=u0s1u1sn1un1o.y_{n}^{-}:=u_{0}s_{1}u_{1}\cdots s_{n-1}u_{n-1}\cdot o.

Letting si=aibis_{i}=a_{i}b_{i} where aia_{i} and bib_{i} are independent and distributed as λ𝒮\lambda_{\mathcal{S}}, we define

yn:=u0s1u1sn1un1anoandyn+:=u0s1u1sn1un1anbno.y_{n}:=u_{0}s_{1}u_{1}\cdots s_{n-1}u_{n-1}a_{n}\cdot o\quad\text{and}\quad y_{n}^{+}:=u_{0}s_{1}u_{1}\cdots s_{n-1}u_{n-1}a_{n}b_{n}\cdot o.

A sequence of pivotal times Pn{1,,n}P_{n}\subseteq\{1,\dots,n\} is defined inductively as in the following: Let P0:=P_{0}:=\emptyset (the empty set). Given Pn1P_{n-1}, let k:=maxPn1k:=\max P_{n-1} if Pn1P_{n-1}\neq\emptyset, and let k=0k=0 and yk:=oy_{k}:=o if Pn1=P_{n-1}=\emptyset. Suppose that Pn1P_{n-1}\neq\emptyset. Let us say that the local geodesic condition is satisfied at time nn if

(yk|yn)ynC0,(yn|yn+)ynC0and(yn|yn+1)yn+C0.(y_{k}|y_{n})_{y_{n}^{-}}\leq C_{0},\quad(y_{n}^{-}|y_{n}^{+})_{y_{n}}\leq C_{0}\quad\text{and}\quad(y_{n}|y_{n+1}^{-})_{y_{n}^{+}}\leq C_{0}. (5.2)

If the local geodesic condition is satisfied at time nn, then we define

Pn:=Pn1{n}.P_{n}:=P_{n-1}\cup\{n\}.

If otherwise, then letting mm be the largest pivotal time in Pn1P_{n-1} such that

yn+1𝒞𝒮ym(ym+,C0+δ),y_{n+1}^{-}\in{\mathcal{C}}{\mathcal{S}}_{y_{m}}(y_{m}^{+},C_{0}+\delta),

we define Pn:=Pn1{1,,m}P_{n}:=P_{n-1}\cap\{1,\dots,m\}, and Pn:=P_{n}:=\emptyset in the case when there is no such mm. Note that the set PnP_{n} depends only on the sequence s1,,sns_{1},\dots,s_{n} for fixed u0,,unu_{0},\dots,u_{n}.

Lemma 5.2.

If Pn:={k1,,kp}P_{n}:=\{k_{1},\dots,k_{p}\} where k1<<kpk_{1}<\cdots<k_{p}, then the sequence

o,yk1,yk2,yk2,,ykp,ykp,yn+1,o,y_{k_{1}},y_{k_{2}}^{-},y_{k_{2}},\dots,y_{k_{p}}^{-},y_{k_{p}},y_{n+1}^{-},

forms a (2C0+4δ,D2C03δ)(2C_{0}+4\delta,D-2C_{0}-3\delta)-chain. Moreover, if D6C0+13δ+1D\geq 6C_{0}+13\delta+1, then for every i=2,,pi=2,\dots,p, the sequence yki,yki,yn+1y_{k_{i}}^{-},y_{k_{i}},y_{n+1}^{-} is a (2C0+5δ,D6C013δ)(2C_{0}+5\delta,D-6C_{0}-13\delta)-chain.

Proof.

The first claim is [Gou22, Lemma 4.4]. The second claim follows from the first claim and (5.1). Indeed, applying to them the sequence yki,yki,,ykp,yn+1y_{k_{i}}^{-},y_{k_{i}},\dots,y_{k_{p}},y_{n+1}^{-} for each i=2,,pi=2,\dots,p shows that (yki|yn+1)yki2C0+5δ(y_{k_{i}}^{-}|y_{n+1}^{-})_{y_{k_{i}}}\leq 2C_{0}+5\delta, and further,

d(yki,yn+1)d(yki,yki+1)2(2C0+4δ)2δD6C013δ.d(y_{k_{i}},y_{n+1}^{-})\geq d(y_{k_{i}},y_{k_{i+1}}^{-})-2(2C_{0}+4\delta)-2\delta\geq D-6C_{0}-13\delta.

The claim follows. ∎

Let s¯=(s1,,sn)\overline{s}=(s_{1},\dots,s_{n}). Let us say that a sequence s¯=(s1,,sn)\overline{s}^{\prime}=(s_{1}^{\prime},\dots,s_{n}^{\prime}) where si=aibis_{i}^{\prime}=a_{i}^{\prime}b_{i}^{\prime} is pivoted from s¯\overline{s} if s¯\overline{s}^{\prime} and s¯\overline{s} have the same pivotal times, bi=bib_{i}^{\prime}=b_{i} for all i=1,,ni=1,\dots,n, and ai=aia_{i}^{\prime}=a_{i} for all ii which is not a pivotal time. The relation that s¯\overline{s}^{\prime} is pivoted from s¯\overline{s} defines an equivalence relation among sequences. In this notation, we understand that bib_{i} for all i=1,,ni=1,\dots,n and aia_{i} for ii which is not a pivotal time are determined in s¯\overline{s}.

Let n(s¯){\mathcal{E}}_{n}(\overline{s}) be the set of sequences which are pivoted from s¯\overline{s}. Note that if u0,u1,u_{0},u_{1},\dots are fixed, then conditioned on n(s¯){\mathcal{E}}_{n}(\overline{s}), all aia_{i} are independent. However, their distributions may depend on ii. For each i=1,,ni=1,\dots,n, let

Ai(s¯):={a𝒮:si=abi for some s¯=(s1,,si,,sn)n(s¯)}.A_{i}(\overline{s}):=\big{\{}a\in{\mathcal{S}}\ :\ \text{$s_{i}^{\prime}=ab_{i}$ for some $\overline{s}^{\prime}=(s_{1}^{\prime},\dots,s_{i}^{\prime},\dots,s_{n}^{\prime})\in{\mathcal{E}}_{n}(\overline{s})$}\big{\}}.

It holds that for each pivotal time ii of s¯\overline{s},

𝐏(ai=an(s¯))=λ𝒮(aAi(s¯))=λ𝒮(a)λ𝒮(Ai(s¯))for aAi(s¯).{\bf P}(a_{i}=a\mid{\mathcal{E}}_{n}(\overline{s}))=\lambda_{\mathcal{S}}(a\mid A_{i}(\overline{s}))=\frac{\lambda_{\mathcal{S}}(a)}{\lambda_{\mathcal{S}}(A_{i}(\overline{s}))}\quad\text{for $a\in A_{i}(\overline{s})$}.

If ii is a pivotal time of s¯\overline{s} and s¯=(s1,,si,,sn)\overline{s}^{\prime}=(s_{1},\dots,s_{i}^{\prime},\dots,s_{n}) in which si=aibis_{i}=a_{i}b_{i} is replaced by si=aibis_{i}^{\prime}=a_{i}^{\prime}b_{i} satisfies the local geodesic condition at time ii, then s¯\overline{s}^{\prime} is pivoted from s¯\overline{s} [Gou22, Lemma 4.7]. By the definition of (ε,C0,D)(\varepsilon,C_{0},D)-Schottky set (Definition 5.1), there are at most 2ε#𝒮2\varepsilon\#{\mathcal{S}} elements for which the local geodesic condition (5.2) does not hold at ii. Therefore for each pivotal time ii in s¯\overline{s},

#Ai(s¯)(12ε)#𝒮.\#A_{i}(\overline{s})\geq(1-2\varepsilon)\#{\mathcal{S}}. (5.3)
Lemma 5.3.

Let D10C0+25δ+1D\geq 10C_{0}+25\delta+1. For s¯n(s¯)\overline{s}^{\prime}\in{\mathcal{E}}_{n}(\overline{s}), if yn+1=yn+1y_{n+1}^{-}=y_{n+1}^{\prime-}, then s¯=s¯\overline{s}=\overline{s}^{\prime}.

Proof.

Let Pn={k1,,kp}P_{n}=\{k_{1},\dots,k_{p}\} with k1<<kpk_{1}<\cdots<k_{p} be the set of pivotal times in s¯\overline{s}. By Lemma 5.2, the sequences yki,yki,yn+1y_{k_{i}}^{-},y_{k_{i}},y_{n+1}^{-} and yki,yki,yn+1y_{k_{i}}^{\prime-},y_{k_{i}}^{\prime},y_{n+1}^{\prime-} are (2C0+5δ,D6C013δ)(2C_{0}+5\delta,D-6C_{0}-13\delta)-chains respectively. For s¯n(s¯)\overline{s}^{\prime}\in{\mathcal{E}}_{n}(\overline{s}) with s¯s¯\overline{s}^{\prime}\neq\overline{s}, let ii be the first ii for which skiskis_{k_{i}}\neq s_{k_{i}}^{\prime}. For ski=akibkis_{k_{i}}=a_{k_{i}}b_{k_{i}} and ski=akibkis_{k_{i}}^{\prime}=a_{k_{i}}^{\prime}b_{k_{i}}^{\prime}, it holds that akiakia_{k_{i}}\neq a_{k_{i}^{\prime}} and these akia_{k_{i}} and akia_{k_{i}}^{\prime} are in the Schottky set 𝒮{\mathcal{S}}, whence (akio|akio)oC0(a_{k_{i}}\cdot o|a_{k_{i}}^{\prime}\cdot o)_{o}\leq C_{0}. This shows that the sequence

yn+1,yki,yki,yki,yn+1,where yki=yki,y_{n+1}^{\prime},y_{k_{i}}^{\prime},y_{k_{i}}^{-},y_{k_{i}},y_{n+1}^{-},\quad\text{where $y_{k_{i}}^{-}=y_{k_{i}}^{\prime-}$},

forms a (2C0+5δ,D6C013δ)(2C_{0}+5\delta,D-6C_{0}-13\delta)-chain. For such DD, one has d(yn+1,yn+1)>0d(y_{n+1}^{\prime-},y_{n+1}^{-})>0 by (5.1), and thus yn+1yn+1y_{n+1}^{\prime-}\neq y_{n+1}^{-}, as required. ∎

5.2. A lower bound for entropy

Theorem 5.4.

Let Γ\Gamma and Γ{\Gamma^{\star}} be countable subgroups in Isom𝒳\operatorname{Isom}{\mathcal{X}} and in Isom𝒳\operatorname{Isom}{{\mathcal{X}}^{\star}} respectively, and 𝚪:=Γ×Γ{\bm{\Gamma}}:=\Gamma\times{\Gamma^{\star}}. Further let us consider a probability measure π\pi on 𝚪{\bm{\Gamma}} of the following form:

π=αλ×λ+(1α)π0\pi=\alpha\lambda\times{\lambda^{\star}}+(1-\alpha)\pi_{0}

for some α(0,1]\alpha\in(0,1], non-elementary probability measures λ\lambda and λ{\lambda^{\star}} on Γ\Gamma and on Γ{\Gamma^{\star}} respectively, and a probability measure π0\pi_{0} on 𝚪{\bm{\Gamma}}. Then for the marginal μ{\mu^{\star}} of π\pi on Γ{\Gamma^{\star}}, it holds that h(π)h(𝒳,μ)>0h(\pi)-h(\partial{{\mathcal{X}}^{\star}},{\mu^{\star}})>0. Moreover, if in addition 𝚪{\bm{\Gamma}} has a finite exponential growth relative to (𝒳,𝐝)({\mathcal{X}},\bm{d}), then for νμ\nu_{{\mu^{\star}}}-almost every η𝒳\eta\in\partial{{\mathcal{X}}^{\star}}, the Hausdorff dimension of the conditional measure νπη\nu_{\pi}^{\eta} is positive.

Fix constants ε=1/100\varepsilon=1/100, C00C_{0}\geq 0 and D20C0+100δ+1D\geq 20C_{0}+100\delta+1, and a (1/100,C0,D)(1/100,C_{0},D)-Schottky set 𝒮{\mathcal{S}} in Γ\Gamma, contained in the support of λM\lambda^{\ast M} for some M+M\in{\mathbb{Z}}_{+}. For N:=2MN:=2M, let us write for some β(0,1]\beta\in(0,1] and for a probability measure λ0\lambda_{0} on Γ\Gamma,

λN=βλ𝒮2+(1β)λ0.\lambda^{\ast N}=\beta\lambda_{\mathcal{S}}^{\ast 2}+(1-\beta)\lambda_{0}.

Let us also write for a probability measure π¯0\overline{\pi}_{0} on 𝚪=Γ×Γ{\bm{\Gamma}}=\Gamma\times{\Gamma^{\star}},

πN=αNβλ𝒮2×λN+(1αNβ)π¯0.\pi^{\ast N}=\alpha^{N}\beta\lambda_{\mathcal{S}}^{\ast 2}\times{\lambda^{\star}}^{\ast N}+(1-\alpha^{N}\beta)\overline{\pi}_{0}.

For a sequence ε1,ε2,\varepsilon_{1},\varepsilon_{2},\dots of independent Bernoulli random variables with the common parameter αNβ\alpha^{N}\beta, let us define a sequence of independent random group elements 𝜸1,𝜸2,{\bm{\gamma}}_{1},{\bm{\gamma}}_{2},\dots where 𝜸i=(γi,γi)𝚪{\bm{\gamma}}_{i}=(\gamma_{i},\gamma^{\star}_{i})\in{\bm{\Gamma}} is distributed as λ𝒮2×λN\lambda_{\mathcal{S}}^{\ast 2}\times{\lambda^{\star}}^{\ast N} if εi=1\varepsilon_{i}=1 and as π¯0\overline{\pi}_{0} if εi=0\varepsilon_{i}=0. Note in particular that for i=1,2,i=1,2,\dots, conditioned on the event {εi=1}\{\varepsilon_{i}=1\}, random group elements γi\gamma_{i} and γi\gamma^{\star}_{i} are independent.

Further we realize a π\pi-random walk 𝒘nN\bm{w}_{nN} at time nNnN by a product 𝜸1𝜸n{\bm{\gamma}}_{1}\cdots{\bm{\gamma}}_{n} through a coupling on an enlarged probability space. Let us define a sequence of pivotal times for zn=wnoz_{n}=w_{n}\cdot o on 𝒳{\mathcal{X}}, where {wn}n+\{w_{n}\}_{n\in{\mathbb{Z}}_{+}} is a μ\mu-random walk and μ\mu is the marginal of π\pi on Γ\Gamma.

Let t1,t2,t_{1},t_{2},\dots be the sequence of ii such that εi=1\varepsilon_{i}=1. For every positive n+n\in{\mathbb{Z}}_{+}, let τ=τ(n)\tau=\tau(n) be the maximum of jj with NtjnNt_{j}\leq n. It holds that

(N(tj1),Ntj](0,n]for all j=1,,τ.(N(t_{j}-1),Nt_{j}]\subset(0,n]\quad\text{for all $j=1,\dots,\tau$}.

For each j=1,,τj=1,\dots,\tau, let stjs_{t_{j}} be γj\gamma_{j}, which is realized as the product of elements xix_{i} over i(N(tj1),Ntj]i\in(N(t_{j}-1),Nt_{j}] in the natural order from +{\mathbb{Z}}_{+}. Let us write sj:=stjs_{j}^{\prime}:=s_{t_{j}} for brevity, and

u0:=x1xN(t11),uj:=xNtj+1xN(tj+11)andu(n):=xNtτ+1xn.u_{0}:=x_{1}\cdots x_{N(t_{1}-1)},\quad u_{j}:=x_{Nt_{j}+1}\cdots x_{N(t_{j+1}-1)}\quad\text{and}\quad u(n):=x_{Nt_{\tau}+1}\cdots x_{n}.

In the above, u0u_{0}, uju_{j} and u(n)u(n) are defined as the identity if they are empty words. For a μ\mu-random walk wnw_{n} at time nn, the orbit znz_{n} is realized as

zn=wno=u0s1u1sτu(n)o.z_{n}=w_{n}\cdot o=u_{0}s_{1}^{\prime}u_{1}\cdots s_{\tau}^{\prime}u(n)\cdot o.

Let P1,,Pτ(n)P_{1},\dots,P_{\tau(n)} be the sequence of pivotal times of znz_{n} given u0,u1,,u(n)u_{0},u_{1},\dots,u(n). Note that Pτ(n)P_{\tau(n)} depends not only on τ(n)\tau(n) but also on nn. It is shown that there exists a constant κ>0\kappa>0 such that

𝐏(#Pτ(n)κn)eκnfor all n+,{\bf P}(\#P_{\tau(n)}\leq\kappa n)\leq e^{-\kappa n}\quad\text{for all $n\in{\mathbb{Z}}_{+}$}, (5.4)

[Gou22, Proposition 4.11].

Note that the sequence {εi}i=1\{\varepsilon_{i}\}_{i=1}^{\infty} determines {ti}i=1\{t_{i}\}_{i=1}^{\infty} and τ=τ(n)\tau=\tau(n) for each nn. Let us define the σ\sigma-algebra

𝒢:=σ(εi,ti,xi for i=1,2, and xi for ij=1τ(n)(N(tj1),Ntj]).{\mathcal{G}}:=\sigma\Big{(}\text{$\varepsilon_{i},t_{i},x^{\star}_{i}$ for $i=1,2,\dots$ and $x_{i}$ for $i\notin\bigcup_{j=1}^{\tau(n)}(N(t_{j}-1),Nt_{j}]$}\Big{)}.

Conditioning on 𝒢{\mathcal{G}} amounts to fixing a typical sequence {εi}i=1\{\varepsilon_{i}\}_{i=1}^{\infty}, {ti}i=1\{t_{i}\}_{i=1}^{\infty} and a trajectory of μ{\mu^{\star}}-random walk {wn}n+\{w^{\star}_{n}\}_{n\in{\mathbb{Z}}_{+}}, increments xix_{i} of μ\mu-random walk outside the time intervals (N(tj1),Ntj](N(t_{j}-1),Nt_{j}] for j=1,,τ(n)j=1,\dots,\tau(n). Under this conditioning, s1,,sτs^{\prime}_{1},\dots,s^{\prime}_{\tau} is a sequence of independent random elements in Γ\Gamma with the common distribution λ𝒮2\lambda_{\mathcal{S}}^{\ast 2}.

Let s¯:=(s1,,sτ)\overline{s}:=(s^{\prime}_{1},\dots,s^{\prime}_{\tau}), and τ(s¯){\mathcal{E}}_{\tau}(\overline{s}) be the set of sequences pivoted from s¯\overline{s}. Conditioned on τ(s¯){\mathcal{E}}_{\tau}(\overline{s}) and 𝒢{\mathcal{G}}, the random group elements aia_{i} at pivotal times where si=aibis^{\prime}_{i}=a_{i}b_{i} are independent and each aia_{i} is distributed as λ𝒮(Ai(s¯))\lambda_{\mathcal{S}}(\,\cdot\mid A_{i}(\overline{s})). Let σ(τ(s¯),𝒢)\sigma({\mathcal{E}}_{\tau}(\overline{s}),{\mathcal{G}}) denote the σ\sigma-algebra generated by τ(s¯){\mathcal{E}}_{\tau}(\overline{s}) and 𝒢{\mathcal{G}}. Given τ(s¯){\mathcal{E}}_{\tau}(\overline{s}) and u0,,u(n)u_{0},\dots,u(n), let us define the map

iPτ(n)Ai(s¯)𝒳,(ai)iPτ(n)wno=u0a1b1u1uτ1aτbτu(n)o.\prod_{i\in P_{\tau(n)}}A_{i}(\overline{s})\to{\mathcal{X}},\quad(a_{i})_{i\in P_{\tau(n)}}\mapsto w_{n}\cdot o=u_{0}a_{1}b_{1}u_{1}\cdots u_{\tau-1}a_{\tau}b_{\tau}u(n)\cdot o.

In the above, we understand that {bi}i=1,,τ(n)\{b_{i}\}_{i=1,\dots,\tau(n)} and {ai}iPτ(n)\{a_{i}\}_{i\notin P_{\tau(n)}} are determined by τ(s¯){\mathcal{E}}_{\tau}(\overline{s}). Under the conditioning, the map is injective by Lemma 5.3. Furthermore the conditional distribution 𝐏(wnoσ(τ(s¯),𝒢)){\bf P}(w_{n}\cdot o\in\cdot\mid\sigma({\mathcal{E}}_{\tau}(\overline{s}),{\mathcal{G}})) is the pushforward by the injective map of the product measure λ𝒮(Ai(s¯))\lambda_{\mathcal{S}}(\,\cdot\mid A_{i}(\overline{s})) over iPτ(n)i\in P_{\tau(n)} almost everywhere in 𝐏{\bf P}.

Proof of Theorem 5.4.

Let us denote by H(w)H(w) the entropy H(μ)H(\mu) for a random variable ww with the distribution μ\mu. If {\mathcal{F}} is a sub σ\sigma-algebra of σ({(wn,wn)}n+)\sigma(\{(w_{n},w^{\star}_{n})\}_{n\in{\mathbb{Z}}_{+}}), then the conditional entropy of wnw_{n} with respect to {\mathcal{F}} is defined by

H(wn):=𝐄[xsuppμn𝐏(wn=x)log𝐏(wn=x)].H(w_{n}\mid{\mathcal{F}}):={\bf E}\,\Big{[}-\sum_{x\in{\rm supp}\,\mu_{n}}{\bf P}(w_{n}=x\mid{\mathcal{F}})\log{\bf P}(w_{n}=x\mid{\mathcal{F}})\Big{]}.

It holds that H(wn)H(wn)H(w_{n})\geq H(w_{n}\mid{\mathcal{F}}), further that if 1{\mathcal{F}}_{1} and 2{\mathcal{F}}_{2} are sub σ\sigma-algebras of σ({(wn,wn)}n+)\sigma(\{(w_{n},w^{\star}_{n})\}_{n\in{\mathbb{Z}}_{+}}) and 12{\mathcal{F}}_{1}\subseteq{\mathcal{F}}_{2}, then H(wn1)H(wn2)H(w_{n}\mid{\mathcal{F}}_{1})\geq H(w_{n}\mid{\mathcal{F}}_{2}). Since the σ\sigma-algebra σ(wn)\sigma(w^{\star}_{n}) generated by wnw^{\star}_{n} is included in 𝒢{\mathcal{G}}, it follows that

H(wnσ(wn))H(wnσ(τ(s¯),𝒢)).H(w_{n}\mid\sigma(w^{\star}_{n}))\geq H(w_{n}\mid\sigma({\mathcal{E}}_{\tau}(\overline{s}),{\mathcal{G}})).

Let us find a lower bound on the right hand side of the following:

H(wnσ(τ(s¯),𝒢))=𝐄[xsuppμn𝐏(wn=xσ(τ(s¯),𝒢))log𝐏(wn=xσ(τ(s¯),𝒢))].H(w_{n}\mid\sigma({\mathcal{E}}_{\tau}(\overline{s}),{\mathcal{G}}))={\bf E}\,\Big{[}-\sum_{x\in{\rm supp}\,\mu_{n}}{\bf P}\left(w_{n}=x\mid\sigma({\mathcal{E}}_{\tau}(\overline{s}),{\mathcal{G}})\right)\log{\bf P}\left(w_{n}=x\mid\sigma({\mathcal{E}}_{\tau}(\overline{s}),{\mathcal{G}})\right)\Big{]}.

Since 𝐏(wnoσ(τ(s¯),𝒢)){\bf P}(w_{n}\cdot o\in\cdot\mid\sigma({\mathcal{E}}_{\tau}(\overline{s}),{\mathcal{G}})) is the pushforward by an injective map of the product measure of λ𝒮(Ai(s¯))\lambda_{\mathcal{S}}(\cdot\mid A_{i}(\overline{s})) over iPτ(n)i\in P_{\tau(n)}, one has 𝐏{\bf P}-almost everywhere,

xsuppμn𝐏(wn=xσ(τ(s¯),𝒢))log𝐏(wn=xσ(τ(s¯),𝒢))=iPτ(n)log1#Ai(s¯).-\sum_{x\in{\rm supp}\,\mu_{n}}{\bf P}(w_{n}=x\mid\sigma({\mathcal{E}}_{\tau}(\overline{s}),{\mathcal{G}}))\log{\bf P}(w_{n}=x\mid\sigma({\mathcal{E}}_{\tau}(\overline{s}),{\mathcal{G}}))=-\sum_{i\in P_{\tau(n)}}\log\frac{1}{\#A_{i}(\overline{s})}.

This shows that for each n+n\in{\mathbb{Z}}_{+} and for κ>0\kappa>0 in (5.4),

H(wnσ(wn))𝐄[(iPτ(n)log#Ai(s¯))𝟏{#Pτ(n)κn}].H(w_{n}\mid\sigma(w^{\star}_{n}))\geq{\bf E}\,\Big{[}\Big{(}\sum_{i\in P_{\tau(n)}}\log\#A_{i}(\overline{s})\Big{)}\cdot{\bf 1}_{\{\#P_{\tau(n)}\geq\kappa n\}}\Big{]}.

If iPτ(n)i\in P_{\tau(n)}, then #Ai(s¯)(12ε)#𝒮\#A_{i}(\overline{s})\geq(1-2\varepsilon)\#{\mathcal{S}} by (5.3) and 𝐏(#Pτ(n)κn)eκn{\bf P}(\#P_{\tau(n)}\leq\kappa n)\leq e^{-\kappa n} by (5.4), the right hand side of the above inequality is at least

κnlog((12ε)#𝒮)𝐏(#Pτ(n)κn)κnlog((12ε)#𝒮)(1eκn).\kappa n\log((1-2\varepsilon)\#{\mathcal{S}})\cdot{\bf P}(\#P_{\tau(n)}\geq\kappa n)\geq\kappa n\log((1-2\varepsilon)\#{\mathcal{S}})\cdot(1-e^{-\kappa n}).

Therefore it holds that

lim infn1nH(wnσ(wn))κlog((12ε)#𝒮)>0.\liminf_{n\to\infty}\frac{1}{n}H(w_{n}\mid\sigma(w^{\star}_{n}))\geq\kappa\log((1-2\varepsilon)\#{\mathcal{S}})>0.

Noting that H(wn,wn)=H(wn)+H(wnσ(wn))H(w_{n},w^{\star}_{n})=H(w^{\star}_{n})+H(w_{n}\mid\sigma(w^{\star}_{n})) and h(𝒳,μ)h(μ)h(\partial{{\mathcal{X}}^{\star}},{\mu^{\star}})\leq h({\mu^{\star}}), we obtain

h(π)h(𝒳,μ)h(π)h(μ)lim infn1nH(wnσ(wn))>0.h(\pi)-h(\partial{{\mathcal{X}}^{\star}},{\mu^{\star}})\geq h(\pi)-h({\mu^{\star}})\geq\liminf_{n\to\infty}\frac{1}{n}H(w_{n}\mid\sigma(w^{\star}_{n}))>0.

(In the above, the second inequality is in fact the equality and the liminf is the limit.) Thus the first claim follows. The second claim follows from the first claim together with Theorem 1.3. ∎

Proof of Theorem 1.4.

This is contained in Theorem 5.4. ∎

Acknowledgments

The author would like to thank the anonymous referee for a meticulous review and suggestions. The author is partially supported by JSPS Grant-in-Aid for Scientific Research JP20K03602 and JP24K06711.

References

  • [Ass83] Patrice Assouad. Plongements lipschitziens dans 𝐑n{\bf R}^{n}. Bull. Soc. Math. France, 111(4):429–448, 1983.
  • [BB23] Itai Benjamini and Jérémie Brieussel. Noise sensitivity of random walks on groups. ALEA Lat. Am. J. Probab. Math. Stat., 20(2):1139–1164, 2023.
  • [BHM11] Sébastien Blachère, Peter Haissinsky, and Pierre Mathieu. Harmonic measures versus quasiconformal measures for hyperbolic groups. Ann. Sci. Éc. Norm. Supér. (4), 44(4):683–721, 2011.
  • [BS00] M. Bonk and O. Schramm. Embeddings of Gromov hyperbolic spaces. Geom. Funct. Anal., 10(2):266–306, 2000.
  • [Der80] Y. Derriennic. Quelques applications du théorème ergodique sous-additif. In Conference on Random Walks (Kleebach, 1979), volume 74 of Astérisque, pages 183–201, Paris, 1980. Soc. Math. France.
  • [Der86] Y. Derriennic. Entropie, théorèmes limite et marches aléatoires. In Probability measures on groups, VIII (Oberwolfach, 1985), volume 1210 of Lecture Notes in Math., pages 241–284. Springer, Berlin, 1986.
  • [DY23] Matthieu Dussaule and Wenyuan Yang. The Hausdorff dimension of the harmonic measure for relatively hyperbolic groups. Trans. Amer. Math. Soc. Ser. B, 10:766–806, 2023.
  • [Fen23] De-Jun Feng. Dimension of invariant measures for affine iterated function systems. Duke Math. J., 172(4):701–774, 2023.
  • [FH09] De-Jun Feng and Huyi Hu. Dimension theory of iterated function systems. Comm. Pure Appl. Math., 62(11):1435–1500, 2009.
  • [Fur08] Hillel Furstenberg. Ergodic fractal measures and dimension conservation. Ergodic Theory Dynam. Systems, 28(2):405–422, 2008.
  • [Gou22] Sébastien Gouëzel. Exponential bounds for random walks on hyperbolic spaces without moment conditions. Tunis. J. Math., 4:635–671, 2022.
  • [Gro87] M. Gromov. Hyperbolic groups. In S. M. Gersten, editor, Essays in group theory, volume 8 of Mathematical Sciences Research Institute Publications, pages 75–263, New York, 1987. Springer-Verlag.
  • [Hei01] Juha Heinonen. Lectures on analysis on metric spaces. Universitext. Springer-Verlag, New York, 2001.
  • [HS17] Michael Hochman and Boris Solomyak. On the dimension of Furstenberg measure for SL2()SL_{2}(\mathbb{R}) random matrix products. Invent. Math., 210(3):815–875, 2017.
  • [Kai98] Vadim A. Kaimanovich. Hausdorff dimension of the harmonic measure on trees. Ergodic Theory Dynam. Systems, 18(3):631–660, 1998.
  • [Kai00] Vadim A. Kaimanovich. The Poisson formula for groups with hyperbolic properties. Ann. of Math. (2), 152(3):659–692, 2000.
  • [KLP11] Vadim A. Kaimanovich and Vincent Le Prince. Matrix random products with singular harmonic measure. Geom. Dedicata, 150:257–279, 2011.
  • [KP96] R. Kenyon and Y. Peres. Measures of full dimension on affine-invariant sets. Ergodic Theory Dynam. Systems, 16(2):307–323, 1996.
  • [KV83] V. A. Kaimanovich and A. M. Vershik. Random walks on discrete groups: boundary and entropy. Ann. Probab., 11(3):457–490, 1983.
  • [Led01] François Ledrappier. Some asymptotic properties of random walks on free groups. In Topics in probability and Lie groups: boundary theory, volume 28 of CRM Proc. Lecture Notes, pages 117–152. Amer. Math. Soc., Providence, RI, 2001.
  • [Les21] Pablo Lessa. Entropy and dimension of disintegrations of stationary measures. Trans. Amer. Math. Soc. Ser. B, 8:105–129, 2021.
  • [LL23] François Ledrappier and Pablo Lessa. Exact dimension of Furstenberg measures. Geom. Funct. Anal., 33(1):245–298, 2023.
  • [Rap21] Ariel Rapaport. Exact dimensionality and Ledrappier-Young formula for the Furstenberg measure. Trans. Amer. Math. Soc., 374(7):5225–5268, 2021.
  • [Tan19] Ryokichi Tanaka. Dimension of harmonic measures in hyperbolic spaces. Ergodic Theory Dynam. Systems, 39(2):474–499, 2019.
  • [Tan24] Ryokichi Tanaka. Non-noise sensitivity for word hyperbolic groups. Ann. Fac. Sci. Toulouse Math. (6), 33(5):1487–1510, 2024.
  • [Vol21] Oleksii Volkov. Random walks on products of hyperbolic groups. PhD thesis, University of Ottawa, 2021.