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

Genericity on submanifolds and application to Universal hitting time statistics

Han Zhang Yau Mathematical Sciences Center, Tsinghua University, Beijing, 100084, China [email protected]
Abstract.

We investigate Birkhoff genericity on submanifolds of homogeneous space X=SLd()(d)k/SLd()(d)kX=SL_{d}(\mathbb{R})\ltimes(\mathbb{R}^{d})^{k}/SL_{d}(\mathbb{Z})\ltimes(\mathbb{Z}^{d})^{k}, where d2d\geq 2 and k1k\geq 1 are fixed integers. The submanifolds we consider are parameterized by unstable horospherical subgroup UU of a diagonal flow ata_{t} in SLd()SL_{d}(\mathbb{R}). As long as the intersection of the submanifold with any affine rational subspace has Lebesgue measure zero, we show that the trajectory of ata_{t} along Lebesgue almost every point on the submanifold gets equidistributed on XX. This generalizes the previous work of Frączek, Shi and Ulcigrai in [8].

Following the scheme developed by Dettmann, Marklof and Strömbergsson in [3], we then deduce an application to universal hitting time statistics for integrable flows.

1. Introduction

Let (X,,μ,R)(X,\mathcal{B},\mu,R) be a probability measure preserving system, where (X,)(X,\mathcal{B}) is a Borel measurable space with probability measure μ\mu, and Rt:XXR^{t}:X\to X is an \mathbb{R}-action (or \mathbb{Z}-action) preserving μ\mu. Assume that RR is ergodic, then Birkhoff’s ergodic theorem (cf. [4]) asserts that, for any fLμ1(X)f\in L^{1}_{\mu}(X),

1T0Tf(Rtx)dtTXfdμ, (or 1Nn=1Nf(Rnx)NXfdμ),\displaystyle\frac{1}{T}\int_{0}^{T}f(R^{t}x)dt\xrightarrow{T\to\infty}\int_{X}fd\mu,\text{ (or }\frac{1}{N}\sum_{n=1}^{N}f(R^{n}x)\xrightarrow{N\to\infty}\int_{X}fd\mu), (1.1)

for μ\mu-almost every xXx\in X. In particular, (1.1) holds for any fCc(X)f\in C_{c}(X), where Cc(X)C_{c}(X) denotes the collection of all compactly supported and continuous functions on XX. Therefore, (1.1) implies that for μ\mu-almost every xXx\in X,

1T0TδRtxdtTμ, (or 1Nn=1NδRnxNμ),\displaystyle\frac{1}{T}\int_{0}^{T}\delta_{R^{t}x}dt\xrightarrow{T\to\infty}\mu,\text{ (or }\frac{1}{N}\sum_{n=1}^{N}\delta_{R^{n}x}\xrightarrow{N\to\infty}\mu), (1.2)

in the weak* topology on the set of all probability measures on XX. Here δ\delta is the Dirac measure on XX.

For xXx\in X, we say that xx is Birkhoff generic with respect to (μ,R)(\mu,R) if xx satisfies (1.2). Given a Radon measure ν\nu on XX (possibly singular to μ\mu), if ν\nu-almost every xXx\in X is Birkhoff generic with respect to (μ,R)(\mu,R), we say that ν\nu is Birkhoff generic with respect to (μ,R)(\mu,R). It is then natural to ask the following

Question 1.1.

Under what conditions the measure ν\nu is Birkhoff generic with respect to (μ,R)(\mu,R)?

This question had been previously studied in the case of X=/X=\mathbb{R}/\mathbb{Z}, μX\mu_{X} is the Lebesgue measure on XX and Rn=×n mod R^{n}=\times n\text{ mod }\mathbb{Z} in [10]. It was shown that for any m,nm,n\in\mathbb{N}, any RmR^{m} invariant ergodic probability measure ν\nu is Birkhoff generic with respect to (Rn,μX)(R^{n},\mu_{X}). This result was strengthened later in [9]. An analogous question was also studied in the context of moduli space of translation surfaces in [29].

We consider Question 1.1 in the setting of homogeneous dynamics. Let X=G/ΓX=G/\Gamma, where GG is a Lie group and Γ\Gamma is a lattice in GG. Here and hereafter let μ=μX\mu=\mu_{X} be the GG-invariant probability measure on XX. Let {Rt}t\{R^{t}\}_{t\in\mathbb{R}} be a one-parameter flow in XX. Assume that Rt=u(t)R^{t}=u(t), where {u(t)}t\{u(t)\}_{t\in\mathbb{R}} is a one-parameter unipotent flow, that is, the adjoint action of u(t)u(t) on the Lie algebra of GG is unipotent. In this case, Ratner’s uniform distribution theorem [20] says that for any xXx\in X, xx is Birkhoff generic with respect to (ν,u(t))(\nu,u(t)), where ν\nu is the u(t)u(t)-invariant probability measure supported on the orbit closure {u(t)x:t}¯\overline{\{u(t)x:t\in\mathbb{R}\}}. By Ratner’s orbit closure theorem [19], these orbit closures are all homogeneous. Thus this provides a satisfactory answer to Question 1.1.

On the other hand, when Rt=atR^{t}=a_{t}, here and hereafter {at}t\{a_{t}\}_{t\in\mathbb{R}} is a one-parameter diagonal flow on GG, that is, the adjoint action of ata_{t} on Lie algebra of GG is semisimple, a description of Birkhoff genericity of xXx\in X under the flow ata_{t} is much harder. Indeed, the question of describing the orbit closures of diagonal action remains open (cf. [15][27, Conjecture 1]).

Nevertheless, there is a natural class of probability measures ν\nu on XX that are interesting to study with respect to Question 1.1. This class of measures is given as follows. We define the unstable horospherical subgroup U+U^{+} with respect to ata_{t} by

U+:={gG:atgattId},\displaystyle U^{+}:=\{g\in G:a_{-t}ga_{t}\xrightarrow{t\to\infty}Id\},

where IdId is the identity element of GG. Let YU+Y\subset U^{+} be a submanifold and ν\nu be a normalized bounded supported volume measure of YY (here and hereafter by normalized measure, we mean that ν\nu is renormalized to be a probability measure). It has been proved in [21][22][23][28] that if YY satisfies certain algebraic conditions, then the translation of the measure ν\nu under ata_{t} converges weakly to μ\mu as tt\to\infty. That is, for any fCc(X)f\in C_{c}(X),

f(atx)𝑑ν(x)tf𝑑μ.\displaystyle\int f(a_{t}x)d\nu(x)\xrightarrow{t\to\infty}\int fd\mu.

As in Question 1.1, it is curious to ask the following

Question 1.2.

Assume that a bounded supported normalized volume measure ν\nu of a submanifold YU+Y\subset U^{+} is such that the translate of ν\nu under ata_{t} is equidistributed with respect to μ\mu, is it true that ν\nu is also Birkhoff generic with respect to (μ,at)(\mu,a_{t})?

Roughly speaking, Question 1.2 is answered when the manifold YY is considerably "large" compared to the unstable horosphere subgroup of ata_{t}. In [25], Shi considered the situation where GG is a semisimple Lie group, Y=UY=U is the ata_{t} expanding subgroup (cf. [24]) of U+U^{+} and ν\nu is a normalized bounded supported Haar measure on UU. By [21], ν\nu satisfies the assumption of Question 1.2. Shi showed that ν\nu is also Birkhoff generic with respect to (μ,at\mu,a_{t}), and thus gave an affirmative answer to Question 1.2. In the special case where G=SLd()G=SL_{d}(\mathbb{R}) and Γ=SLd()\Gamma=SL_{d}(\mathbb{Z}), authors in [13] also obtained the effective convergence rate of (1.2).

In [8], the authors considered the setting where G=SL2()2G=SL_{2}(\mathbb{R})\ltimes\mathbb{R}^{2}, Γ=SL2()2\Gamma=SL_{2}(\mathbb{Z})\ltimes\mathbb{Z}^{2} and X=G/ΓX=G/\Gamma. One of the main results in [8] asserts that if YY is a C1C^{1} curve in U+U^{+} that intersects any affine rational line in a Lebesgue null set, then a normalized bounded supported volume measure ν\nu on YY is Birkhoff generic with respect to (μ,at)(\mu,a_{t}). By the equidistribution result of translation of such ν\nu under ata_{t} in [3], the result in [8] also gives an affirmative answer to Question 1.2 in the case where YY is a curve.

Let X=SL3()/SL3()X=SL_{3}(\mathbb{R})/SL_{3}(\mathbb{Z}). In a recent preprint [12, Theorem 1.4], it is shown that when the natural measure ν\nu on the planar line LU+L\subset U^{+} gets equidistributed under ata_{t}, then for ν\nu almost every point xx in LL, the orbit {atx}t0\{a_{t}x\}_{t\geq 0} is dense in XX. This supports an affirmative answer to Question 1.2.

The aim of this paper is to generalize the genericity results in [8] to X=SLd()(d)k/SLd()(d)kX=SL_{d}(\mathbb{R})\ltimes(\mathbb{R}^{d})^{k}/SL_{d}(\mathbb{Z})\ltimes(\mathbb{Z}^{d})^{k} and gives an affirmative answer to Question 1.2 when the manifold YY satisfies certain diophantine condition. We also deduce an application of our results to the statistics of universal hitting time for integrable flows.

1.1. Notations

From now on, any vector in the Euclidean space will be taken to be a column vector, and we will use boldface letters to denote vectors and matrices. Also, a.e. will be the shorthand for Lebesgue almost everywhere. |||\cdot| will denote Lebesgue measure of measurable subsets of Euclidean space or absolute value of real numbers. \left\lVert\cdot\right\rVert will denote the standard Euclidean norm and \left\lVert\cdot\right\rVert_{\infty} the sup norm of a vector or matrix. Throughout this article, for two matrices AA and BB, ABA\cdot B will denote matrix multiplication.

For m,nm,n\in\mathbb{N}, Matm×n()Mat_{m\times n}(\mathbb{R}) will denote the space of mm by nn real matrices. (n)m(\mathbb{R}^{n})^{m} is the direct product of m copies of n\mathbb{R}^{n}.

Fix integers d2d\geq 2, k1k\geq 1. Let G=SLd()G^{\prime}=SL_{d}(\mathbb{R}), G=SLd()(d)kG=SL_{d}(\mathbb{R})\ltimes(\mathbb{R}^{d})^{k}, Γ=SLd()\Gamma^{\prime}=SL_{d}(\mathbb{Z}) and Γ=SLd()(d)k\Gamma=SL_{d}(\mathbb{Z})\ltimes(\mathbb{Z}^{d})^{k}. It is well known that Γ\Gamma^{\prime} is a lattice in GG^{\prime} and Γ\Gamma is a lattice in GG.

Let X=G/ΓX=G/\Gamma. Denote μX\mu_{X} the GG-invariant probability measure on G/ΓG/\Gamma. Note that the action of GG^{\prime} on (d)k(\mathbb{R}^{d})^{k} is given by

g𝐯=(g𝐯𝟏,,g𝐯𝐤),\displaystyle g\cdot\mathbf{v}=(g\cdot\mathbf{v_{1}},\cdots,g\cdot\mathbf{v_{k}}),

where gGg\in G^{\prime} and 𝐯=(𝐯𝟏,,𝐯k)\mathbf{v}=(\mathbf{v_{1}},\cdots,\mathbf{v}_{k}) with 𝐯𝐢d\mathbf{v_{i}}\in\mathbb{R}^{d}. The multiplication law in GG is given by

(g,𝐯)(g,𝐯)=(gg,𝐯+g𝐯).\displaystyle(g,\mathbf{v})\cdot(g^{\prime},\mathbf{v^{\prime}})=(g\cdot g^{\prime},\mathbf{v}+g\cdot\mathbf{v^{\prime}}).

GG^{\prime} is naturally embedded into GG by

G(G,𝟎)G.\displaystyle G^{\prime}\cong(G^{\prime},\mathbf{0})\leq G.

Fix an r{1,,d1}r\in\{1,\cdots,d-1\}. For tt\in\mathbb{R} and 𝐬Matr×(dr)()\mathbf{s}\in Mat_{r\times(d-r)}(\mathbb{R}), denote

at=diag[e(dr)t,,e(dr)t,ert,,ert],\displaystyle a_{t}=diag[e^{(d-r)t},\cdots,e^{(d-r)t},e^{-rt},\cdots,e^{-rt}], (1.3)
u(𝐬)=[𝟏r𝐬𝟎dr,r𝟏dr],\displaystyle u(\mathbf{s})=\begin{bmatrix}\mathbf{1}_{r}&\mathbf{s}\\ \mathbf{0}_{d-r,r}&\mathbf{1}_{d-r}\end{bmatrix}, (1.4)
U:={u(𝐬):𝐬Matr×(dr)()}Matr×(dr)().\displaystyle U:=\{u(\mathbf{s}):\mathbf{s}\in Mat_{r\times(d-r)}(\mathbb{R})\}\cong Mat_{r\times(d-r)}(\mathbb{R}). (1.5)

For a column vector or a matrix 𝐯\mathbf{v}, let (𝐯)r(\mathbf{v})_{\leq r} (or (𝐯)>r)(\mathbf{v})_{>r}) be the first rr rows (or last drd-r rows) of 𝐯\mathbf{v}. For example, if 𝐯(d)k\mathbf{v}\in(\mathbb{R}^{d})^{k}, then

(𝐯)r(r)k, (𝐯)>r(dr)k.\displaystyle(\mathbf{v})_{\leq r}\in(\mathbb{R}^{r})^{k},\textbf{ }(\mathbf{v})_{>r}\in(\mathbb{R}^{d-r})^{k}.

With the above notations, the unstable horospherical subgroup U+U^{+} of ata_{t} in GG is

U+=U{(Id,[(𝐯)r𝟎]):𝐯(d)k}.\displaystyle U^{+}=U\cdot\{(Id,\begin{bmatrix}(\mathbf{v})_{\leq r}\\ \mathbf{0}\end{bmatrix}):\mathbf{v}\in(\mathbb{R}^{d})^{k}\}.

Lastly, for a map 𝝋:Matr×(dr)()(d)k\bm{\varphi}:Mat_{r\times(d-r)}(\mathbb{R})\to(\mathbb{R}^{d})^{k}, we write 𝝋(𝐬)=(φij(𝐬))1id,1jk\bm{\varphi}(\mathbf{s})=(\varphi_{ij}(\mathbf{s}))_{1\leq i\leq d,1\leq j\leq k}.

We also write

u𝝋(𝐬):=u(𝐬)(Id,𝝋(𝐬)).\displaystyle u_{\bm{\varphi}}(\mathbf{s}):=u(\mathbf{s})\cdot(Id,\bm{\varphi}(\mathbf{s})). (1.6)

1.2. Main results

For any 𝐬Matr×(dr)()\mathbf{s}\in Mat_{r\times(d-r)}(\mathbb{R}) and T>0T>0, define the probability measure

μ𝐬,T=1T0Tδatu𝝋(𝐬)Γ𝑑t.\displaystyle\mu_{\mathbf{s},T}=\frac{1}{T}\int_{0}^{T}\delta_{a_{t}u_{\bm{\varphi}}(\mathbf{s})\Gamma}dt. (1.7)

As before, we say that u𝝋(𝐬)Γu_{\bm{\varphi}}(\mathbf{s})\Gamma is Birkhoff generic with respect to (X,μX,at)(X,\mu_{X},a_{t}) if μ𝐬,T\mu_{\mathbf{s},T} converges to μX\mu_{X} in the weak*-topology as TT\to\infty.

One of our main results is the following:

Theorem 1.3.

Let 𝒰Matr×(dr)()\mathcal{U}\subset Mat_{r\times(d-r)}(\mathbb{R}) be a bounded open subset. Let 𝛗:𝒰(d)k\bm{\varphi}:\mathcal{U}\to(\mathbb{R}^{d})^{k} be a C1C^{1}-map satisfying (𝛗(𝐬))>r𝟎(\bm{\varphi}(\mathbf{s}))_{>r}\equiv\mathbf{0} for any 𝐬𝒰\mathbf{s}\in\mathcal{U}. Assume that for any 𝐦k{𝟎}\mathbf{m}\in\mathbb{Z}^{k}\setminus\{\mathbf{0}\},

|{𝐬𝒰:(𝝋(𝐬))r𝐦𝐬dr+r}|=0,\displaystyle|\{\mathbf{s}\in\mathcal{U}:(\bm{\varphi}(\mathbf{s}))_{\leq r}\cdot\mathbf{m}\in\mathbf{s}\cdot\mathbb{Z}^{d-r}+\mathbb{Z}^{r}\}|=0, (1.8)

then for Lebesgue a.e. 𝐬𝒰\mathbf{s}\in\mathcal{U}, u𝛗(𝐬)Γu_{\bm{\varphi}}(\mathbf{s})\Gamma is Birkhoff generic with respect to (X,μX,at)(X,\mu_{X},a_{t}).

Using the observation that if a trajectory equidistributes with respect to μX\mu_{X}, then all other parallel trajectories will also equidistribute with respect to μX\mu_{X} (see Lemma 2.3), we can remove the assumption that (𝝋(𝐬))>r𝟎(\bm{\varphi}(\mathbf{s}))_{>r}\equiv\mathbf{0} and strengthen Theorem 1.3 to the following

Corollary 1.4.

Let 𝒰\mathcal{U} be a bounded open subset of Matr×(dr)()Mat_{r\times(d-r)}(\mathbb{R}). Let 𝛗:𝒰(d)k\bm{\varphi}:\mathcal{U}\to{(\mathbb{R}^{d})^{k}} be a C1C^{1} map. If for any 𝐦k{𝟎}\mathbf{m}\in\mathbb{Z}^{k}\setminus\{\mathbf{0}\},

|{𝐬𝒰:(𝝋(𝐬))r𝐦𝐬dr+r}|=0,\displaystyle|\{\mathbf{s}\in\mathcal{U}:(\bm{\varphi}(\mathbf{s}))_{\leq r}\cdot\mathbf{m}\in\mathbf{s}\cdot\mathbb{Z}^{d-r}+\mathbb{Z}^{r}\}|=0,

then for Lebesgue a.e. 𝐬𝒰\mathbf{s}\in\mathcal{U}, u𝛗(𝐬)Γu_{\bm{\varphi}}(\mathbf{s})\Gamma is Birkhoff generic with respect to (X,μX,at)(X,\mu_{X},a_{t}).

We also obtain the following variants of Theorem 1.3.

Theorem 1.5.

Let 𝒰\mathcal{U} be a bounded open subset of Matr×(dr)()Mat_{r\times(d-r)}(\mathbb{R}). Let 𝛗:𝒰(d)k\bm{\varphi}:\mathcal{U}\to(\mathbb{R}^{d})^{k} be a C1C^{1} map. Let 𝐌SLd()\mathbf{M}\in SL_{d}(\mathbb{R}). If for any 𝐦k{𝟎}\mathbf{m}\in\mathbb{Z}^{k}\setminus\{\mathbf{0}\},

|{𝐬𝒰:𝝋(𝐬)𝐦𝐌1u(𝐬)[𝟎dr]+d}|=0.\displaystyle|\{\mathbf{s}\in\mathcal{U}:\bm{\varphi}(\mathbf{s})\cdot\mathbf{m}\in\mathbf{M}^{-1}u(-\mathbf{s})\cdot\begin{bmatrix}\mathbf{0}\\ \mathbb{R}^{d-r}\end{bmatrix}+{\mathbb{Z}^{d}}\}|=0. (1.9)

Then for Lebesgue a.e. 𝐬𝒰\mathbf{s}\in\mathcal{U}, u(𝐬)𝐌(Id,𝛗(𝐬))Γu(\mathbf{s})\mathbf{M}(Id,\bm{\varphi}(\mathbf{s}))\Gamma is Birkhoff generic with respect to (X,μX,at)(X,\mu_{X},a_{t}).

Remark 1.6.

By the equidistribution result in [3], Theorem 1.5 gives an affirmative answer to Question 1.2.

Remark 1.7.

The conditions (1.8) and (1.9) are indeed necessary. For example in Theorem 1.3, suppose that (𝝋(𝐬))r𝐦𝐬dr+r(\bm{\varphi}(\mathbf{s}))_{\leq r}\cdot\mathbf{m}\in\mathbf{s}\cdot\mathbb{Z}^{d-r}+\mathbb{Z}^{r} for some 𝐬𝒰\mathbf{s}\in\mathcal{U} and 𝐦k{𝟎}\mathbf{m}\in\mathbb{Z}^{k}\setminus\{\mathbf{0}\}, then the ata_{t} trajectory along u𝝋(𝐬)Γu_{\bm{\varphi}}(\mathbf{s})\Gamma will concentrate on a proper submanifold of G/ΓG/\Gamma.

Corollary 1.8.

Let 𝒰\mathcal{U} be a bounded open subset of Matr×(dr)()Mat_{r\times(d-r)}(\mathbb{R}). Let 𝛗:𝒰(d)k\bm{\varphi}:\mathcal{U}\to(\mathbb{R}^{d})^{k} be a C1C^{1} map. Let (𝐌,𝐯)G(\mathbf{M},\mathbf{v})\in G. If for any 𝐦k{𝟎}\mathbf{m}\in\mathbb{Z}^{k}\setminus\{\mathbf{0}\},

|{𝐬𝒰:(𝝋(𝐬)+𝐯)𝐦u(𝐬)[𝟎dr]+𝐌d}|=0,\displaystyle|\{\mathbf{s}\in\mathcal{U}:(\bm{\varphi}(\mathbf{s})+\mathbf{v})\cdot\mathbf{m}\in u(-\mathbf{s})\cdot\begin{bmatrix}\mathbf{0}\\ \mathbb{R}^{d-r}\end{bmatrix}+\mathbf{M}\cdot\mathbb{Z}^{d}\}|=0,

then for Lebesgue a.e. 𝐬𝒰\mathbf{s}\in\mathcal{U}, u𝛗(𝐬)(𝐌,𝐯)Γu_{\bm{\varphi}}(\mathbf{s})(\mathbf{M},\mathbf{v})\Gamma is Birkhoff generic with respect to (X,μX,at)(X,\mu_{X},a_{t}).

For 1id1\leq i\leq d, let 𝐞id\mathbf{e}_{i}\in\mathbb{R}^{d} be the column vector such that ii-th row of 𝐞i\mathbf{e}_{i} is 11 and others are 0.

Corollary 1.9.

Let 𝒰\mathcal{U} be a bounded open subset of Matr×(dr)()Mat_{r\times(d-r)}(\mathbb{R}). Let 𝐄𝟏:𝒰SOd()\mathbf{E_{1}}:\mathcal{U}\to SO_{d}(\mathbb{R}) be a smooth map such that the map 𝐬𝐄𝟏(𝐬)1[𝐞r+1,,𝐞d]\mathbf{s}\mapsto\mathbf{E_{1}}(\mathbf{s})^{-1}\cdot[\mathbf{e}_{r+1},\cdots,\mathbf{e}_{d}] has a nonsingular differential at Lebesgue almost every 𝐬𝒰\mathbf{s}\in\mathcal{U}. Let 𝛗:𝒰(d)k\bm{\varphi}:\mathcal{U}\to(\mathbb{R}^{d})^{k} be a C1C^{1} map. Assume that for any 𝐦k{𝟎}\mathbf{m}\in\mathbb{Z}^{k}\setminus\{\mathbf{0}\},

|{𝐬𝒰:𝝋(𝐬)𝐦𝐄1(𝐬)1[𝟎dr]+d}|=0,\displaystyle|\{\mathbf{s}\in\mathcal{U}:\bm{\varphi}(\mathbf{s})\cdot\mathbf{m}\in\mathbf{E}_{1}(\mathbf{s})^{-1}\cdot\begin{bmatrix}\mathbf{0}\\ \mathbb{R}^{d-r}\end{bmatrix}+{\mathbb{Z}^{d}}\}|=0,

then for Lebesgue a.e. 𝐬𝒰\mathbf{s}\in\mathcal{U}, 𝐄1(𝐬)(Id,𝛗(𝐬))Γ\mathbf{E}_{1}(\mathbf{s})(Id,\bm{\varphi}(\mathbf{s}))\Gamma is Birkhoff generic with respect to (X,μX,at)(X,\mu_{X},a_{t}).

Corollary 1.9 will allow us to deduce an application to universal hitting time for integrable flows in dd-torus 𝕋d\mathbb{T}^{d} (see Theorem 9.5).

1.3. Ingredients of the proof

Proof of Theorem 1.3 follows the similar strategy as in [8]. However, some new ingredients are required. We need the description of orbit closures of GG^{\prime} in XX. This can be done using Ratner’s orbit closure theorem following the approach in [3].

We need to construct a suitable mixed height function in our situation, which measures the distance of point to the cusp and singular submanifolds.

Also due to higher rank, some technical difficulties arise in the proof of uniform contraction property of the mixed height function. To overcome these difficulties, we apply a linear algebra lemma (see Lemma 6.13) which is inspired by the proof of [11, Proposition 3.4].

1.4. Overview

In Section 2, we make some reductions and give a proof of Theorem 1.3 and Corollary 1.4.

In Section 3, we will investigate the orbit closure of GG^{\prime} in XX using Ratner’s orbit closure theorem.

In Section 4, we prove that for a.e. 𝐬𝒰\mathbf{s}\in\mathcal{U}, the limit measure is invariant under the unipotent group UU. This enables us to apply Ratner’s measure classification theorem.

In Section 5 and Section 6, we will construct mixed height function β𝐦\beta_{\mathbf{m}} for 𝐦k{𝟎}\mathbf{m}\in\mathbb{Z}^{k}\setminus\{\mathbf{0}\}, and give a proof of its uniform contraction property.

In Section 7, we prove Proposition 2.1 using mixed height function.

In Section 8, we deduce variants of Theorem 1.3.

In Section 9, we deduce an application to universal hitting time statistics.

2. Reductions and proof of Theorem 1.3

In this section, assuming several Theorems/Propositions/Lemmas that will be proved later, we give a proof of Theorem 1.3.

By Proposition 4.1, for a.e. 𝐬𝒰\mathbf{s}\in\mathcal{U}, after possibly passing to a subsequence the weak* limit μ𝐬\mu_{\mathbf{s}} of μ𝐬,T\mu_{\mathbf{s},T} is UU-invariant. From the definition of μ𝐬,T\mu_{\mathbf{s},T} (see (1.7)), it follows that μ𝐬\mu_{\mathbf{s}} is also D={at:t}D=\{a_{t}:t\in\mathbb{R}\}-invariant. Hence for a.e. 𝐬𝒰\mathbf{s}\in\mathcal{U}, μ𝐬\mu_{\mathbf{s}} is DUDU-invariant. Note that DUDU is an epimorphic subgroup of G=SLd()G^{\prime}=SL_{d}(\mathbb{R}). By [17], as μ𝐬\mu_{\mathbf{s}} is a probability measure invariant under DUDU, μ𝐬\mu_{\mathbf{s}} is GG^{\prime}-invariant. By Ratner’s measure classification theorem, any GG^{\prime} invariant and ergodic probability measure is supported on an orbit closure of GG^{\prime} on XX.

A consequence of Ratner’s orbit closure theorem (Theorem 3.1) shows that any orbit closure of GG^{\prime} is either

(1) the whole XX, or

(2) concentrated in a proper closed submanifold X𝐦X_{\mathbf{m}} for some 𝐦k{𝟎}\mathbf{m}\in\mathbb{Z}^{k}\setminus\{\mathbf{0}\}, where

X𝐦={(g,g𝐯)Γ:gG,𝐯𝐦d}.\displaystyle X_{\mathbf{m}}=\{(g,g\mathbf{v})\Gamma:g\in G^{\prime},\mathbf{v}\cdot\mathbf{m}\in\mathbb{Z}^{d}\}.

Therefore, it remains to show that for a.e. 𝐬𝒰\mathbf{s}\in\mathcal{U}, μ𝐬\mu_{\mathbf{s}} is a probability measure on XX and μ𝐬(X𝐦)=0\mu_{\mathbf{s}}(X_{\mathbf{m}})=0 for any 𝐦k{𝟎}\mathbf{m}\in\mathbb{Z}^{k}\setminus\{\mathbf{0}\}.

Let

M1:=N1(max1ir,1jdrsups𝒰ij𝝋(𝐬))+1,\displaystyle M_{1}:=N_{1}\cdot(\max_{1\leq i\leq r,1\leq j\leq d-r}\sup_{s\in\mathcal{U}}\left\lVert\partial_{ij}\bm{\varphi}(\mathbf{s})\right\rVert_{\infty})+1, (2.1)

where N1=8r2k1/2(dr)N_{1}=8r^{2}k^{1/2}(d-r), and ij𝝋\partial_{ij}\bm{\varphi} is a dd by kk matrix whose (p,q)(p,q)-th entry is φpq/sij\partial\varphi_{pq}/\partial s_{ij}. Here the choice of N1N_{1} is flexible, we just choose a value for N1N_{1} that is convenient for us.

By assumption (1.8) of Theorem 1.3, for any 𝐦k{𝟎}\mathbf{m}\in\mathbb{Z}^{k}\setminus\{\mathbf{0}\}, the set

Bad𝐦={𝐬𝒰:𝐚dr,𝐛r such that (𝝋(𝐬))r𝐦=𝐬𝐚+𝐛,\displaystyle Bad_{\mathbf{m}}=\{\mathbf{s}\in\mathcal{U}:\exists\mathbf{a}\in\mathbb{Z}^{d-r},\mathbf{b}\in\mathbb{Z}^{r}\text{ such that }(\bm{\varphi}(\mathbf{s}))_{\leq r}\bm{\cdot}\mathbf{m}=\mathbf{s}\bm{\cdot}\mathbf{a}+\mathbf{b},
and 𝐚M1𝐦}\displaystyle\text{ and }\left\lVert\mathbf{a}\right\rVert_{\infty}\leq M_{1}\left\lVert\mathbf{m}\right\rVert\} (2.2)

has Lebesgue measure zero.

Since there are only finitely many 𝐚dr\mathbf{a}\in\mathbb{Z}^{d-r} such that 𝐚M1𝐦\left\lVert\mathbf{a}\right\rVert_{\infty}\leq M_{1}\cdot\left\lVert\mathbf{m}\right\rVert, Bad𝐦Bad_{\mathbf{m}} is a closed set with Lebesgue measure zero. Thus to prove Theorem 1.3, it suffices to prove it for a closed cube contained in 𝒰Bad𝐦\mathcal{U}\setminus Bad_{\mathbf{m}}. Now let’s fix a closed cube I𝒰Bad𝒎I\subset\mathcal{U}\setminus Bad_{\bm{m}}.

Let KK be a measurable subset of XX. For any T>0T>0, we define the average operator 𝒜KT:𝒰[0,1]\mathcal{A}_{K}^{T}:\mathcal{U}\to[0,1] by

𝒜KT(𝐬)=1T0TχK(atu𝝋(𝐬)Γ)𝑑t=μ𝐬,T(K),\displaystyle\mathcal{A}_{K}^{T}(\mathbf{s})=\frac{1}{T}\int_{0}^{T}\chi_{K}(a_{t}u_{\bm{\varphi}}(\mathbf{s})\Gamma)dt=\mu_{\mathbf{s},T}(K),

where χK\chi_{K} is the characteristic function of KK.

The key proposition, which ensures that μ𝐬\mu_{\mathbf{s}} is a probability measure putting zero mass on X𝐦X_{\mathbf{m}} for a.e. 𝐬I\mathbf{s}\in I, is the following:

Proposition 2.1.

Let 𝐦k{𝟎}\mathbf{m}\in\mathbb{Z}^{k}\setminus\{\mathbf{0}\}. Let 𝛗:I(d)k\bm{\varphi}:I\to(\mathbb{R}^{d})^{k} be a C1C^{1} map satisfying (𝛗(𝐬))>r𝟎(\bm{\varphi}(\mathbf{s}))_{>r}\equiv\mathbf{0} for any 𝐬I\mathbf{s}\in I. Suppose that

inf𝐬I{(𝝋(𝐬))r𝐦𝐬𝐚𝐛:𝐚M1𝐦,𝐚dr,𝐛r}>0.\displaystyle\inf_{\mathbf{s}\in I}\{\left\lVert(\bm{\varphi}(\mathbf{s}))_{\leq r}\bm{\cdot}\mathbf{m}-\mathbf{s}\bm{\cdot}\mathbf{a}-\mathbf{b}\right\rVert_{\infty}:\left\lVert\mathbf{a}\right\rVert_{\infty}\leq M_{1}\left\lVert\mathbf{m}\right\rVert,\mathbf{a}\in\mathbb{Z}^{d-r},\mathbf{b}\in\mathbb{Z}^{r}\}>0. (2.3)

Then for any ϵ>0\epsilon>0, there exists a compact subset KϵXX𝐦{K_{\epsilon}}\subset X\setminus X_{\mathbf{m}} and 𝔳>0\mathfrak{v}>0 such that for any T>0T>0,

|{𝐬I:𝒜KϵT(𝐬)1ϵ}|e𝔳T|I|.\displaystyle|\{\mathbf{s}\in I:{\mathcal{A}_{K_{\epsilon}}^{T}}(\mathbf{s})\leq 1-\epsilon\}|\leq e^{-\mathfrak{v}T}|I|. (2.4)

It will be proved in Lemma 6.1 that condition (2.3) in Proposition 2.1 follows from condition (1.8) in Theorem 1.3.

Proposition 2.1 will be proved in Section 6. Combining Borel-Cantelli lemma, a direct consequence of Proposition 2.1 is the following:

Proposition 2.2.

Under the assumptions of Theorem 1.3, for a.e. 𝐬𝒰\mathbf{s}\in\mathcal{U}, by possibly passing to a subsequence, μ𝐬,T\mu_{\mathbf{s},T} converges to a probability measure μ𝐬\mu_{\mathbf{s}} on XX in weak*-topology as TT\to\infty, and μ𝐬(X𝐦)=0\mu_{\mathbf{s}}(X_{\mathbf{m}})=0 for any 𝐦k{𝟎}\mathbf{m}\in\mathbb{Z}^{k}\setminus\{\mathbf{0}\}.

Proof.

Fix an 𝐦k{𝟎}\mathbf{m}\in\mathbb{Z}^{k}\setminus\{\mathbf{0}\} and ϵ>0\epsilon>0. By Proposition 2.1, we can choose a compact subset KϵK_{\epsilon} of XX𝐦X\setminus X_{\mathbf{m}} such that (2.4) holds for any T>0T>0. Let T=nT=n\in\mathbb{N} and apply Borel-Cantelli lemma to the collection of the sets

{𝐬I:𝒜Kϵn(𝐬)1ϵ},n.\displaystyle\{\mathbf{s}\in I:\mathcal{A}_{K_{\epsilon}}^{n}(\mathbf{s})\leq 1-\epsilon\},n\in\mathbb{N}.

We can find a measurable subset I𝐦ϵI_{\mathbf{m}}^{\epsilon} of II with full measure such that for any 𝐬I𝐦ϵ\mathbf{s}\in I_{\mathbf{m}}^{\epsilon}, 𝒜Kϵn(𝐬)>1ϵ\mathcal{A}_{K_{\epsilon}}^{n}(\mathbf{s})>1-\epsilon for all sufficiently large nn\in\mathbb{N}. Therefore, for any 𝐬I𝐦ϵ\mathbf{s}\in I_{\mathbf{m}}^{\epsilon}, μ𝐬(X)1ϵ\mu_{\mathbf{s}}(X)\geq 1-\epsilon and μ𝐬(X𝐦)ϵ\mu_{\mathbf{s}}(X_{\mathbf{m}})\leq\epsilon. Let I𝐦=n=1I𝐦1nI_{\mathbf{m}}=\cap_{n=1}^{\infty}I_{\mathbf{m}}^{\frac{1}{n}}, then I𝐦I_{\mathbf{m}} has full Lebesgue measure in II, and for any 𝐬I𝐦\mathbf{s}\in I_{\mathbf{m}}, μ𝐬(X)=1\mu_{\mathbf{s}}(X)=1 and μ𝐬(X𝐦)=0\mu_{\mathbf{s}}(X_{\mathbf{m}})=0.

To complete the proof, we let I=𝐦k{𝟎}I𝐦I^{\prime}=\bigcap_{\mathbf{m}\in\mathbb{Z}^{k}\setminus\{\mathbf{0}\}}I_{\mathbf{m}}. Then II^{\prime} has full Lebesgue measure in II and the proposition holds for all 𝐬I\mathbf{s}\in I^{\prime}. ∎

Assuming Proposition 2.2, Proposition 4.1 and Theorem 3.1, we are ready to prove Theorem 1.3:

Proof of Theorem 1.3.

By Proposition 2.2 and Proposition 4.1, we conclude that for a.e. 𝐬𝒰\mathbf{s}\in\mathcal{U}, the weak* limit μ𝐬\mu_{\mathbf{s}} of μ𝐬,T\mu_{\mathbf{s},T} as TT\to\infty is

(1) a probability measure on XX, and μ𝐬(X𝐦)=0\mu_{\mathbf{s}}(X_{\mathbf{m}})=0 for any 𝐦k{𝟎}\mathbf{m}\in\mathbb{Z}^{k}\setminus\{\mathbf{0}\};

(2) DUDU-invariant.

Since DUDU is an epimorphic subgroup of G=SLd()G^{\prime}=SL_{d}(\mathbb{R}), and μ𝐬\mu_{\mathbf{s}} is a DUDU-invariant probability measure on XX, μ𝐬\mu_{\mathbf{s}} is GG^{\prime}-invariant by [17, Theorem 1].

By Ratner’s measure classification theorem [19], any ergodic component of such μ𝐬\mu_{\mathbf{s}} is supported on an orbit closure of GG^{\prime} on XX. Theorem 3.1 describes all the possible orbit closures of GG^{\prime} on XX: either it is XX or it is concentrated on X𝐦X_{\mathbf{m}} for some 𝐦k{𝟎}\mathbf{m}\in\mathbb{Z}^{k}\setminus\{\mathbf{0}\}.

Since μ𝐬(X𝐦)=0\mu_{\mathbf{s}}(X_{\mathbf{m}})=0 for any 𝐦k{𝟎}\mathbf{m}\in\mathbb{Z}^{k}\setminus\{\mathbf{0}\}, we conclude that for a.e. 𝐬𝒰\mathbf{s}\in\mathcal{U}, μ𝐬=μX\mu_{\mathbf{s}}=\mu_{X}. ∎

We note the following

Lemma 2.3.

Assume that for some xX=G/Γx\in X=G/\Gamma,

1T0Tδatx𝑑tTμG/Γ, in weak*-topology,\displaystyle\frac{1}{T}\int_{0}^{T}\delta_{a_{t}x}dt\xrightarrow{T\to\infty}\mu_{G/\Gamma},\text{ in weak*-topology},

and for some gGg\in G, atgatIdGa_{t}ga_{-t}\to Id\in G as tt\to\infty, then

1T0Tδatgx𝑑tTμG/Γ, in weak*-topology.\displaystyle\frac{1}{T}\int_{0}^{T}\delta_{a_{t}gx}dt\xrightarrow{T\to\infty}\mu_{G/\Gamma},\text{ in weak*-topology}.

For any 𝝋:Matr×(dr)()(d)k\bm{\varphi}:Mat_{r\times(d-r)}(\mathbb{R})\to(\mathbb{R}^{d})^{k}, and any 𝐬Matr×(dr)()\mathbf{s}\in Mat_{r\times(d-r)}(\mathbb{R}), we can write

atu𝝋(𝐬)\displaystyle a_{t}u_{\bm{\varphi}}(\mathbf{s}) =at(u(𝐬),𝝋(𝐬))\displaystyle=a_{t}(u(\mathbf{s}),\bm{\varphi}(\mathbf{s}))
=at(Id,[𝟎(𝝋(𝐬))>r])atat(u(𝐬),[(𝝋(𝐬))r𝟎]),\displaystyle=a_{t}(Id,\begin{bmatrix}\mathbf{0}\\ (\bm{\varphi}(\mathbf{s}))_{>r}\end{bmatrix})a_{-t}\cdot a_{t}(u(\mathbf{s}),\begin{bmatrix}(\bm{\varphi}(\mathbf{s}))_{\leq r}\\ \mathbf{0}\end{bmatrix}),

where

𝝋(𝐬)=[𝟎(𝝋(𝐬))>r]+[(𝝋(𝐬))r𝟎].\displaystyle\bm{\varphi}(\mathbf{s})=\begin{bmatrix}\mathbf{0}\\ (\bm{\varphi}(\mathbf{s}))_{>r}\end{bmatrix}+\begin{bmatrix}(\bm{\varphi}(\mathbf{s}))_{\leq r}\\ \mathbf{0}\end{bmatrix}.

Since

at(Id,[𝟎(𝝋(𝐬))>r])at(Id,𝟎),\displaystyle a_{t}(Id,\begin{bmatrix}\mathbf{0}\\ (\bm{\varphi}(\mathbf{s}))_{>r}\end{bmatrix})a_{-t}\to(Id,\mathbf{0}),

by Lemma 2.3 and Theorem 1.3, Corollary 1.4 is proven.

3. Orbit closure

In this section, we will classify all orbit closures of GG^{\prime} in XX following [3]. Recall that G=SLd()G^{\prime}=SL_{d}(\mathbb{R}) and G=SLd()(d)kG=SL_{d}(\mathbb{R})\ltimes(\mathbb{R}^{d})^{k}.

Consider a base point (Id,𝝃)G(Id,\bm{\xi})\in G. Since GG^{\prime} is a simple Lie group, an application of Ratner’s orbit closure theorem gives the following theorem describing the orbit closure of G(Id,𝝃)Γ/ΓG^{\prime}\cdot(Id,\bm{\xi})\Gamma/\Gamma in G/ΓG/\Gamma:

Theorem 3.1.

The orbit closure G(Id,𝛏)Γ/Γ¯\overline{G^{\prime}\cdot(Id,\bm{\xi})\Gamma/\Gamma} is G/ΓG/\Gamma if and only if for any 𝐦k{𝟎}\mathbf{m}\in\mathbb{Z}^{k}\setminus\{\mathbf{0}\}, 𝛏𝐦d\bm{\xi}\cdot\mathbf{m}\notin\mathbb{Z}^{d}.

By Ratner’s orbit closure theorem ([20]), for any 𝝃(d)k\bm{\xi}\in(\mathbb{R}^{d})^{k}, there exists a closed subgroup HH of GG containing GG^{\prime} such that

G(Id,𝝃)Γ/Γ¯=H(Id,𝝃)Γ/Γ,\displaystyle\overline{G^{\prime}\cdot(Id,\bm{\xi})\Gamma/\Gamma}=H\cdot(Id,\bm{\xi})\Gamma/\Gamma,

and H(Id,𝝃)Γ/ΓH\cdot(Id,\bm{\xi})\Gamma/\Gamma admits an HH-invariant probability measure.

It can be checked that if there exists 𝐦k{𝟎}\mathbf{m}\in\mathbb{Z}^{k}\setminus\{\mathbf{0}\} such that 𝝃𝐦d\bm{\xi}\cdot\mathbf{m}\in\mathbb{Z}^{d}, then G(Id,𝝃)ΓX𝐦G^{\prime}\cdot(Id,\bm{\xi})\Gamma\subset X_{\mathbf{m}}, where

X𝐦={(g,g𝐯)Γ:gG,𝐯(d)k such that 𝐯𝐦d}.\displaystyle X_{\mathbf{m}}=\{(g,g\mathbf{v})\Gamma:g\in G^{\prime},\mathbf{v}\in(\mathbb{R}^{d})^{k}\text{ such that }\mathbf{v}\cdot\mathbf{m}\in\mathbb{Z}^{d}\}. (3.1)

This is a closed submanifold of XX of codimension dd. In this case, the orbit G(Id,𝝃)Γ/ΓG^{\prime}\cdot(Id,\bm{\xi})\Gamma/\Gamma does not equidistribute in XX.

The converse of Theorem 3.1 will follow from Lemmas 3.2-3.4. We will follow the proof strategy of [3, Theorem 3]. Let 𝝃\bm{\xi}, HH be as above.

Lemma 3.2.

There is a linear subspace 𝐔k\bm{U}\subset\mathbb{R}^{k} such that H=SLd()L(𝐔)H=SL_{d}(\mathbb{R})\ltimes L(\bm{U}), where L(𝐔)L(\bm{U}) is a subset of Matd×k()Mat_{d\times k}(\mathbb{R}) such that for any element 𝐯\mathbf{v} of L(𝐔)L(\bm{U}), each row vector of 𝐯\mathbf{v} is a vector in 𝐔\bm{U}.

Proof.

Let 𝑳={𝐯(d)k:(Id,𝐯)H}\bm{L}=\{\mathbf{v}\in(\mathbb{R}^{d})^{k}:(Id,\mathbf{v})\in H\}. Because GHG^{\prime}\subset H, for any 𝐯𝑳\mathbf{v}\in\bm{L}, we have (g,𝟎)(Id,𝐯)(g,𝟎)1=(Id,g𝐯)H.(g,\mathbf{0})\cdot(Id,\mathbf{v})\cdot(g,\mathbf{0})^{-1}=(Id,g\mathbf{v})\in H. It follows that 𝑳\bm{L} is GG^{\prime}-invariant and SLd()𝑳HSL_{d}(\mathbb{R})\ltimes\bm{L}\subset H. For any (g,𝐯)H(g,\mathbf{v})\in H, we have (g1,𝟎)(g,𝐯)=(Id,g1𝐯)H(g^{-1},\mathbf{0})\cdot(g,\mathbf{v})=(Id,g^{-1}\mathbf{v})\in H, so g1𝐯𝑳g^{-1}\mathbf{v}\in\bm{L}. Since 𝑳\bm{L} is GG^{\prime}-invariant, 𝐯𝑳\mathbf{v}\in\bm{L}. Therefore H=SLd()𝑳H=SL_{d}(\mathbb{R})\ltimes\bm{L}.

Let A𝔰𝔩d=Lie(G)A\in\mathfrak{sl}_{d}=Lie(G^{\prime}), then for any tt\in\mathbb{R}, and any 𝐯𝑳\mathbf{v}\in\bm{L}

exp(tA)𝐯𝐯t𝑳.\displaystyle\frac{exp(tA)\mathbf{v}-\mathbf{v}}{t}\in\bm{L}.

Let t0t\to 0, we obtain A𝐯𝑳A\cdot\mathbf{v}\in\bm{L}. Recall that 𝔰𝔩d\mathfrak{sl}_{d} consists of all trace zero d×dd\times d matrices. Let 𝐄ij\mathbf{E}_{ij} be the d×dd\times d matrix with 11 in the (i,j)(i,j)-th entry and zero for all other entries. Then for any iji\neq j, 𝐄ij𝐯𝑳\mathbf{E}_{ij}\mathbf{v}\in\bm{L}. Since 𝐄ij𝐄ji=𝐄ii\mathbf{E}_{ij}\cdot\mathbf{E}_{ji}=\mathbf{E}_{ii}, for any ii we have 𝐄ii𝐯𝑳\mathbf{E}_{ii}\mathbf{v}\in\bm{L} as well. Therefore, 𝑳\bm{L} is invariant under left multiplication of all d×dd\times d real matrices. Since left multiplication is row operation, there is a linear subspace 𝑼k\bm{U}\subset\mathbb{R}^{k} such that 𝑳=L(𝑼)\bm{L}=L(\bm{U}). ∎

Let π1:GG\pi_{1}:G\to G^{\prime} be the natural projection map and ΓL=L(𝑼)Γ.\Gamma_{L}=L(\bm{U})\cap\Gamma.

Lemma 3.3.

Let 𝐔\bm{U} be the linear subspace of k\mathbb{R}^{k} obtained by the Lemma 3.2. Then 𝐔k\bm{U}\cap\mathbb{Z}^{k} is a lattice in 𝐔\bm{U} and 𝛏(d)k+L(𝐔)\bm{\xi}\in(\mathbb{Q}^{d})^{k}+L(\bm{U}).

Proof.

By Lemma 3.2, H=SLd()L(𝑼)H=SL_{d}(\mathbb{R})\ltimes L(\bm{U}) and H(Id,𝝃)Γ/ΓH\cdot(Id,\bm{\xi})\Gamma/\Gamma is closed and admits an HH-invariant probability measure, therefore ΓH=(Id,𝝃)Γ(Id,𝝃)H\Gamma_{H}=(Id,\bm{\xi})\Gamma(Id,-\bm{\xi})\cap H is a lattice in HH.

By [18, Corollary 8.28], ΓL\Gamma_{L} is a lattice in L(𝑼)L(\bm{U}), that is, (d)kL(𝑼)(\mathbb{Z}^{d})^{k}\cap L(\bm{U}) is a lattice in L(𝑼)L(\bm{U}). Thus 𝑼\bm{U} has a basis belonging to k\mathbb{Z}^{k}, and it follows that k𝑼\mathbb{Z}^{k}\cap\bm{U} is a lattice in 𝑼\bm{U}.

Recall that Γ=SLd()\Gamma^{\prime}=SL_{d}(\mathbb{Z}). Now consider π1(ΓH)={γΓ:𝝃γ𝝃(d)k+L(𝑼)}\pi_{1}(\Gamma_{H})=\{\gamma\in\Gamma^{\prime}:\bm{\xi}-\gamma\cdot\bm{\xi}\in(\mathbb{Z}^{d})^{k}+L(\bm{U})\}. Again by [18, Corollary 8.28], π1(ΓH)\pi_{1}(\Gamma_{H}) is a lattice in GG^{\prime}. Therefore π1(ΓH)\pi_{1}(\Gamma_{H}) is a finite index subgroup of Γ\Gamma^{\prime}. Pick a γπ1(ΓH)\gamma\in\pi_{1}(\Gamma_{H}) such that IdγId-\gamma is invertible, then 𝝃(d)k+L(𝑼)\bm{\xi}\in(\mathbb{Q}^{d})^{k}+L(\bm{U}). ∎

Lemma 3.4.

Let 𝐔\bm{U} be the linear subspace of k\mathbb{R}^{k} obtained by Lemma 3.2. If for any 𝐦k{𝟎}\mathbf{m}\in\mathbb{Z}^{k}\setminus\{\mathbf{0}\}, 𝛏𝐦d\bm{\xi}\cdot\mathbf{m}\notin\mathbb{Z}^{d}. Then 𝐔=k\bm{U}=\mathbb{R}^{k} and hence, H=GH=G.

Proof.

Suppose 𝑼k\bm{U}\neq\mathbb{R}^{k}, then dim𝑼<k\dim\bm{U}<k. Since 𝑼k\bm{U}\cap\mathbb{Z}^{k} is a lattice in 𝑼\bm{U}, there exists a nonzero 𝐯k𝑼\mathbf{v}\in\mathbb{Z}^{k}\cap\bm{U}^{\perp}. Since 𝝃𝐯(d)k𝐯+L(𝑼)𝐯=(d)k𝐯\bm{\xi}\cdot\mathbf{v}\in(\mathbb{Q}^{d})^{k}\cdot\mathbf{v}+L(\bm{U})\cdot\mathbf{v}=(\mathbb{Q}^{d})^{k}\cdot\mathbf{v}, we can choose 𝐦\mathbf{m} to be a suitable integral multiple of 𝐯\mathbf{v} such that 𝝃𝐦d\bm{\xi}\cdot\mathbf{m}\in\mathbb{Z}^{d}, this contradicts to the assumption of the lemma. ∎

4. Unipotent invariance

The collection of all probability measures on the one point compactification XX^{*} of XX is a compact space in weak*-topology. Therefore, for any 𝐬𝒰\mathbf{s}\in\mathcal{U}, after possibly passing to a subsequence, we have

1T0Tδatu𝝋(𝐬)Γ𝑑tTμ𝐬 in weak* topology,\displaystyle\frac{1}{T}\int_{0}^{T}\delta_{a_{t}u_{\bm{\varphi}}(\mathbf{s})\Gamma}dt\xrightarrow{T\to\infty}\mu_{\mathbf{s}}\text{ in weak* topology},

for some probability measure μ𝐬\mu_{\mathbf{s}} on XX^{*}. Throughout this section, the function 𝝋\bm{\varphi} is assumed to be C1C^{1} and satisfy (𝝋(𝐬))>r𝟎(\bm{\varphi}(\mathbf{s}))_{>r}\equiv\mathbf{0}.

Proposition 4.1.

For a.e. 𝐬𝒰\mathbf{s}\in\mathcal{U}, μ𝐬\mu_{\mathbf{s}} is UU-invariant.

Proof.

Since 𝒰\mathcal{U} is a bounded open subset of Matr×(dr)()Mat_{r\times(d-r)}(\mathbb{R}), it is enough to prove the proposition for a.e. 𝐬\mathbf{s} in an open cube of 𝒰\mathcal{U}.

We may choose an open interval 𝕀\mathbb{I}\subset\mathbb{R} such that 𝕀r(dr)𝒰\mathbb{I}^{r(d-r)}\subset\mathcal{U}. For 1ir,1jdr1\leq i\leq r,1\leq j\leq d-r, let 𝐄ijMatr×(dr)()\mathbf{E}_{ij}\in Mat_{r\times(d-r)}(\mathbb{R}) be the matrix with 11 in (i,j)(i,j)-th entry and zero otherwise.

If s1,s2s_{1},s_{2} are two real numbers linearly independent over \mathbb{Q}, then the closure of the subgroup generated by {u(s1𝐄ij),u(s2𝐄ij):1ir,1jdr}\{u(s_{1}\mathbf{E}_{ij}),u(s_{2}\mathbf{E}_{ij}):1\leq i\leq r,1\leq j\leq d-r\} is UU.

Therefore, given ss^{\prime}\in\mathbb{R}, without loss of generality, it suffices to prove that for a.e. 𝐬𝕀r(dr)\mathbf{s}\in\mathbb{I}^{r(d-r)}, the limit measure μ𝐬\mu_{\mathbf{s}} is invariant under u(s𝐄11)u(s^{\prime}\mathbf{E}_{11}).

Note that there exists a countable dense subset of Cc(G/Γ)C_{c}(G/\Gamma) consisting of smooth functions. Let ψCc(G/Γ)\psi\in C_{c}^{\infty}(G/\Gamma). For t>0t>0 and 𝐰Matr×(dr)()\mathbf{w}\in Mat_{r\times(d-r)}(\mathbb{R}), define

ψt(𝐰)=ψ(atu𝝋(𝐰)Γ)ψ(u(s𝐄11)atu𝝋(𝐰)Γ).\displaystyle\psi_{t}(\mathbf{w})=\psi(a_{t}u_{\bm{\varphi}}(\mathbf{w})\Gamma)-\psi(u(s^{\prime}\mathbf{E}_{11})a_{t}u_{\bm{\varphi}}(\mathbf{w})\Gamma).

Hence, we only need to show that for this ψ\psi, for a.e. 𝐰𝕀r(dr)\mathbf{w}\in\mathbb{I}^{r(d-r)},

1T0Tψt(𝐰)𝑑tT0.\displaystyle\frac{1}{T}\int_{0}^{T}\psi_{t}(\mathbf{w})dt\xrightarrow{T\to\infty}0.

This follows from Theorem 4.2 and Lemma 4.3 as follows. ∎

Theorem 4.2.

[13, Theorem 3.1] Let (Y,μ)(Y,\mu) be a probability space. Let F:Y×+F:Y\times\mathbb{R}^{+}\to\mathbb{R} be a bounded measurable function. Suppose that there exist δ>0\delta>0 and c>0c>0 such that for any lt0l\geq t\geq 0,

|YF(x,t)F(x,l)𝑑μ(x)|ceδmin(t,lt),\displaystyle|\int_{Y}F(x,t)F(x,l)d\mu(x)|\leq c\cdot e^{-\delta\min(t,l-t)}, (4.1)

then given any ϵ>0\epsilon>0, for μ\mu-a.e. yYy\in Y,

1T0TF(y,t)𝑑t=o(T12log23+ϵT).\displaystyle\frac{1}{T}\int_{0}^{T}F(y,t)dt=o(T^{-\frac{1}{2}}\cdot log^{\frac{2}{3}+\epsilon}T).
Lemma 4.3.

There exist c>0c>0 such that for any t,l>0t,l>0,

|𝕀r(dr)ψt(𝐰)ψl(𝐰)𝑑𝐰|ce|lt|.\displaystyle|\int_{\mathbb{I}^{r(d-r)}}\psi_{t}(\mathbf{w})\psi_{l}(\mathbf{w})d\mathbf{w}|\leq c\cdot e^{-|l-t|}.
Proof.

In the following proof, for positive valued functions f,gf,g, we write f=O(g)f=O(g) if there exists a positive constant CC depending only on ψ,𝝋,s\psi,\bm{\varphi},s^{\prime} and |𝕀||\mathbb{I}| (these are fixed throughout the proof) such that fCgf\leq Cg. Also, for any positive number ϵ>0\epsilon>0, we let OG(ϵ)O_{G}(\epsilon) denote a group element in a O(ϵ)O(\epsilon)-neighborhood of IdId in GG.

Without loss of generality, we assume that ltl\geq t. For s0𝕀s_{0}\in\mathbb{I}, consider the interval

𝕀(s0)=(s0|𝕀|ed(l+t)2,s0+|𝕀|ed(l+t)2)\displaystyle\mathbb{I}(s_{0})=(s_{0}-|\mathbb{I}|e^{-\frac{d(l+t)}{2}},s_{0}+|\mathbb{I}|e^{-\frac{d(l+t)}{2}})

such that 𝕀(s0)𝕀\mathbb{I}(s_{0})\subset\mathbb{I}. For any s𝕀(s0)s\in\mathbb{I}(s_{0}), and any 𝐰{0}×𝕀r(dr)1\mathbf{w}\in\{0\}\times\mathbb{I}^{r(d-r)-1},

|ψt(s𝐄11+𝐰)ψt(s0𝐄11+𝐰)||ψ(atu𝝋(s𝐄11+𝐰)Γ)ψ(atu𝝋(s0𝐄11+𝐰)Γ)|\displaystyle|\psi_{t}(s\mathbf{E}_{11}+\mathbf{w})-\psi_{t}(s_{0}\mathbf{E}_{11}+\mathbf{w})|\leq|\psi(a_{t}u_{\bm{\varphi}}(s\mathbf{E}_{11}+\mathbf{w})\Gamma)-\psi(a_{t}u_{\bm{\varphi}}(s_{0}\mathbf{E}_{11}+\mathbf{w})\Gamma)|
+|ψ(u(s𝐄𝟏𝟏)atu𝝋(s𝐄𝟏𝟏+𝐰)Γ)ψ(u(s𝐄𝟏𝟏)atu𝝋(s0𝐄𝟏𝟏+𝐰)Γ)|.\displaystyle+|\psi(u(s^{\prime}\mathbf{\mathbf{E}_{11}})a_{t}u_{\bm{\varphi}}(s\mathbf{\mathbf{E}_{11}}+\mathbf{w})\Gamma)-\psi(u(s^{\prime}\mathbf{\mathbf{E}_{11}})a_{t}u_{\bm{\varphi}}(s_{0}\mathbf{\mathbf{E}_{11}}+\mathbf{w})\Gamma)|.

Note that

atu𝝋(s𝐄𝟏𝟏+𝐰)=at(u(s𝐄11+𝐰),𝝋(s𝐄11+𝐰))\displaystyle a_{t}u_{\bm{\varphi}}(s\mathbf{\mathbf{E}_{11}}+\mathbf{w})=a_{t}(u(s\mathbf{E}_{11}+\mathbf{w}),\bm{\varphi}(s\mathbf{E}_{11}+\mathbf{w}))
=at(u(s0𝐄11+𝐰+(ss0)𝐄11),𝝋(s0𝐄11+𝐰)+𝝋(s𝐄11+𝐰)𝝋(s0𝐄11+𝐰))\displaystyle=a_{t}(u(s_{0}\mathbf{E}_{11}+\mathbf{w}+(s-s_{0})\mathbf{E}_{11}),\bm{\varphi}(s_{0}\mathbf{E}_{11}+\mathbf{w})+\bm{\varphi}(s\mathbf{E}_{11}+\mathbf{w})-\bm{\varphi}(s_{0}\mathbf{E}_{11}+\mathbf{w}))
=(u(edt(ss0)𝐄𝟏𝟏),e(dr)t(ss0)11𝝋)atu𝝋(s0𝐄𝟏𝟏+𝐰).\displaystyle=(u(e^{dt}(s-s_{0})\mathbf{\mathbf{E}_{11}}),e^{(d-r)t}(s-s_{0})\partial_{11}\bm{\varphi})a_{t}u_{\bm{\varphi}}(s_{0}\mathbf{\mathbf{E}_{11}}+\mathbf{w}).

where the last equality follows by mean value theorem (for simplicity of notations, by uniform boundedness of 11φ||\partial_{11}\varphi||_{\infty} on 𝒰\mathcal{U}, we write 11𝝋\partial_{11}\bm{\varphi} for 11𝝋(s~𝐄𝟏𝟏+𝐰)\partial_{11}\bm{\varphi}(\tilde{s}\mathbf{\mathbf{E}_{11}}+\mathbf{w}) with arbitrary s~\tilde{s} ). Since

e(dr)t|ss0|edt|ss0|ed(l+t)2edt|𝕀|=ed(lt)2|𝕀|,\displaystyle e^{(d-r)t}|s-s_{0}|\leq e^{dt}|s-s_{0}|\leq e^{-\frac{d(l+t)}{2}}\cdot e^{dt}|\mathbb{I}|=e^{-\frac{d(l-t)}{2}}|\mathbb{I}|,

we have

atu𝝋(s𝐄𝟏𝟏+𝐰)\displaystyle a_{t}u_{\bm{\varphi}}(s\mathbf{\mathbf{E}_{11}}+\mathbf{w}) =OG(ed(lt)2|𝕀|)atu𝝋(s0𝐄𝟏𝟏+𝐰)\displaystyle=O_{G}(e^{-\frac{d(l-t)}{2}}|\mathbb{I}|)a_{t}u_{\bm{\varphi}}(s_{0}\mathbf{\mathbf{E}_{11}}+\mathbf{w})
=OG(ed(lt)2)atu𝝋(s0𝐄𝟏𝟏+𝐰).\displaystyle=O_{G}(e^{-\frac{d(l-t)}{2}})a_{t}u_{\bm{\varphi}}(s_{0}\mathbf{\mathbf{E}_{11}}+\mathbf{w}).

Likewise,

u(s𝐄𝟏𝟏)atu𝝋(s𝐄𝟏𝟏+𝐰)=OG(ed(lt)2)u(s𝐄𝟏𝟏)atu𝝋(s0𝐄𝟏𝟏+𝐰).\displaystyle u(s^{\prime}\mathbf{\mathbf{E}_{11}})a_{t}u_{\bm{\varphi}}(s\mathbf{\mathbf{E}_{11}}+\mathbf{w})=O_{G}(e^{-\frac{d(l-t)}{2}})u(s^{\prime}\mathbf{\mathbf{E}_{11}})a_{t}u_{\bm{\varphi}}(s_{0}\mathbf{\mathbf{E}_{11}}+\mathbf{w}).

Since ψCc(G/Γ)\psi\in C_{c}^{\infty}(G/\Gamma), ψ\psi is Lipschitz, and hence

|ψt(s𝐄𝟏𝟏+𝐰)ψt(s0𝐄𝟏𝟏+𝐰)|=O(ed(lt)2).\displaystyle|\psi_{t}(s\mathbf{\mathbf{E}_{11}}+\mathbf{w})-\psi_{t}(s_{0}\mathbf{\mathbf{E}_{11}}+\mathbf{w})|=O(e^{-\frac{d(l-t)}{2}}).

Therefore,

𝕀(s0)ψt(s𝐄𝟏𝟏+𝐰)ψl(s𝐄𝟏𝟏+𝐰)𝑑s\displaystyle\int_{\mathbb{I}(s_{0})}\psi_{t}(s\mathbf{\mathbf{E}_{11}}+\mathbf{w})\psi_{l}(s\mathbf{\mathbf{E}_{11}}+\mathbf{w})ds
=𝕀(s0)(ψt(s0𝐄𝟏𝟏+𝐰)+O(ed(lt)2))ψl(s𝐄𝟏𝟏+𝐰)𝑑s\displaystyle=\int_{\mathbb{I}(s_{0})}(\psi_{t}(s_{0}\mathbf{\mathbf{E}_{11}}+\mathbf{w})+O(e^{-\frac{d(l-t)}{2}}))\cdot\psi_{l}(s\mathbf{\mathbf{E}_{11}}+\mathbf{w})ds
=ψt(s0𝐄𝟏𝟏+𝐰)𝕀(s0)ψl(s𝐄𝟏𝟏+𝐰)𝑑s+O(ed(lt)2)|𝕀(s0)|.\displaystyle=\psi_{t}(s_{0}\mathbf{\mathbf{E}_{11}}+\mathbf{w})\int_{\mathbb{I}(s_{0})}\psi_{l}(s\mathbf{\mathbf{E}_{11}}+\mathbf{w})ds+O(e^{-\frac{d(l-t)}{2}})\cdot|\mathbb{I}(s_{0})|. (4.2)

Now we estimate 𝕀(s0)ψl(s𝐄𝟏𝟏+𝐰)𝑑s\int_{\mathbb{I}(s_{0})}\psi_{l}(s\mathbf{\mathbf{E}_{11}}+\mathbf{w})ds. Note that

u(s𝐄11)alu𝝋(s𝐄𝟏𝟏+𝐰)\displaystyle u(s^{\prime}\mathbf{E}_{11})a_{l}u_{\bm{\varphi}}(s\mathbf{\mathbf{E}_{11}}+\mathbf{w})
=al(u((s+edls)𝐄𝟏𝟏+𝐰),𝝋(s𝐄𝟏𝟏+𝐰))\displaystyle=a_{l}(u((s+e^{-dl}s^{\prime})\mathbf{\mathbf{E}_{11}}+\mathbf{w}),\bm{\varphi}(s\mathbf{\mathbf{E}_{11}}+\mathbf{w}))
=al(Id,𝝋(s𝐄𝟏𝟏+𝐰)𝝋((edls+s)𝐄𝟏𝟏+𝐰))alalu𝝋((s+edls)𝐄𝟏𝟏+𝐰)\displaystyle=a_{l}(Id,\bm{\varphi}(s\mathbf{\mathbf{E}_{11}}+\mathbf{w})-\bm{\varphi}((e^{-dl}s^{\prime}+s)\mathbf{\mathbf{E}_{11}}+\mathbf{w}))\cdot a_{-l}a_{l}u_{\bm{\varphi}}((s+e^{-dl}s^{\prime})\mathbf{\mathbf{E}_{11}}+\mathbf{w})
=al(Id,edls11𝝋)alalu𝝋((s+edls)𝐄𝟏𝟏+𝐰)\displaystyle=a_{l}(Id,e^{-dl}s^{\prime}\partial_{11}\bm{\varphi})a_{-l}a_{l}u_{\bm{\varphi}}((s+e^{-dl}s^{\prime})\mathbf{\mathbf{E}_{11}}+\mathbf{w})
=OG(erl)alu𝝋((s+edls)𝐄𝟏𝟏+𝐰).\displaystyle=O_{G}(e^{-rl})a_{l}u_{\bm{\varphi}}((s+e^{-dl}s^{\prime})\mathbf{\mathbf{E}_{11}}+\mathbf{w}).

As ψ\psi is Lipschitz,

ψl(s𝐄𝟏𝟏+𝐰)=ψ(alu𝝋(s𝐄𝟏𝟏+𝐰)Γ)ψ(alu𝝋((s+edls)𝐄𝟏𝟏+𝐰)Γ)+O(erl).\displaystyle\psi_{l}(s\mathbf{\mathbf{E}_{11}}+\mathbf{w})=\psi(a_{l}u_{\bm{\varphi}}(s\mathbf{\mathbf{E}_{11}}+\mathbf{w})\Gamma)-\psi(a_{l}u_{\bm{\varphi}}((s+e^{-dl}s^{\prime})\mathbf{\mathbf{E}_{11}}+\mathbf{w})\Gamma)+O(e^{-rl}).

Since 𝕀(s0)\mathbb{I}(s_{0}) and 𝕀(s0)+edls\mathbb{I}(s_{0})+e^{-dl}s^{\prime} overlap except for a length of O(edl)O(e^{-dl}), we have

𝕀(s0)ψl(s𝐄𝟏𝟏+𝐰)𝑑s\displaystyle\int_{\mathbb{I}(s_{0})}\psi_{l}(s\mathbf{\mathbf{E}_{11}}+\mathbf{w})ds
=𝕀(s0)ψ(alu𝝋(s𝐄𝟏𝟏+𝐰)Γ)ψ(alu𝝋((s+edls)𝐄𝟏𝟏+𝐰)Γ)+O(erl)ds\displaystyle=\int_{\mathbb{I}(s_{0})}\psi(a_{l}u_{\bm{\varphi}}(s\mathbf{\mathbf{E}_{11}}+\mathbf{w})\Gamma)-\psi(a_{l}u_{\bm{\varphi}}((s+e^{-dl}s^{\prime})\mathbf{\mathbf{E}_{11}}+\mathbf{w})\Gamma)+O(e^{-rl})ds
=O(edl)+O(erl)|𝕀(s0)|\displaystyle=O(e^{-dl})+O(e^{-rl})|\mathbb{I}(s_{0})|
=O(ed(lt)2)|𝕀(s0)|+O(e(lt))|𝕀(s0)|\displaystyle=O(e^{-\frac{d(l-t)}{2}})|\mathbb{I}(s_{0})|+O(e^{-(l-t)})|\mathbb{I}(s_{0})|
=O(e(lt))|𝕀(s0)|.\displaystyle=O(e^{-(l-t)})|\mathbb{I}(s_{0})|.

Now we consider the partition 𝕀=j=1p𝕀j\mathbb{I}=\bigcup_{j=1}^{p}\mathbb{I}_{j} such that 𝕀j=[sj1,sj]\mathbb{I}_{j}=[s_{j-1},s_{j}] with sjsj1=2ed(l+t)2|𝕀|s_{j}-s_{j-1}=2e^{-\frac{d(l+t)}{2}}|\mathbb{I}| for 1jp11\leq j\leq p-1, and spsp12ed(l+t)2|𝕀|s_{p}-s_{p-1}\leq 2e^{-\frac{d(l+t)}{2}}|\mathbb{I}|. By (4), we have

𝕀ψt(s𝐄𝟏𝟏+𝐰)ψl(s𝐄𝟏𝟏+𝐰)𝑑s\displaystyle\int_{\mathbb{I}}\psi_{t}(s\mathbf{\mathbf{E}_{11}}+\mathbf{w})\psi_{l}(s\mathbf{\mathbf{E}_{11}}+\mathbf{w})ds
=j=1p𝕀jψt(s𝐄𝟏𝟏+𝐰)ψl(s𝐞𝟏𝟏+𝐰)𝑑s\displaystyle=\sum_{j=1}^{p}\int_{\mathbb{I}_{j}}\psi_{t}(s\mathbf{\mathbf{E}_{11}}+\mathbf{w})\psi_{l}(s\mathbf{\mathbf{e}_{11}}+\mathbf{w})ds
=j=1p1𝕀jψt(s𝐄𝟏𝟏+𝐰)ψl(s𝐄𝟏𝟏+𝐰)𝑑s+O(ed(l+t)2)|𝕀|\displaystyle=\sum_{j=1}^{p-1}\int_{\mathbb{I}_{j}}\psi_{t}(s\mathbf{\mathbf{E}_{11}}+\mathbf{w})\psi_{l}(s\mathbf{\mathbf{E}_{11}}+\mathbf{w})ds+O(e^{-\frac{d(l+t)}{2}})|\mathbb{I}|
=j=1p1O(e(lt))|𝕀j|+O(ed(l+t)2)|𝕀|\displaystyle=\sum_{j=1}^{p-1}O(e^{-(l-t)})|\mathbb{I}_{j}|+O(e^{-\frac{d(l+t)}{2}})|\mathbb{I}|
=O(e(lt))|𝕀|=O(e(lt)).\displaystyle=O(e^{-(l-t)})|\mathbb{I}|=O(e^{-(l-t)}).

The above estimate holds for any 𝐰{𝟎}×𝕀r(dr)1\mathbf{w}\in\{\mathbf{0}\}\times\mathbb{I}^{r(d-r)-1}. Now the lemma follows from the above estimate and Fubini’s theorem. ∎

5. Margulis’ height function

In this section, we will recall the definition of Margulis’ height function on SLd()/SLd()SL_{d}(\mathbb{R})/SL_{d}(\mathbb{Z}) and its uniform contraction property .

Margulis’ height function was first introduced in [6] and later developed in several papers (see for example [1][25]). It measures the depth of elements of XX into the cusps. It has been used to study equidistribution problem for certain unbounded functions (cf.[6][7][16]) and random walks on homogeneous spaces (cf.[1][5]).

We start with the vector space 𝐕=d=0idid\mathbf{V}=\wedge^{*}\mathbb{R}^{d}=\bigoplus_{0\leq i\leq d}\wedge^{i}\mathbb{R}^{d}, where G=SLd()G^{\prime}=SL_{d}(\mathbb{R}) acts on 𝐕\mathbf{V} naturally.

Let Δ\Delta be a lattice in d\mathbb{R}^{d}. We say that a subspace LL of d\mathbb{R}^{d} is Δ\Delta-rational if LΔL\cap\Delta is a lattice in LL. For any Δ\Delta-rational subspace LL, denote d(L)d(L) or dΔ(L)d_{\Delta}(L) the volume of L/LΔL/L\cap\Delta. Note that d(L)d(L) is the norm of u1u2ulu_{1}\wedge u_{2}\wedge\cdots\wedge u_{l} in 𝐕\mathbf{V}, where {ui}1il\{u_{i}\}_{1\leq i\leq l} is a \mathbb{Z}-basis of LΔL\cap\Delta. If L={𝟎}L=\{\mathbf{0}\}, we set d(L)=1d(L)=1.

For any lattice Δ\Delta, we define for 0id0\leq i\leq d,

αi(Δ):=sup{1d(L):L is a Δ-rational subspace of dimension i}.\displaystyle\alpha_{i}(\Delta):=\sup\left\{\frac{1}{d(L)}:L\text{ is a }\Delta\text{-rational subspace of dimension }i\right\}.
Proposition 5.1.

There exists a continuous map α~:SLd()/SLd()[1,]\tilde{\alpha}:SL_{d}(\mathbb{R})/SL_{d}(\mathbb{Z})\to[1,\infty] and b1>0b_{1}>0 such that for BB a bounded open box in Matr×(dr)()Mat_{r\times(d-r)}(\mathbb{R}), for all t>0t>0 large enough and for any unimodular lattice Λ\Lambda of d\mathbb{R}^{d},

1|B|Bα~(atu(𝐬)Λ)𝑑𝐬<2r(dr)2α~(Λ)+b1,\displaystyle\frac{1}{|B|}\int_{B}\tilde{\alpha}(a_{t}u(\mathbf{s})\Lambda)d\mathbf{s}<2^{-r(d-r)-2}\tilde{\alpha}(\Lambda)+b_{1},

and there exists ν>0\nu>0 such that

α1(Λ)να~(Λ).\displaystyle\alpha_{1}(\Lambda)^{\nu}\leq\tilde{\alpha}(\Lambda).

Moreover, a measurable subset KK of SLd()/SLd()SL_{d}(\mathbb{R})/SL_{d}(\mathbb{Z}) is precompact if there exists N>0N>0 such that

K{xX:α~(x)N}.\displaystyle K\subset\{x\in X:\tilde{\alpha}(x)\leq N\}.
Proof.

Define α~=ϵq(1)i=0dϵq(i)αiν\tilde{\alpha}=\epsilon^{-q(1)}\cdot\sum_{i=0}^{d}\epsilon^{q(i)}\cdot\alpha_{i}^{\nu}, where ϵ,ν>0\epsilon,\nu>0 are sufficiently small numbers and q(i)=i(di)q(i)=i(d-i). The fact that α~\tilde{\alpha} satisfies conclusion of Proposition 5.1 will follow from [25, Lemma 4.1]. ∎

The function α~\tilde{\alpha} above is the Margulis’ height function that we need in our setting.

Remark 5.2.

The function α~\tilde{\alpha} satisfies Lipschitz property as follows: For any bounded neighborhood 𝒱\mathcal{V} of ee of SLd()SL_{d}(\mathbb{R}), there exists M¯>0\overline{M}>0 such that for any xSLd()/SLd()x\in SL_{d}(\mathbb{R})/SL_{d}(\mathbb{Z}), any g𝒱g\in\mathcal{V},

α~(gx)M¯α~(x).\displaystyle\tilde{\alpha}(gx)\leq\overline{M}\tilde{\alpha}(x).

Indeed, M¯>0\overline{M}>0 is the maximum of operator norms of elements in 𝒱\mathcal{V} acting on 𝐕\mathbf{V}.

6. Mixed height function

In this section, we will construct a mixed height function, which is crucial for us to prove Proposition 2.1. The main result of this section is the following:

Proposition 6.1.

Let 𝛗\bm{\varphi} be a C1C^{1} map from 𝒰\mathcal{U} to (d)k(\mathbb{R}^{d})^{k} satisfying (𝛗(𝐬))>r𝟎(\bm{\varphi}(\mathbf{s}))_{>r}\equiv\mathbf{0} for any 𝐬𝒰\mathbf{s}\in\mathcal{U}. For any 𝐦k{𝟎}\mathbf{m}\in\mathbb{Z}^{k}\setminus\{\mathbf{0}\}, any closed cube I𝒰Bad𝐦I\subset\mathcal{U}\setminus Bad_{\mathbf{m}} (for Bad𝐦Bad_{\mathbf{m}}, see (2)), there are t>0t>0 sufficiently large (depending on II, 𝐦\mathbf{m}) and measurable function β𝐦:G/Γ(0,]\beta_{\mathbf{m}}:G/\Gamma\to(0,\infty] such that the following hold:

(1) For any l>0l>0, {xG/Γ:β𝐦(x)l}\{x\in G/\Gamma:\beta_{\mathbf{m}}(x)\leq l\} is compact;

(2) For any xG/Γx\in G/\Gamma, β𝐦(x)=\beta_{\mathbf{m}}(x)=\infty if and only if xX𝐦x\in X_{\mathbf{m}};

(3) Given any n0n\in\mathbb{Z}_{\geq 0}, a box JIJ\subset I with J=i=1r(dr)JiJ=\prod_{i=1}^{r(d-r)}J_{i}, where JiJ_{i}\subset\mathbb{R} and |Ji|2ednt|J_{i}|\leq 2e^{-dnt} for all ii. There exists M~1>0\tilde{M}_{1}>0 such that for any 𝐬,𝐬~J\mathbf{s},\tilde{\mathbf{s}}\in J, one has

β𝐦(antu𝝋(𝐬~)Γ)M~1β𝐦(antu𝝋(𝐬)Γ);\displaystyle\beta_{\mathbf{m}}(a_{nt}u_{\bm{\varphi}}(\tilde{\mathbf{s}})\Gamma)\leq\tilde{M}_{1}\beta_{\mathbf{m}}(a_{nt}u_{\bm{\varphi}}(\mathbf{s})\Gamma);

(4) There exists M~2>0\tilde{M}_{2}>0 depending on tt, for any n0n\in\mathbb{Z}_{\geq 0}, any 𝐬I\mathbf{s}\in I and any τ\tau\in\mathbb{R} with |τ|t|\tau|\leq t, one has

β𝐦(aτantu𝝋(𝐬)Γ)M~2β𝐦(antu𝝋(𝐬)Γ);\displaystyle\beta_{\mathbf{m}}(a_{\tau}a_{nt}u_{\bm{\varphi}}(\mathbf{s})\Gamma)\leq\tilde{M}_{2}\beta_{\mathbf{m}}(a_{nt}u_{\bm{\varphi}}(\mathbf{s})\Gamma);

(5) There exists b>0b>0 such that the following holds: for any n0n\in\mathbb{Z}_{\geq 0} and any box JIJ\subset I with J=i=1r(dr)JiJ=\prod_{i=1}^{r(d-r)}J_{i} satisfying either n1n\geq 1 and |Ji|ednt|J_{i}|\geq e^{-dnt} for all 1ir(dr)1\leq i\leq r(d-r), or n=0n=0 and J=IJ=I, one has

Jβ𝐦(a(n+1)tu𝝋(𝐬)Γ)𝑑𝐬12Jβ𝐦(antu𝝋(𝐬)Γ)𝑑𝐬+b|J|.\displaystyle\int_{J}\beta_{\mathbf{m}}(a_{(n+1)t}u_{\bm{\varphi}}(\mathbf{s})\Gamma)d\mathbf{s}\leq\frac{1}{2}\int_{J}\beta_{\mathbf{m}}(a_{nt}u_{\bm{\varphi}}(\mathbf{s})\Gamma)d\mathbf{s}+b|J|.
Remark 6.2.

The function β𝐦\beta_{\mathbf{m}} in Proposition 6.1 is the desired mixed height function.

From now on until the end of this section, we will fix a closed cube I𝒰Bad𝐦I\subset\mathcal{U}\setminus Bad_{\mathbf{m}}. Recall that the finite number M1M_{1} is defined as in (2.1). By the choice of II, and the fact that there are only finitely many 𝐚dr\mathbf{a}\in\mathbb{Z}^{d-r} satisfying 𝐚M1𝐦\left\lVert\mathbf{a}\right\rVert_{\infty}\leq M_{1}\left\lVert\mathbf{m}\right\rVert, we obtain σ>0\sigma>0 such that

inf𝐬I{(𝝋(𝐬))r𝐦𝐬𝐚𝐛:𝐚M1𝐦,𝐚dr,𝐛r}=σ.\displaystyle\inf_{\mathbf{s}\in I}\{\left\lVert(\bm{\varphi}(\mathbf{s}))_{\leq r}\cdot\mathbf{m}-\mathbf{s}\cdot\mathbf{a}-\mathbf{b}\right\rVert_{\infty}:\left\lVert\mathbf{a}\right\rVert_{\infty}\leq M_{1}\left\lVert\mathbf{m}\right\rVert,\mathbf{a}\in\mathbb{Z}^{d-r},\mathbf{b}\in\mathbb{Z}^{r}\}=\sigma. (6.1)
Remark 6.3.

By (6.1), we can choose a closed neighborhood II^{\prime} of II such that II^{\prime} is a closed cube contained in 𝒰\mathcal{U} and satisfies

inf𝐬I{(𝝋(𝐬))r𝐦𝐬𝐚𝐛:𝐚M1𝐦,𝐚dr,𝐛r}=σ2.\displaystyle\inf_{\mathbf{s}\in I^{\prime}}\{\left\lVert(\bm{\varphi}(\mathbf{s}))_{\leq r}\cdot\mathbf{m}-\mathbf{s}\cdot\mathbf{a}-\mathbf{b}\right\rVert_{\infty}:\left\lVert\mathbf{a}\right\rVert_{\infty}\leq M_{1}\left\lVert\mathbf{m}\right\rVert,\mathbf{a}\in\mathbb{Z}^{d-r},\mathbf{b}\in\mathbb{Z}^{r}\}=\frac{\sigma}{2}.

Next we construct a suitable function measuring the distance to the closed submanifold X𝐦X_{\mathbf{m}}. For 𝐦k{𝟎}\mathbf{m}\in\mathbb{Z}^{k}\setminus\{\mathbf{0}\}, consider the quotient space

(d)𝐦k:={𝐯(d)k:𝐯𝐦d}/,\displaystyle(\mathbb{R}^{d})^{k}_{\mathbf{m}}:=\{\mathbf{v}\in(\mathbb{R}^{d})^{k}:\mathbf{v}\cdot\mathbf{m}\in\mathbb{Z}^{d}\}/\sim,

where 𝐯𝐯\mathbf{v}\sim\mathbf{v}^{\prime} if and only if 𝐯𝐦=𝐯𝐦\mathbf{v}\cdot\mathbf{m}=\mathbf{v}^{\prime}\cdot\mathbf{m}. One can directly verify that \sim is an equivalence relation.

Lemma 6.4.

For any (g,𝐯)G(g,\mathbf{v})\in G, there exists at most one 𝐯0(d)𝐦k\mathbf{v}_{0}\in(\mathbb{R}^{d})^{k}_{\mathbf{m}} such that

(𝐯g𝐯0)𝐦<12inf𝐰d{𝟎}g𝐰.\displaystyle\left\lVert(\mathbf{v}-g\mathbf{v}_{0})\cdot\mathbf{m}\right\rVert<\frac{1}{2}\inf_{\mathbf{w}\in\mathbb{Z}^{d}\setminus\{\mathbf{0}\}}\left\lVert g\mathbf{w}\right\rVert. (6.2)
Proof.

Suppose there are two vectors 𝐯0\mathbf{v}_{0} and 𝐯0\mathbf{v}_{0}^{\prime} in (d)𝐦k(\mathbb{R}^{d})^{k}_{\mathbf{m}} satisfying (6.2) such that 𝐯0≁𝐯0\mathbf{v}_{0}\not\sim\mathbf{v}_{0}^{\prime}. Then

g(𝐯0𝐯0)𝐦(𝐯g𝐯0)𝐦+(𝐯g𝐯0)𝐦<inf𝐰d{𝟎}g𝐰.\displaystyle\left\lVert g(\mathbf{v}_{0}-\mathbf{v}_{0}^{\prime})\cdot\mathbf{m}\right\rVert\leq\left\lVert(\mathbf{v}-g\mathbf{v}_{0})\cdot\mathbf{m}\right\rVert+\left\lVert(\mathbf{v}-g\mathbf{v}_{0}^{\prime})\cdot\mathbf{m}\right\rVert<\inf_{\mathbf{w}\in\mathbb{Z}^{d}\setminus\{\mathbf{0}\}}\left\lVert g\mathbf{w}\right\rVert.

But since 𝐯0≁𝐯0\mathbf{v}_{0}\not\sim\mathbf{v}_{0}^{\prime}, (𝐯0𝐯0)𝐦d{𝟎}(\mathbf{v}_{0}-\mathbf{v}_{0}^{\prime})\cdot\mathbf{m}\in\mathbb{Z}^{d}\setminus\{\mathbf{0}\}. This is a contradiction. ∎

Definition 6.5.

Let 𝐦k{𝟎}\mathbf{m}\in\mathbb{Z}^{k}\setminus\{\mathbf{0}\}. For any (g,𝐯)G(g,\mathbf{v})\in G, we say that 𝝃g,v(d)𝐦k\bm{\xi}_{g,v}\in(\mathbb{R}^{d})_{\mathbf{m}}^{k} exists if 𝝃g,v\bm{\xi}_{g,v} satisfies (6.2) in the place of 𝐯0\mathbf{v}_{0}. By convention, we set 𝝃g,𝐯=\bm{\xi}_{g,\mathbf{v}}=\infty if it does not exist. Define the function

α𝐦(g,𝐯)={(𝐯g𝝃g,𝐯)𝐦1, If 𝝃g,𝐯 exists; 1, if 𝝃g,𝐯=.\displaystyle\alpha_{\mathbf{m}}(g,\mathbf{v})=\begin{cases}\left\lVert(\mathbf{v}-g\bm{\xi}_{g,\mathbf{v}})\cdot\mathbf{m}\right\rVert^{-1},&\text{ If }\bm{\xi}_{g,\mathbf{v}}\text{ exists; }\\ 1,&\text{ if }\bm{\xi}_{g,\mathbf{v}}=\infty.\end{cases} (6.3)
Remark 6.6.

By Lemma 6.4, α𝐦\alpha_{\mathbf{m}} is a well-defined function on GG. By Minkowski’s first theorem, there is a constant 0<μd10<\mu_{d}\leq 1 such that if 𝝃g,𝐯\bm{\xi}_{g,\mathbf{v}} exists for (g,𝐯)(g,\mathbf{v}), then α𝐦(g,𝐯)>μd\alpha_{\mathbf{m}}(g,\mathbf{v})>\mu_{d}. Therefore, by definition of α𝐦\alpha_{\mathbf{m}}, we have α𝐦(g,𝐯)>μd\alpha_{\mathbf{m}}(g,\mathbf{v})>\mu_{d} for any (g,𝐯)G(g,\mathbf{v})\in G. Moreover, by (6.1), for any 𝐬I\mathbf{s}\in I, α𝐦(u𝝋(𝐬))σ1\alpha_{\mathbf{m}}(u_{\bm{\varphi}}(\mathbf{s}))\leq\sigma^{-1}.

Lemma 6.7.

α𝐦\alpha_{\mathbf{m}} is a well-defined function on G/ΓG/\Gamma. Moreover, α𝐦\alpha_{\mathbf{m}} is lower semi-continuous.

Proof.

Take any (g,𝐯)G(g,\mathbf{v})\in G and any (γ,𝐯)Γ(\gamma,\mathbf{v}^{\prime})\in\Gamma.

If 𝝃g,𝐯\bm{\xi}_{g,\mathbf{v}} exists, then γ1(𝝃g,𝐯+𝐯)(d)𝐦k\gamma^{-1}(\bm{\xi}_{g,\mathbf{v}}+\mathbf{v}^{\prime})\in(\mathbb{R}^{d})^{k}_{\mathbf{m}}. Note that

(𝐯+g𝐯gγγ1(𝝃g,𝐯+𝐯))𝐦=(𝐯g𝝃g,𝐯)𝐦<12inf𝐰d{𝟎}g𝐰.\displaystyle\left\lVert(\mathbf{v}+g\mathbf{v}^{\prime}-g\gamma\cdot\gamma^{-1}(\bm{\xi}_{g,\mathbf{v}}+\mathbf{v}^{\prime}))\cdot\mathbf{m}\right\rVert=\left\lVert(\mathbf{v}-g\bm{\xi}_{g,\mathbf{v}})\cdot\mathbf{m}\right\rVert<\frac{1}{2}\inf_{\mathbf{w}\in\mathbb{Z}^{d}\setminus\{\mathbf{0}\}}\left\lVert g\mathbf{w}\right\rVert.

Thus 𝝃gγ,𝐯+g𝐯=γ(𝝃g,𝐯+𝐯)\bm{\xi}_{g\gamma,\mathbf{v}+g\mathbf{v}^{\prime}}=\gamma(\bm{\xi}_{g,\mathbf{v}}+\mathbf{v}^{\prime}) exists and α𝐦(g,𝐯)=α𝐦(gγ,𝐯+g𝐯)\alpha_{\mathbf{m}}(g,\mathbf{v})=\alpha_{\mathbf{m}}(g\gamma,\mathbf{v}+g\mathbf{v}^{\prime}).

If 𝝃g,𝐯\bm{\xi}_{g,\mathbf{v}} does not exist, same argument as above shows that 𝝃gγ,𝐯+g𝐯\bm{\xi}_{g\gamma,\mathbf{v}+g\mathbf{v}^{\prime}} does not exist neither.

If 𝝃g,𝐯\bm{\xi}_{g,\mathbf{v}} exists, it is locally constant. Therefore, α𝐦\alpha_{\mathbf{m}} is lower semi-continuous. ∎

Let ν(0,1r(dr))\nu\in(0,\frac{1}{r(d-r)}) be a number satisfying Proposition 5.1. Let c=4(10r2d)ν2r(dr)c=4\cdot(10r^{2}d)^{\nu}\cdot 2^{r(d-r)} and t>0t>0 be a sufficiently large number (to be specified later). Define

β𝐦=α𝐦ν+ceνrtα~.\displaystyle\beta_{\mathbf{m}}=\alpha_{\mathbf{m}}^{\nu}+ce^{\nu rt}\tilde{\alpha}. (6.4)

We will prove that β𝐦\beta_{\mathbf{m}} satisfies properties (1)-(5) of Proposition 6.1.

Proof of Proposition 6.1 (1).

Since

{xX:β𝐦(x)l}{xX:α~(x)lc1eνrt},\displaystyle\{x\in X:\beta_{\mathbf{m}}(x)\leq l\}\subset\{x\in X:\tilde{\alpha}(x)\leq lc^{-1}e^{-\nu rt}\},

{xX:β𝐦(x)l}\{x\in X:\beta_{\mathbf{m}}(x)\leq l\} is precompact. As α~\tilde{\alpha} is continuous and α𝐦\alpha_{\mathbf{m}} is lower semicontinuous, {xX:β𝐦(x)l}\{x\in X:\beta_{\mathbf{m}}(x)\leq l\} is closed and thus compact. ∎

Proof of Proposition 6.1 (2).

If for some xG/Γx\in G/\Gamma, β𝐦(x)=\beta_{\mathbf{m}}(x)=\infty, then α𝐦(x)=\alpha_{\mathbf{m}}(x)=\infty as α~(x)<\tilde{\alpha}(x)<\infty. Let x=(g,𝐯)Γ/Γx=(g,\mathbf{v})\Gamma/\Gamma. By definition of α𝐦\alpha_{\mathbf{m}}, we have (𝐯g𝝃)𝐦=𝟎(\mathbf{v}-g\bm{\xi})\cdot\mathbf{m}=\mathbf{0} for some 𝝃(d)𝐦k\bm{\xi}\in(\mathbb{R}^{d})^{k}_{\mathbf{m}}. Note that g1𝐯𝐦=𝝃𝐦dg^{-1}\mathbf{v}\cdot\mathbf{m}=\bm{\xi}\cdot\mathbf{m}\in\mathbb{Z}^{d}. Hence (g,𝐯)Γ=(g,gg1𝐯)ΓX𝐦(g,\mathbf{v})\Gamma=(g,gg^{-1}\mathbf{v})\Gamma\in X_{\mathbf{m}}.

Conversely, if x=(g,𝐯)Γ/ΓX𝐦x=(g,\mathbf{v})\Gamma/\Gamma\in X_{\mathbf{m}}, then by definition of X𝐦X_{\mathbf{m}}, we have g1𝐯𝐦dg^{-1}\mathbf{v}\cdot\mathbf{m}\in\mathbb{Z}^{d}. Choose any 𝝃(d)k\bm{\xi}\in(\mathbb{R}^{d})^{k} such that g1𝐯𝐦=𝝃𝐦g^{-1}\mathbf{v}\cdot\mathbf{m}=\bm{\xi}\cdot\mathbf{m}, then (𝐯g𝝃)𝐦=g(g1𝐯𝐦𝝃𝐦)=𝟎(\mathbf{v}-g\bm{\xi})\cdot\mathbf{m}=g(g^{-1}\mathbf{v}\cdot\mathbf{m}-\bm{\xi}\cdot\mathbf{m})=\mathbf{0}. Therefore, α𝐦(x)=.\alpha_{\mathbf{m}}(x)=\infty.

Notations. Let’s fix some simplified notations for the rest of the proof. We will fix an 𝐦k{𝟎}\mathbf{m}\in\mathbb{Z}^{k}\setminus\{\mathbf{0}\} till the end of this section. In the following, t>0t>0 is a sufficiently large number.

  • For any nn\in\mathbb{N}, any 𝐬𝒰\mathbf{s}\in\mathcal{U}, denote antu𝝋(𝐬)=(gn(𝐬),𝐯n(𝐬))a_{nt}u_{\bm{\varphi}}(\mathbf{s})=(g_{n}(\mathbf{s}),\mathbf{v}_{n}(\mathbf{s})).

  • If 𝝃gn(𝐬),𝐯n(𝐬)\bm{\xi}_{g_{n}(\mathbf{s}),\mathbf{v}_{n}(\mathbf{s})} exists, denote 𝝃gn(𝐬),𝐯n(𝐬)=𝝃n,𝐬\bm{\xi}_{g_{n}(\mathbf{s}),\mathbf{v}_{n}(\mathbf{s})}=\bm{\xi}_{n,\mathbf{s}}.

  • For any 𝐯(d)𝐦k\mathbf{v}\in(\mathbb{R}^{d})^{k}_{\mathbf{m}}, nn\in\mathbb{N}, 𝐬𝒰\mathbf{s}\in\mathcal{U}, let

    w(n,𝐬,𝐯)\displaystyle w(n,\mathbf{s},\mathbf{v}) =(𝐯n(𝐬)gn(𝐬)𝐯)𝐦\displaystyle=(\mathbf{v}_{n}(\mathbf{s})-g_{n}(\mathbf{s})\mathbf{v})\cdot\mathbf{m}
    =[e(dr)nt[(𝝋(𝐬))r(𝐯)r𝐬(𝐯)>r]𝐦ernt(𝐯)>r𝐦]=[w1(n,𝐬,𝐯)wd(n,𝐬,𝐯)].\displaystyle=\begin{bmatrix}e^{(d-r)nt}[(\bm{\varphi}(\mathbf{s}))_{\leq r}-(\mathbf{v})_{\leq r}-\mathbf{s}\cdot(\mathbf{v})_{>r}]\cdot\mathbf{m}\\ e^{-rnt}(\mathbf{v})_{>r}\cdot\mathbf{m}\end{bmatrix}=\begin{bmatrix}w_{1}(n,\mathbf{s},\mathbf{v})\\ \vdots\\ w_{d}(n,\mathbf{s},\mathbf{v})\end{bmatrix}.

    We note that if 𝝃n,𝐬\bm{\xi}_{n,\mathbf{s}} exists, then α𝐦(antu𝝋(𝐬))=w(n,𝐬,𝝃n,𝐬)1\alpha_{\mathbf{m}}(a_{nt}u_{\bm{\varphi}}(\mathbf{s}))=\left\lVert w(n,\mathbf{s},\bm{\xi}_{n,\mathbf{s}})\right\rVert^{-1}.

  • For any differentiable function ψ:Matr×(dr)()\psi:Mat_{r\times(d-r)}(\mathbb{R})\to\mathbb{R}, by mean value theorem in several variables, for any 𝐬,𝐬~Matr×(dr)()\mathbf{s},\tilde{\mathbf{s}}\in Mat_{r\times(d-r)}(\mathbb{R}), there is a 𝐬^\hat{\mathbf{s}} such that

    ψ(𝐬~)ψ(𝐬)=i=1rj=1drψsij(𝐬^)(s~ijsij).\displaystyle\psi(\tilde{\mathbf{s}})-\psi(\mathbf{s})=\sum_{i=1}^{r}\sum_{j=1}^{d-r}\frac{\partial\psi}{\partial s_{ij}}(\hat{\mathbf{s}})\cdot(\tilde{s}_{ij}-s_{ij}).

    Since the functions that we consider have bounded first derivative on a bounded set, we will omit this 𝐬^\hat{\mathbf{s}} for simplicity.

  • For 1idr1\leq i\leq d-r, let 𝐞i\mathbf{e}_{i} denote the column vector in dr\mathbb{R}^{d-r} with 11 in ii-th row and 0 elsewhere. Let <,><,> denote the usual inner product of column vectors.

Lemma 6.8.

Let n1n\geq 1 be an integer and t>log(2μd1σ1)t>log(2\mu_{d}^{-1}\sigma^{-1}), where μd>0\mu_{d}>0 is the constant given as in Remark 6.6. For any 𝐬I\mathbf{s}\in I^{\prime}, where II^{\prime} is given as in Remark 6.3, if 𝛏n,𝐬\bm{\xi}_{n,\mathbf{s}} exists, then (𝛏n,𝐬)>r𝐦>M1𝐦\left\lVert(\bm{\xi}_{n,\mathbf{s}})_{>r}\cdot\mathbf{m}\right\rVert_{\infty}>M_{1}\left\lVert\mathbf{m}\right\rVert.

Proof.

Suppose (𝝃n,𝐬)>r𝐦M1𝐦\left\lVert(\bm{\xi}_{n,\mathbf{s}})_{>r}\cdot\mathbf{m}\right\rVert_{\infty}\leq M_{1}\left\lVert\mathbf{m}\right\rVert. By definition of II^{\prime},

(𝐯n(𝐬)gn(𝐬)𝝃n,𝐬)𝐦\displaystyle\left\lVert(\mathbf{v}_{n}(\mathbf{s})-g_{n}(\mathbf{s})\bm{\xi}_{n,\mathbf{s}})\cdot\mathbf{m}\right\rVert (𝐯n(𝐬)gn(𝐬)𝝃n,𝐬)r𝐦\displaystyle\geq\left\lVert(\mathbf{v}_{n}(\mathbf{s})-g_{n}(\mathbf{s})\bm{\xi}_{n,\mathbf{s}})_{\leq r}\cdot\mathbf{m}\right\rVert_{\infty}
e(dr)nt[(𝝋(𝐬))r(𝝃n,𝐬)r𝐬(𝝃n,𝐬)>r]𝐦\displaystyle\geq e^{(d-r)nt}\left\lVert[(\bm{\varphi}(\mathbf{s}))_{\leq r}-(\bm{\xi}_{n,\mathbf{s}})_{\leq r}-\mathbf{s}(\bm{\xi}_{n,\mathbf{s}})_{>r}]\cdot\mathbf{m}\right\rVert_{\infty}
etσ2>μd1.\displaystyle\geq e^{t}\frac{\sigma}{2}>\mu_{d}^{-1}.

By Remark 6.6, this contradicts the existence of 𝝃n,𝐬\bm{\xi}_{n,\mathbf{s}}. ∎

Proof of Proposition 6.1 (3).

If n=0n=0, by Remark 6.6, we have μdνα𝐦ν(u𝝋(𝐬~))σν{\mu_{d}^{\nu}}\leq\alpha_{\mathbf{m}}^{\nu}(u_{\bm{\varphi}}(\tilde{\mathbf{s}}))\leq\sigma^{-\nu}. Since α~\tilde{\alpha} is continuous and bounded on compact sets, there exists 0<m<M0<m<M such that for any 𝐬I\mathbf{s}\in I,

mα~(u𝝋(𝐬)Γ)M.\displaystyle m\leq\tilde{\alpha}(u_{\bm{\varphi}}(\mathbf{s})\Gamma)\leq M.

Therefore,

β𝐦(u𝝋(𝐬~)Γ)σν+ceνrtMσν+ceνrtMμdν+ceνrtmβ𝐦(u𝝋(𝐬)Γ).\displaystyle\beta_{\mathbf{m}}(u_{\bm{\varphi}}(\tilde{\mathbf{s}})\Gamma)\leq\sigma^{-\nu}+ce^{\nu rt}\cdot M\leq{\frac{\sigma^{-\nu}+ce^{\nu rt}M}{\mu_{d}^{\nu}+ce^{\nu rt}m}\beta_{\mathbf{m}}(u_{\bm{\varphi}}(\mathbf{s})\Gamma).}

Assume n1n\geq 1. If 𝝃n,𝐬~\bm{\xi}_{n,\tilde{\mathbf{s}}} does not exist, then by definition, α𝐦ν(antu𝝋(𝐬~)Γ)=1\alpha_{\mathbf{m}}^{\nu}(a_{nt}u_{\bm{\varphi}}(\tilde{\mathbf{s}})\Gamma)=1. Now we suppose that 𝝃n,𝐬~\bm{\xi}_{n,\tilde{\mathbf{s}}} exists.

Case 1: (w(n,𝐬,𝝃n,𝐬~))rN2(w(n,𝐬,𝝃n,𝐬~))>r\left\lVert(w(n,\mathbf{s},\bm{\xi}_{n,\tilde{\mathbf{s}}}))_{\leq r}\right\rVert_{\infty}\geq N_{2}\left\lVert(w(n,\mathbf{s},\bm{\xi}_{n,\tilde{\mathbf{s}}}))_{>r}\right\rVert_{\infty}, where N2=2(r+1)(dr)N_{2}=2(r+1)(d-r).

Choose t>log(2μd1σ1)t>log(2\mu_{d}^{-1}\sigma^{-1}), by Lemma 6.8, the Lipschity continuity of 𝝋\bm{\varphi}, and the choices of M1M_{1} and JJ, we have

(w(n,𝐬~,𝝃n,𝐬~))r(w(n,𝐬,𝝃n,𝐬~))r\displaystyle\left\lVert(w(n,\tilde{\mathbf{s}},\bm{\xi}_{n,\tilde{\mathbf{s}}}))_{\leq r}-(w(n,\mathbf{s},\bm{\xi}_{n,\tilde{\mathbf{s}}}))_{\leq r}\right\rVert_{\infty} 2ernt(r+1)(dr)(𝝃n,𝐬~)>r𝐦.\displaystyle\leq 2e^{-rnt}(r+1)(d-r)\left\lVert(\bm{\xi}_{n,\tilde{\mathbf{s}}})_{>r}\cdot\mathbf{m}\right\rVert_{\infty}.
=2(r+1)(dr)(w(n,𝐬~,𝝃n,𝐬~))>r.\displaystyle=2(r+1)(d-r)\left\lVert(w(n,\tilde{\mathbf{s}},\bm{\xi}_{n,\tilde{\mathbf{s}}}))_{>r}\right\rVert_{\infty}.

Hence, by the assumption of Case 1,

w(n,𝐬~,𝝃n,𝐬~)\displaystyle\left\lVert w(n,\tilde{\mathbf{s}},\bm{\xi}_{n,\tilde{\mathbf{s}}})\right\rVert (w(n,𝐬~,𝝃n,𝐬~))r\displaystyle\geq\left\lVert(w(n,\tilde{\mathbf{s}},\bm{\xi}_{n,\tilde{\mathbf{s}}}))_{\leq r}\right\rVert_{\infty}
(w(n,𝐬,𝝃n,𝐬~))r2(r+1)(dr)(w(n,𝐬,𝝃n,𝐬~))>r\displaystyle\geq\left\lVert(w(n,\mathbf{s},\bm{\xi}_{n,\tilde{\mathbf{s}}}))_{\leq r}\right\rVert_{\infty}-2(r+1)(d-r)\left\lVert(w(n,\mathbf{s},\bm{\xi}_{n,\tilde{\mathbf{s}}}))_{>r}\right\rVert_{\infty}
1d2w(n,𝐬,𝝃n,𝐬~).\displaystyle\geq{\frac{1}{d^{2}}\left\lVert w(n,\mathbf{s},\bm{\xi}_{n,\tilde{\mathbf{s}}})\right\rVert.}

Case 2: (w(n,𝐬,𝝃n,𝐬~))r<N2(w(n,𝐬,𝝃n,𝐬~))>r\left\lVert(w(n,\mathbf{s},\bm{\xi}_{n,\tilde{\mathbf{s}}}))_{\leq r}\right\rVert_{\infty}<N_{2}\left\lVert(w(n,\mathbf{s},\bm{\xi}_{n,\tilde{\mathbf{s}}}))_{>r}\right\rVert_{\infty}. Then

w(n,𝐬~,𝝃n,𝐬~)\displaystyle\left\lVert w(n,\tilde{\mathbf{s}},\bm{\xi}_{n,\tilde{\mathbf{s}}})\right\rVert (w(n,𝐬~,𝝃n,𝐬~))>r=(w(n,𝐬,𝝃n,𝐬~))>r\displaystyle\geq\left\lVert(w(n,\tilde{\mathbf{s}},\bm{\xi}_{n,\tilde{\mathbf{s}}}))_{>r}\right\rVert_{\infty}=\left\lVert(w(n,\mathbf{s},\bm{\xi}_{n,\tilde{\mathbf{s}}}))_{>r}\right\rVert_{\infty}
1rN22+drw(n,𝐬,𝝃n,𝐬~)15dr2w(n,𝐬,𝝃n,𝐬~).\displaystyle\geq\frac{1}{\sqrt{rN_{2}^{2}+d-r}}\left\lVert w(n,\mathbf{s},\bm{\xi}_{n,\tilde{\mathbf{s}}})\right\rVert\geq\frac{1}{5dr^{2}}\left\lVert w(n,\mathbf{s},\bm{\xi}_{n,\tilde{\mathbf{s}}})\right\rVert.

By construction of 𝝃n,𝐬\bm{\xi}_{n,\mathbf{s}}, we have w(n,𝐬,𝝃n,𝐬~)w(n,𝐬,𝝃n,𝐬)\left\lVert w(n,\mathbf{s},\bm{\xi}_{n,\tilde{\mathbf{s}}})\right\rVert\geq\left\lVert w(n,\mathbf{s},\bm{\xi}_{n,\mathbf{s}})\right\rVert. Combining Case 1 and Case 2, we have

w(n,𝐬~,𝝃n,𝐬~)min{15dr2,1d2}w(n,𝐬,𝝃n,𝐬).\displaystyle\left\lVert w(n,\tilde{\mathbf{s}},\bm{\xi}_{n,\tilde{\mathbf{s}}})\right\rVert\geq\min\{\frac{1}{5dr^{2}},\frac{1}{d^{2}}\}\left\lVert w(n,\mathbf{s},\bm{\xi}_{n,\mathbf{s}})\right\rVert.

Therefore,

α𝐦ν(antu𝝋(𝐬~)Γ)max(d2ν,(5dr2)ν)α𝐦ν(antu𝝋(𝐬)Γ).\displaystyle\alpha_{\mathbf{m}}^{\nu}(a_{nt}u_{\bm{\varphi}}(\tilde{\mathbf{s}})\Gamma)\leq\max(d^{2\nu},(5dr^{2})^{\nu})\cdot\alpha_{\mathbf{m}}^{\nu}(a_{nt}u_{\bm{\varphi}}(\mathbf{s})\Gamma).

For α~\tilde{\alpha}, by Remark 5.2, we have for large enough M¯>0\overline{M}>0,

α~(antu𝝋(𝐬~)Γ)=α~(u(ednt(𝐬~𝐬))antu𝝋(𝐬)Γ)M¯α~(antu𝝋(𝐬)Γ).\displaystyle\tilde{\alpha}(a_{nt}u_{\bm{\varphi}}(\tilde{\mathbf{s}})\Gamma)=\tilde{\alpha}(u(e^{dnt}(\tilde{\mathbf{s}}-\mathbf{s}))a_{nt}u_{\bm{\varphi}}(\mathbf{s})\Gamma)\leq\overline{M}\tilde{\alpha}(a_{nt}u_{\bm{\varphi}}(\mathbf{s})\Gamma).

By the above, let M~1=2max{M¯,d2ν,(5dr2)ν,σν+ceνrtMμdν+ceνrtm}\tilde{M}_{1}=2\max\{\overline{M},d^{2\nu},(5dr^{2})^{\nu},\frac{\sigma^{-\nu}+ce^{\nu rt}M}{\mu_{d}^{\nu}+ce^{\nu rt}m}\}, then

β𝐦(antu𝝋(𝐬~)Γ)M~1β𝐦(antu𝝋(𝐬)Γ),n0.\displaystyle\beta_{\mathbf{m}}(a_{nt}u_{\bm{\varphi}}(\tilde{\mathbf{s}})\Gamma)\leq\tilde{M}_{1}\beta_{\mathbf{m}}(a_{nt}u_{\bm{\varphi}}(\mathbf{s})\Gamma),\forall n\in\mathbb{Z}_{\geq 0}.

Proof of Proposition 6.1 (4).

By Remark 5.2, since |τ|t|\tau|\leq t, we have

α~(aτantu𝝋(𝐬)Γ)er(dr)νtα~(antu𝝋(𝐬)Γ).\displaystyle\tilde{\alpha}(a_{\tau}a_{nt}u_{\bm{\varphi}}(\mathbf{s})\Gamma)\leq e^{r(d-r)\nu t}\tilde{\alpha}(a_{nt}u_{\bm{\varphi}}(\mathbf{s})\Gamma).

Let aτantu𝝋(𝐬)=(aτgn(𝐬),aτ𝐯n(𝐬))a_{\tau}a_{nt}u_{\bm{\varphi}}(\mathbf{s})=(a_{\tau}g_{n}(\mathbf{s}),a_{\tau}\mathbf{v}_{n}(\mathbf{s})).

If 𝝃aτgn(𝐬),aτ𝐯n(𝐬)\bm{\xi}_{a_{\tau}g_{n}(\mathbf{s}),a_{\tau}\mathbf{v}_{n}(\mathbf{s})} does not exist, then α𝐦(aτantu𝝋(𝐬)Γ)=1\alpha_{\mathbf{m}}(a_{\tau}a_{nt}u_{\bm{\varphi}}(\mathbf{s})\Gamma)=1.

If 𝝃aτgn(𝐬),aτ𝐯n(𝐬)=𝐯\bm{\xi}_{a_{\tau}g_{n}(\mathbf{s}),a_{\tau}\mathbf{v}_{n}(\mathbf{s})}=\mathbf{v} exists, then

α𝐦ν(aτantu𝝋(𝐬)Γ)\displaystyle\alpha_{\mathbf{m}}^{\nu}(a_{\tau}a_{nt}u_{\bm{\varphi}}(\mathbf{s})\Gamma) =(aτ(𝐯n(𝐬)gn(𝐬)𝐯)𝐦ν\displaystyle=\left\lVert(a_{\tau}(\mathbf{v}_{n}(\mathbf{s})-g_{n}(\mathbf{s})\mathbf{v})\cdot\mathbf{m}\right\rVert^{-\nu}
e(dr)νt(𝐯n(𝐬)gn(𝐬)𝐯)𝐦ν\displaystyle\leq e^{(d-r)\nu t}\left\lVert(\mathbf{v}_{n}(\mathbf{s})-g_{n}(\mathbf{s})\mathbf{v})\cdot\mathbf{m}\right\rVert^{-\nu}
{e(dr)νtα𝐦ν(antu𝝋(𝐬)Γ)If 𝐯=𝝃gn(𝐬),𝐯n(𝐬)e(dr)νt2να~(antu𝝋(𝐬)Γ)If 𝐯𝝃gn(𝐬),𝐯n(𝐬)\displaystyle\leq\begin{cases}e^{(d-r)\nu t}\alpha_{\mathbf{m}}^{\nu}(a_{nt}u_{\bm{\varphi}}(\mathbf{s})\Gamma)&\text{If }\mathbf{v}=\bm{\xi}_{g_{n}(\mathbf{s}),\mathbf{v}_{n}(\mathbf{s})}\\ e^{(d-r)\nu t}2^{\nu}\tilde{\alpha}(a_{nt}u_{\bm{\varphi}}(\mathbf{s})\Gamma)&\text{If }\mathbf{v}\neq\bm{\xi}_{g_{n}(\mathbf{s}),\mathbf{v}_{n}(\mathbf{s})}\end{cases}

Let

M~2=max(e(dr)νt,2νe(dr)νt+cer(dr)νteνrtceνrt),\displaystyle\tilde{M}_{2}=\max(e^{(d-r)\nu t},\frac{2^{\nu}e^{(d-r)\nu t}+c\cdot e^{r(d-r)\nu t}e^{\nu rt}}{c\cdot e^{\nu rt}}),

then by the above estimates, Proposition 6.1 (4) is proved. ∎

To prove property (5) of Proposition 6.1, we record the following lemmas:

Lemma 6.9.

[25, Lemma 4.8] Let n0n\in\mathbb{Z}_{\geq 0} and t>0t>0. Let I0=[1,1]r(dr)I_{0}=[-1,1]^{r(d-r)}, J=i=1r(dr)JiJ=\prod_{i=1}^{r(d-r)}J_{i}, where JiJ_{i} is an interval with |Ji|ednt|J_{i}|\geq e^{-dnt} for each ii. Let Ψ:G/Γ+\Psi:G/\Gamma\to\mathbb{R}_{+} be a measurable function. Then

JΨ(a(n+1)tu𝝋(𝐬)Γ)𝑑𝐬JI0Ψ(a(n+1)tu𝝋(𝐬+𝐬~ednt)Γ)𝑑𝐬~𝑑𝐬.\displaystyle\int_{J}\Psi(a_{(n+1)t}u_{\bm{\varphi}}(\mathbf{s})\Gamma)d\mathbf{s}\leq\int_{J}\int_{I_{0}}\Psi(a_{(n+1)t}u_{\bm{\varphi}}(\mathbf{s}+\tilde{\mathbf{s}}e^{-dnt})\Gamma)d\tilde{\mathbf{s}}d\mathbf{s}. (6.5)
Lemma 6.10.

Let κ:I0=[1,1]r(dr)+\kappa:I_{0}=[-1,1]^{r(d-r)}\to\mathbb{R}_{+} be a measurable function. Suppose that there exists C>0C>0 such that for any ϵ>0\epsilon>0,

|{𝐬I0:κ(𝐬)<ϵ}|Cϵ1r(dr).\displaystyle|\{\mathbf{s}\in I_{0}:\kappa(\mathbf{s})<\epsilon\}|\leq C\cdot\epsilon^{\frac{1}{r(d-r)}}.

Then for any 0<ν<1r(dr)0<\nu<\frac{1}{r(d-r)}, there exists cν>0c_{\nu}>0 such that

I0κ(𝐬)ν𝑑𝐬Cνr(dr)cν.\displaystyle\int_{I_{0}}\kappa(\mathbf{s})^{-\nu}d\mathbf{s}\leq C^{\nu\cdot r(d-r)}\cdot c_{\nu}.
Proof.

This is a direct generalization of [8, Lemma 6.10]. ∎

The following lemma is a special case of [11, Lemma 3.3].

Lemma 6.11.

Let VV be a bounded open subset of Matr×(dr)()Mat_{r\times(d-r)}(\mathbb{R}), and let fC1(V)f\in C^{1}(V) be such that for some constants A1,A2>0A_{1},A_{2}>0, one has

A2|ijf(𝐬)|A1,1ir,1jdr,𝐬V,\displaystyle A_{2}\leq|\partial_{ij}f(\mathbf{s})|\leq A_{1},\forall 1\leq i\leq r,1\leq j\leq d-r,\forall\mathbf{s}\in V,
and fVA1,\displaystyle\text{ and }\left\lVert f\right\rVert_{V}\leq A_{1},

where V\left\lVert\cdot\right\rVert_{V} denote the sup norm of a function on VV. Then for any box (or ball) BVB\subset V, any ϵ>0\epsilon>0, one has

|{𝐬B:|f(𝐬)|<ϵ}|r(dr)CA1,A2(ϵfB)1r(dr)|B|,\displaystyle|\{\mathbf{s}\in B:|f(\mathbf{s})|<\epsilon\}|\leq r(d-r)\cdot C_{A_{1},A_{2}}(\frac{\epsilon}{\left\lVert f\right\rVert_{B}})^{\frac{1}{r(d-r)}}|B|,

with CA1,A2=12A1A2C_{A_{1},A_{2}}=\frac{12A_{1}}{A_{2}}.

Lemma 6.12.

Let N1N\geq 1 be an integer. Let f:Nf:\mathbb{R}^{N}\to\mathbb{R} be a C1C^{1} map. Given 𝐁SON()\mathbf{B}\in SO_{N}(\mathbb{R}), for any 𝐬N\mathbf{s}\in\mathbb{R}^{N}, let 𝐬=𝐁𝐬\mathbf{s}^{\prime}=\mathbf{B}\mathbf{s}. Then

(fs1,,fsN)t=𝐁(fs1,,fsN)t,\displaystyle(\frac{\partial f}{\partial s_{1}^{\prime}},\cdots,\frac{\partial f}{\partial s_{N}^{\prime}})^{t}=\mathbf{B}(\frac{\partial f}{\partial s_{1}},\cdots,\frac{\partial f}{\partial s_{N}})^{t},

where superscript tt denotes the transpose of the vector.

Proof.

Since 𝐬=𝐁𝐬\mathbf{s}^{\prime}=\mathbf{B}\mathbf{s}, 𝐬=𝐁1𝐬\mathbf{s}=\mathbf{B}^{-1}\mathbf{s}^{\prime}. Using chain rule, it can be verified that

(fs1,,fsN)t=(𝐁1)t(fs1,,fsN)t.\displaystyle(\frac{\partial f}{\partial s_{1}^{\prime}},\cdots,\frac{\partial f}{\partial s_{N}^{\prime}})^{t}=(\mathbf{B}^{-1})^{t}\cdot(\frac{\partial f}{\partial s_{1}},\cdots,\frac{\partial f}{\partial s_{N}})^{t}.

As 𝐁SON()\mathbf{B}\in SO_{N}(\mathbb{R}), we have (𝐁1)t=𝐁(\mathbf{B}^{-1})^{t}=\mathbf{B}. This proves the lemma. ∎

Lemma 6.13.

Let N1N\geq 1 be an integer. Given real numbers 0<c2<C2<c1<C10<c_{2}<C_{2}<c_{1}<C_{1}, and a partition {1,2,3}\{\mathcal{I}_{1},\mathcal{I}_{2},\mathcal{I}_{3}\} of {1,,N}\{1,\cdots,N\} such that 1\mathcal{I}_{1}\neq\emptyset. There is 𝐁SON()\mathbf{B}\in SO_{N}(\mathbb{R}) (depending only on the partition) such that the following holds: For any vector 𝐯=(v1,,vN)tN\mathbf{v}=(v_{1},\cdots,v_{N})^{t}\in\mathbb{R}^{N} (here superscript tt denote the transpose of the corresponding vector) satisfying

  • For any i{1,,N}i\in\{1,\cdots,N\}, |vi|C1|v_{i}|\leq C_{1};

  • For any i1i\in\mathcal{I}_{1}, |vi|c1|v_{i}|\geq c_{1};

  • For any i2i\in\mathcal{I}_{2}, |vi|C2|v_{i}|\geq C_{2};

  • For any i3i\in\mathcal{I}_{3}, |vi|c2|v_{i}|\leq c_{2}.

If we denote 𝐯=(v1,,vN)t=𝐁𝐯\mathbf{v}^{\prime}=(v_{1}^{\prime},\cdots,v_{N}^{\prime})^{t}=\mathbf{B}\mathbf{v}, then for any i=1,,Ni=1,\cdots,N,

min{c1NNc2,C2}|vi|C1+Nc2.\displaystyle\min\{\frac{c_{1}}{\sqrt{N}}-\sqrt{N}c_{2},C_{2}\}\leq|v_{i}^{\prime}|\leq C_{1}+\sqrt{N}c_{2}.
Proof.

If 3=\mathcal{I}_{3}=\emptyset, then the lemma is trivial. Now we assume that 3\mathcal{I}_{3}\neq\emptyset. Let pp\in\mathbb{N} be such that p1=#3p-1=\#\mathcal{I}_{3}, then 2pN2\leq p\leq N. Without loss of generality, we may assume that 111\in\mathcal{I}_{1} and 3={2,,p}\mathcal{I}_{3}=\{2,\cdots,p\}. Choose a 𝐁=(bij)SON()\mathbf{B}=(b_{ij})\in SO_{N}(\mathbb{R}) satisfying

  • bi1=1pb_{i1}=\frac{1}{\sqrt{p}}, for 1ip1\leq i\leq p;

  • bij=0b_{ij}=0 if jp<ij\leq p<i, or ip<ji\leq p<j;

  • bij=δijb_{ij}=\delta_{ij} if ip+1i\geq p+1 and jp+1j\geq p+1.

Note that for any i=1,,pi=1,\cdots,p, vi=1/pv1+j=2pbijvjv_{i}^{\prime}=1/\sqrt{p}\cdot v_{1}+\sum_{j=2}^{p}b_{ij}v_{j}. Therefore, for any 1ip1\leq i\leq p, we have the lower bound

|vi|1p|v1|j=2p|bij||vj|c1p(j=2p|c2|2)12c1NNc2,\displaystyle|v_{i}^{\prime}|\geq\frac{1}{\sqrt{p}}|v_{1}|-\sum_{j=2}^{p}|b_{ij}||v_{j}|\geq\frac{c_{1}}{\sqrt{p}}-(\sum_{j=2}^{p}|c_{2}|^{2})^{\frac{1}{2}}\geq\frac{c_{1}}{\sqrt{N}}-\sqrt{N}c_{2},

where in the second inequality we apply Cauchy-Schwartz inequality. Also for any 1ip1\leq i\leq p, we have the upper bound

|vi|1p|v1|+j=2p|bij||vj|C1+Nc2.\displaystyle|v_{i}^{\prime}|\leq\frac{1}{\sqrt{p}}|v_{1}|+\sum_{j=2}^{p}|b_{ij}||v_{j}|\leq C_{1}+\sqrt{N}c_{2}.

On the other hand, for any p+1iNp+1\leq i\leq N, we have vi=viv_{i}^{\prime}=v_{i}. Therefore, for any 1iN1\leq i\leq N,

min{c1NNc2,C2}|vi|C1+Nc2.\displaystyle\min\{\frac{c_{1}}{\sqrt{N}}-\sqrt{N}c_{2},C_{2}\}\leq|v_{i}^{\prime}|\leq C_{1}+\sqrt{N}c_{2}.

Roughly speaking, Lemma 6.13 says that one can find a suitable rotation 𝐁SON()\mathbf{B}\in SO_{N}(\mathbb{R}) depending only on the partition of {1,,N}\{1,\cdots,N\} such that for any vector 𝐯N\mathbf{v}\in\mathbb{R}^{N}, as long as there is a coordinate of 𝐯\mathbf{v} with large enough absolute value, the absolute value of all coordinates of the new vector 𝐁𝐯\mathbf{B}\mathbf{v} are bounded below by a suitable constant.

Proof of (5) of Proposition 6.1.

If n=0n=0 and J=IJ=I. Then for any 𝐬I\mathbf{s}\in I, by Proposition 6.1 (4),

β𝐦(atu𝝋(𝐬)Γ)M~2β𝐦(u𝝋(𝐬)Γ)M~2(σν+M).\displaystyle\beta_{\mathbf{m}}(a_{t}u_{\bm{\varphi}}(\mathbf{s})\Gamma)\leq\tilde{M}_{2}\beta_{\mathbf{m}}(u_{\bm{\varphi}}(\mathbf{s})\Gamma)\leq\tilde{M}_{2}(\sigma^{-\nu}+M).

Then for any bb\in\mathbb{R} such that b>M~2(σν+M)b>\tilde{M}_{2}(\sigma^{-\nu}+M),

Iβ𝐦(atu𝝋(𝐬)Γ)𝑑𝐬12Iβ𝐦(u𝝋(𝐬)Γ)𝑑𝐬+b|I|.\displaystyle\int_{I}\beta_{\mathbf{m}}(a_{t}u_{\bm{\varphi}}(\mathbf{s})\Gamma)d\mathbf{s}\leq\frac{1}{2}\int_{I}\beta_{\mathbf{m}}(u_{\bm{\varphi}}(\mathbf{s})\Gamma)d\mathbf{s}+b|I|.

Now we assume n1n\geq 1 and let t>0t>0 be a sufficiently large number (to be specified later). By Lemma 6.9,

Jβ𝐦(a(n+1)tu𝝋(𝐬)Γ)𝑑𝐬JI0β𝐦(a(n+1)tu𝝋(𝐬+𝐬~ednt)Γ)𝑑𝐬~𝑑𝐬.\displaystyle\int_{J}\beta_{\mathbf{m}}(a_{(n+1)t}u_{\bm{\varphi}}(\mathbf{s})\Gamma)d\mathbf{s}\leq\int_{J}\int_{I_{0}}\beta_{\mathbf{m}}(a_{(n+1)t}u_{\bm{\varphi}}(\mathbf{s}+\tilde{\mathbf{s}}e^{-dnt})\Gamma)d\tilde{\mathbf{s}}d\mathbf{s}.

By Proposition 5.1, for t>0t>0 sufficiently large, there exists b1>0b_{1}>0 such that for any 𝐬J\mathbf{s}\in J,

I0α~(atu(𝐬~)antu(𝐬)Γ)𝑑𝐬~14α~(antu𝝋(𝐬)Γ)+b1.\displaystyle\int_{I_{0}}\tilde{\alpha}(a_{t}u(\tilde{\mathbf{s}})a_{nt}u(\mathbf{s})\Gamma)d\tilde{\mathbf{s}}\leq\frac{1}{4}\tilde{\alpha}(a_{nt}u_{\bm{\varphi}}(\mathbf{s})\Gamma)+b_{1}. (6.6)

Note that since

I0α~(atu(𝐬~)antu(𝐬)Γ)𝑑𝐬~\displaystyle\int_{I_{0}}\tilde{\alpha}(a_{t}u(\tilde{\mathbf{s}})a_{nt}u(\mathbf{s})\Gamma)d\tilde{\mathbf{s}} =I0α~(a(n+1)tu𝝋(𝐬+𝐬~ednt)Γ)𝑑𝐬~\displaystyle=\int_{I_{0}}\tilde{\alpha}(a_{(n+1)t}u_{\bm{\varphi}}(\mathbf{s}+\tilde{\mathbf{s}}e^{-dnt})\Gamma)d\tilde{\mathbf{s}}
14α~(antu𝝋(𝐬)Γ)+b1,\displaystyle\leq\frac{1}{4}\tilde{\alpha}(a_{nt}u_{\bm{\varphi}}(\mathbf{s})\Gamma)+b_{1},

by definition of β𝐦\beta_{\mathbf{m}}, it remains to estimate the following integral for any 𝐬J\mathbf{s}\in J:

I0α𝐦ν(a(n+1)tu𝝋(𝐬+𝐬~ednt)Γ)𝑑𝐬~.\displaystyle\int_{I_{0}}\alpha_{\mathbf{m}}^{\nu}(a_{(n+1)t}u_{\bm{\varphi}}(\mathbf{s}+\tilde{\mathbf{s}}e^{-dnt})\Gamma)d\tilde{\mathbf{s}}. (6.7)

Define for any 𝐬J\mathbf{s}\in J,

I01(𝐬):={𝐬~I0:𝐬^=𝐬+𝐬~ednt,𝝃n+1,𝐬^ exists,𝝃n+1,𝐬^𝝃n,𝐬},\displaystyle I_{01}(\mathbf{s}):=\{\tilde{\mathbf{s}}\in I_{0}:\hat{\mathbf{s}}=\mathbf{s}+\tilde{\mathbf{s}}e^{-dnt},\bm{\xi}_{n+1,\hat{\mathbf{s}}}\text{ exists},\bm{\xi}_{n+1,\hat{\mathbf{s}}}\neq\bm{\xi}_{n,\mathbf{s}}\},
I02(𝐬):={𝐬~I0:𝐬^=𝐬+𝐬~ednt,𝝃n+1,𝐬^ exists,𝝃n+1,𝐬^=𝝃n,𝐬},\displaystyle I_{02}(\mathbf{s}):=\{\tilde{\mathbf{s}}\in I_{0}:\hat{\mathbf{s}}=\mathbf{s}+\tilde{\mathbf{s}}e^{-dnt},\bm{\xi}_{n+1,\hat{\mathbf{s}}}\text{ exists},\bm{\xi}_{n+1,\hat{\mathbf{s}}}=\bm{\xi}_{n,\mathbf{s}}\},
I03(𝐬):={𝐬~I0:𝐬^=𝐬+𝐬~ednt,𝝃n+1,𝐬^ does not exist}.\displaystyle I_{03}(\mathbf{s}):=\{\tilde{\mathbf{s}}\in I_{0}:\hat{\mathbf{s}}=\mathbf{s}+\tilde{\mathbf{s}}e^{-dnt},\bm{\xi}_{n+1,\hat{\mathbf{s}}}\text{ does not exist}\}.

Since for 𝐬~I03\tilde{\mathbf{s}}\in I_{03}, α𝐦ν\alpha_{\mathbf{m}}^{\nu} is dominated by α~\tilde{\alpha}, by Lemmas 6.14, 6.15 given as follows, we have

I0α𝐦ν(a(n+1)tu𝝋(𝐬+𝐬~ednt)Γ)𝑑𝐬~=I01(𝐬)α𝐦ν𝑑𝐬~+I02(𝐬)α𝐦ν𝑑𝐬~+I03(𝐬)α𝐦ν𝑑𝐬~\displaystyle\int_{I_{0}}\alpha_{\mathbf{m}}^{\nu}(a_{(n+1)t}u_{\bm{\varphi}}(\mathbf{s}+\tilde{\mathbf{s}}e^{-dnt})\Gamma)d\tilde{\mathbf{s}}=\int_{I_{01}(\mathbf{s})}\alpha_{\mathbf{m}}^{\nu}d\tilde{\mathbf{s}}+\int_{I_{02}(\mathbf{s})}\alpha_{\mathbf{m}}^{\nu}d\tilde{\mathbf{s}}+\int_{I_{03}(\mathbf{s})}\alpha_{\mathbf{m}}^{\nu}d\tilde{\mathbf{s}}
(10r2d)νeνrt2r(dr)α~(antu𝝋(𝐬)Γ)+14α𝐦ν(antu𝝋(𝐬)Γ)+2ν+r(dr).\displaystyle\leq(10r^{2}d)^{\nu}e^{\nu rt}\cdot 2^{r(d-r)}\cdot\tilde{\alpha}(a_{nt}u_{\bm{\varphi}}(\mathbf{s})\Gamma)+\frac{1}{4}\alpha_{\mathbf{m}}^{\nu}(a_{nt}u_{\bm{\varphi}}(\mathbf{s})\Gamma)+2^{\nu+r(d-r)}.

As we choose b>2ν+r(dr)+cb1erνtb>2^{\nu+r(d-r)}+cb_{1}e^{r\nu t}, recall that c=4(10r2d)ν2r(dr)c=4\cdot(10r^{2}d)^{\nu}\cdot 2^{r(d-r)}, we have

Jβ𝐦(a(n+1)tu𝝋(𝐬)Γ)𝑑𝐬JI0β𝐦(a(n+1)tu𝝋(𝐬+𝐬~ednt)Γ)𝑑𝐬~𝑑𝐬\displaystyle\int_{J}\beta_{\mathbf{m}}(a_{(n+1)t}u_{\bm{\varphi}}(\mathbf{s})\Gamma)d\mathbf{s}\leq\int_{J}\int_{I_{0}}\beta_{\mathbf{m}}(a_{(n+1)t}u_{\bm{\varphi}}(\mathbf{s}+\tilde{\mathbf{s}}e^{-dnt})\Gamma)d\tilde{\mathbf{s}}d\mathbf{s}
J[14cerνtα~(antu𝝋(𝐬)Γ)+14α𝐦ν(antu𝝋(𝐬)Γ)+2ν+r(dr)+cb1erνt]𝑑𝐬,\displaystyle\leq\int_{J}[\frac{1}{4}c\cdot e^{r\nu t}\tilde{\alpha}(a_{nt}u_{\bm{\varphi}}(\mathbf{s})\Gamma)+\frac{1}{4}\alpha_{\mathbf{m}}^{\nu}(a_{nt}u_{\bm{\varphi}}(\mathbf{s})\Gamma)+2^{\nu+r(d-r)}+cb_{1}e^{r\nu t}]d\mathbf{s},
12Jβ𝐦(antu𝝋(𝐬)Γ)𝑑𝐬+b|J|.\displaystyle\leq\frac{1}{2}\int_{J}\beta_{\mathbf{m}}(a_{nt}u_{\bm{\varphi}}(\mathbf{s})\Gamma)d\mathbf{s}+b|J|.

This finishes the proof of property (5) of Proposition 6.1, modulo Lemmas 6.14, 6.15.

Lemma 6.14.

Let JJ be the box as in Proposition 6.1 (5). There is t>0t>0 sufficiently large such that for any 𝐬J\mathbf{s}\in J,

I01(𝐬)α𝐦ν(a(n+1)tu𝝋(𝐬+𝐬~ednt)Γ)𝑑𝐬~eνrt(10r2d)ν2r(dr)α~(antu𝝋(𝐬)Γ).\displaystyle\int_{I_{01}(\mathbf{s})}\alpha_{\mathbf{m}}^{\nu}(a_{(n+1)t}u_{\bm{\varphi}}(\mathbf{s}+\tilde{\mathbf{s}}e^{-dnt})\Gamma)d\tilde{\mathbf{s}}\leq e^{\nu rt}(10r^{2}d)^{\nu}2^{r(d-r)}\cdot\tilde{\alpha}(a_{nt}u_{\bm{\varphi}}(\mathbf{s})\Gamma).
Proof.

We will prove that for t>0t>0 sufficiently large, for any 𝐬~I01(𝐬)\tilde{\mathbf{s}}\in I_{01}(\mathbf{s}),

α𝐦ν(a(n+1)tu𝝋(𝐬+𝐬~ednt)Γ)eνrt(10r2d)να~(antu𝝋(𝐬)Γ).\displaystyle\alpha_{\mathbf{m}}^{\nu}(a_{(n+1)t}u_{\bm{\varphi}}(\mathbf{s}+\tilde{\mathbf{s}}e^{-dnt})\Gamma)\leq e^{\nu rt}(10r^{2}d)^{\nu}\cdot\tilde{\alpha}(a_{nt}u_{\bm{\varphi}}(\mathbf{s})\Gamma).

For 𝐬~I0\tilde{\mathbf{s}}\in I_{0}, denote 𝐬^=𝐬+𝐬~ednt\hat{\mathbf{s}}=\mathbf{s}+\tilde{\mathbf{s}}e^{-dnt}.

Case 1: (w(n,𝐬,𝝃n+1,𝐬^))rN2(w(n,𝐬,𝝃n+1,𝐬^))>r\left\lVert(w(n,\mathbf{s},\bm{\xi}_{n+1,\hat{\mathbf{s}}}))_{\leq r}\right\rVert_{\infty}\geq N_{2}\left\lVert(w(n,\mathbf{s},\bm{\xi}_{n+1,\hat{\mathbf{s}}}))_{>r}\right\rVert_{\infty}, where N2=2(r+1)(dr)N_{2}=2(r+1)(d-r).

By definition of M1M_{1} (cf.(2.1)), the choice of N1N_{1} and Lipschity continuity of 𝝋\bm{\varphi}, we have for any i,j,p,qi,j,p,q,

|q=1k𝝋pqsijmq|𝐦(q=1k|𝝋pqsij|2)12𝐦k12M1N1(𝝃n+1,𝐬^)>r𝐦.\displaystyle|\sum_{q=1}^{k}\frac{\partial\bm{\varphi}_{pq}}{\partial s_{ij}}\cdot m_{q}|\leq\left\lVert\mathbf{m}\right\rVert\cdot(\sum_{q=1}^{k}|\frac{\partial\bm{\varphi}_{pq}}{\partial s_{ij}}|^{2})^{\frac{1}{2}}\leq\left\lVert\mathbf{m}\right\rVert k^{\frac{1}{2}}\frac{M_{1}}{N_{1}}\leq\left\lVert(\bm{\xi}_{n+1,\hat{\mathbf{s}}})_{>r}\cdot\mathbf{m}\right\rVert_{\infty}.

By Lemma 6.8, the choices of the sidelength of the box and N2N_{2},

w(n+1,𝐬^,𝝃n+1,𝐬^)(w(n+1,𝐬^,𝝃n+1,𝐬^)r\displaystyle\left\lVert w(n+1,\hat{\mathbf{s}},\bm{\xi}_{n+1,\hat{\mathbf{s}}})\right\rVert\geq\left\lVert(w(n+1,\hat{\mathbf{s}},\bm{\xi}_{n+1,\hat{\mathbf{s}}})_{\leq r}\right\rVert_{\infty}
=e(dr)(n+1)t[(𝝋(𝐬^))r(𝝃n+1,𝐬^)r𝐬^(𝝃n+1,𝐬^)>r]𝐦\displaystyle=e^{(d-r)(n+1)t}\left\lVert[(\bm{\varphi}(\bm{\hat{\mathbf{s}}}))_{\leq r}-(\bm{\xi}_{n+1,\hat{\mathbf{s}}})_{\leq r}-\hat{\mathbf{s}}\cdot(\bm{\xi}_{n+1,\hat{\mathbf{s}}})_{>r}]\cdot\mathbf{m}\right\rVert_{\infty}
e(dr)t(w(n,𝐬,𝝃n+1,𝐬^))r(1(r+1)(dr)N2)e(dr)t(w(n,𝐬,𝝃n+1,𝐬^))r\displaystyle\geq e^{(d-r)t}\left\lVert(w(n,\mathbf{s},\bm{\xi}_{n+1,\hat{\mathbf{s}}}))_{\leq r}\right\rVert_{\infty}-(1-\frac{(r+1)(d-r)}{N_{2}})e^{(d-r)t}\left\lVert(w(n,\mathbf{s},\bm{\xi}_{n+1,\hat{\mathbf{s}}}))_{\leq r}\right\rVert_{\infty}
e(dr)t12dw(n,𝐬,𝝃n+1,𝐬^).\displaystyle\geq e^{(d-r)t}\frac{1}{2d}\left\lVert w(n,\mathbf{s},\bm{\xi}_{n+1,\hat{\mathbf{s}}})\right\rVert.

Choose t>0t>0 large enough such that e(dr)t12d>1e^{(d-r)t}\frac{1}{2d}>1, we obtain

α𝐦ν(a(n+1)tu𝝋(𝐬+𝐬~ednt)Γ)\displaystyle\alpha_{\mathbf{m}}^{\nu}(a_{(n+1)t}u_{\bm{\varphi}}(\mathbf{s}+\tilde{\mathbf{s}}e^{-dnt})\Gamma) =w(n+1,𝐬^,𝝃n+1,𝐬^)ν\displaystyle=\left\lVert w(n+1,\hat{\mathbf{s}},\bm{\xi}_{n+1,\hat{\mathbf{s}}})\right\rVert^{-\nu}
eν(dr)t(2d)νw(n,𝐬,𝝃n+1,𝐬^)ν\displaystyle\leq e^{-\nu(d-r)t}(2d)^{\nu}\left\lVert w(n,\mathbf{s},\bm{\xi}_{n+1,\hat{\mathbf{s}}})\right\rVert^{-\nu}
eν(dr)t(2d)ν2νsup𝐰d{𝟎}antu(𝐬)𝐰ν\displaystyle\leq e^{-\nu(d-r)t}(2d)^{\nu}\cdot 2^{\nu}\sup_{\mathbf{w}\in\mathbb{Z}^{d}\setminus\{\mathbf{0}\}}\left\lVert a_{nt}u(\mathbf{s})\mathbf{w}\right\rVert^{-\nu}
α~(antu𝝋(𝐬)Γ).\displaystyle\leq\tilde{\alpha}(a_{nt}u_{\bm{\varphi}(\mathbf{s})}\Gamma).

Case 2: (w(n,𝐬,𝝃n+1,𝐬^))r<N2(w(n,𝐬,𝝃n+1,𝐬^))>r\left\lVert(w(n,\mathbf{s},\bm{\xi}_{n+1,\hat{\mathbf{s}}}))_{\leq r}\right\rVert_{\infty}<N_{2}\left\lVert(w(n,\mathbf{s},\bm{\xi}_{n+1,\hat{\mathbf{s}}}))_{>r}\right\rVert_{\infty}.

Then by the choice of N2N_{2}, we have

w(n+1,𝐬^,𝝃n+1,𝐬^)ert(w(n,𝐬,𝝃n+1,𝐬^))>rert15r2dw(n,𝐬,𝝃n+1,𝐬^).\displaystyle\left\lVert w(n+1,\hat{\mathbf{s}},\bm{\xi}_{n+1,\hat{\mathbf{s}}})\right\rVert\geq e^{-rt}\left\lVert(w(n,\mathbf{s},\bm{\xi}_{n+1,\hat{\mathbf{s}}}))_{>r}\right\rVert_{\infty}\geq e^{-rt}\frac{1}{5r^{2}d}\left\lVert w(n,\mathbf{s},\bm{\xi}_{n+1,\hat{\mathbf{s}}})\right\rVert.

Therefore,

α𝐦ν(a(n+1)tu𝝋(𝐬+𝐬~ednt)Γ)\displaystyle\alpha_{\mathbf{m}}^{\nu}(a_{(n+1)t}u_{\bm{\varphi}}(\mathbf{s}+\tilde{\mathbf{s}}e^{-dnt})\Gamma) =w(n+1,𝐬^,𝝃n+1,𝐬^)ν\displaystyle=\left\lVert w(n+1,\hat{\mathbf{s}},\bm{\xi}_{n+1,\hat{\mathbf{s}}})\right\rVert^{-\nu}
eνrt(5r2d)νw(n,𝐬,𝝃n+1,𝐬^)ν\displaystyle\leq e^{\nu rt}(5r^{2}d)^{\nu}\left\lVert w(n,\mathbf{s},\bm{\xi}_{n+1,\hat{\mathbf{s}}})\right\rVert^{-\nu}
eνrt(10r2d)νsup𝐰d{𝟎}antu(𝐬)𝐰ν\displaystyle\leq e^{\nu rt}(10r^{2}d)^{\nu}\sup_{\mathbf{w}\in\mathbb{Z}^{d}\setminus\{\mathbf{0}\}}\left\lVert a_{nt}u(\mathbf{s})\mathbf{w}\right\rVert^{-\nu}
eνrt(10r2d)να~(antu𝝋(𝐬)Γ).\displaystyle\leq e^{\nu rt}(10r^{2}d)^{\nu}\tilde{\alpha}(a_{nt}u_{\bm{\varphi}(\mathbf{s})}\Gamma).

Combining cases 1 and 2, the lemma is proven. ∎

Lemma 6.15.

There exists t>0t>0 sufficiently large such that for any 𝐬J\mathbf{s}\in J,

I02(𝐬)α𝐦ν(a(n+1)tu𝝋(𝐬+𝐬~ednt)Γ)𝑑𝐬~14α𝐦ν(antu𝝋(𝐬)Γ).\displaystyle\int_{I_{02}(\mathbf{s})}\alpha_{\mathbf{m}}^{\nu}(a_{(n+1)t}u_{\bm{\varphi}}(\mathbf{s}+\tilde{\mathbf{s}}e^{-dnt})\Gamma)d\tilde{\mathbf{s}}\leq\frac{1}{4}\alpha_{\mathbf{m}}^{\nu}(a_{nt}u_{\bm{\varphi}}(\mathbf{s})\Gamma).
Proof.

We fix 𝐬J\mathbf{s}\in J for the rest of the proof. For 𝐬~I0\tilde{\mathbf{s}}\in I_{0}, denote 𝐬^=𝐬+𝐬~ednt\hat{\mathbf{s}}=\mathbf{s}+\tilde{\mathbf{s}}e^{-dnt}. Since 𝐬~I02(𝐬)\tilde{\mathbf{s}}\in I_{02}(\mathbf{s}), for simplicity we denote 𝝃n+1,𝐬^=𝝃n,𝐬=𝐯\bm{\xi}_{n+1,\hat{\mathbf{s}}}=\bm{\xi}_{n,\mathbf{s}}=\mathbf{v}.

Case 1: (w(n,𝐬,𝐯))rN2(w(n,𝐬,𝐯))>r\left\lVert(w(n,\mathbf{s},\mathbf{v}))_{\leq r}\right\rVert_{\infty}\geq N_{2}\left\lVert(w(n,\mathbf{s},\mathbf{v}))_{>r}\right\rVert_{\infty}, where N2=2(r+1)(dr)N_{2}=2(r+1)(d-r).

Then we have by Case 1 of Lemma 6.14,

w(n+1,𝐬^,𝐯)e(dr)t12dw(n,𝐬,𝐯).\displaystyle\left\lVert w(n+1,\hat{\mathbf{s}},\mathbf{v})\right\rVert\geq e^{(d-r)t}\frac{1}{2d}\left\lVert w(n,\mathbf{s},\mathbf{v})\right\rVert.

Hence, for t>0t>0 sufficiently large such that e(dr)νt(2d)ν14e^{-(d-r)\nu t}\cdot(2d)^{\nu}\leq\frac{1}{4}, we obtain

α𝐦ν(a(n+1)tu𝝋(𝐬+𝐬~ednt)Γ)14α𝐦ν(antu𝝋(𝐬)Γ).\displaystyle\alpha_{\mathbf{m}}^{\nu}(a_{(n+1)t}u_{\bm{\varphi}}(\mathbf{s}+\tilde{\mathbf{s}}e^{-dnt})\Gamma)\leq\frac{1}{4}\alpha_{\mathbf{m}}^{\nu}(a_{nt}u_{\bm{\varphi}}(\mathbf{s})\Gamma).

Case 2: (w(n,𝐬,𝐯))r<N2(w(n,𝐬,𝐯))>r\left\lVert(w(n,\mathbf{s},\mathbf{v}))_{\leq r}\right\rVert_{\infty}<N_{2}\left\lVert(w(n,\mathbf{s},\mathbf{v}))_{>r}\right\rVert_{\infty}.

Recall that I0=[1,1]r(dr)I_{0}=[-1,1]^{r(d-r)}. For any 𝐁SOr(dr)()\mathbf{B}\in SO_{r(d-r)}(\mathbb{R}), we have 𝐁I0I0\mathbf{B}\cdot I_{0}\subset I_{0}^{\prime}, where I0I_{0}^{\prime} is the unit ball in r(dr)\mathbb{R}^{r(d-r)}. We may choose t>0t>0 large enough such that for any 𝐬I\mathbf{s}\in I, any 𝐬~I0\tilde{\mathbf{s}}\in I_{0}^{\prime}, we have 𝐬+ednt𝐬~I\mathbf{s}+e^{-dnt}\tilde{\mathbf{s}}\in I^{\prime}, where II^{\prime} is given as in Remark 6.3. Define a function SS on I0I_{0}^{\prime} by

S(𝐬~)=i=1rwi(n+1,𝐬+𝐬~ednt,𝐯),𝐬~I0.\displaystyle S(\tilde{\mathbf{s}})=\sum_{i=1}^{r}w_{i}(n+1,\mathbf{s}+\tilde{\mathbf{s}}e^{-dnt},\mathbf{v}),\forall\tilde{\mathbf{s}}\in I_{0}^{\prime}.

Note that w(n+1,𝐬^,𝐯)1r|S(𝐬~)|\left\lVert w(n+1,\hat{\mathbf{s}},\mathbf{v})\right\rVert\geq\frac{1}{r}|S(\tilde{\mathbf{s}})|.

We will apply Lemma 6.13 to find 𝐁SOr(dr)()\mathbf{B}\in SO_{r(d-r)}(\mathbb{R}) such that after the change of basis 𝐬~=𝐁𝐬~\tilde{\mathbf{s}}^{\prime}=\mathbf{B}\tilde{\mathbf{s}}, for

A1=e(dr)tmax{4r(dr),2(r(dr))3/2+r}w(n,𝐬,𝐯),\displaystyle A_{1}=e^{(d-r)t}\max\{4\sqrt{r(d-r)},2(r(d-r))^{3/2}+r\}\left\lVert w(n,\mathbf{s},\mathbf{v})\right\rVert,
A2=e(dr)t140r3d(dr)w(n,𝐬,𝐯),\displaystyle A_{2}=e^{(d-r)t}\frac{1}{40r^{3}d(d-r)}\left\lVert w(n,\mathbf{s},\mathbf{v})\right\rVert, (6.8)

we have

A2SI0A1; A2|Ss~ij(𝐬~)|A1,i,j,𝐬~I0.\displaystyle A_{2}\leq\left\lVert S\right\rVert_{I_{0}^{\prime}}\leq A_{1};\text{ }A_{2}\leq|\frac{\partial S}{\partial\tilde{s}_{ij}^{\prime}}(\tilde{\mathbf{s}}^{\prime})|\leq A_{1},\forall i,j,\forall\tilde{\mathbf{s}}^{\prime}\in I_{0}^{\prime}. (6.9)

Applying Lemma 6.11 to S(𝐬~)S(\tilde{\mathbf{s}}^{\prime}), since 𝐁\mathbf{B} preserves Lebesgue measure, we obtain that for any ϵ>0\epsilon>0,

|{𝐬~I0:|S(𝐬~)|ϵ}|\displaystyle|\{\tilde{\mathbf{s}}\in I_{0}:|S(\tilde{\mathbf{s}})|\leq\epsilon\}| |{𝐬~I0:|S(𝐬~)|ϵ}|\displaystyle\leq|\{\tilde{\mathbf{s}}^{\prime}\in I_{0}^{\prime}:|S(\tilde{\mathbf{s}}^{\prime})|\leq\epsilon\}|
r(dr)12A1A2(ϵSI0)1r(dr)|I0|\displaystyle\leq r(d-r)\frac{12A_{1}}{A_{2}}\cdot(\frac{\epsilon}{\left\lVert S\right\rVert}_{I_{0}^{\prime}})^{\frac{1}{r(d-r)}}\cdot|I_{0}^{\prime}|
C~e1rtϵ1r(dr)w(n,𝐬,𝐯)1r(dr),\displaystyle\leq\tilde{C}\cdot e^{-\frac{1}{r}t}\cdot\epsilon^{\frac{1}{r(d-r)}}\cdot\left\lVert w(n,\mathbf{s},\mathbf{v})\right\rVert^{-\frac{1}{r(d-r)}},

where C~\tilde{C} is a constant depending only on rr and dd. Choose t>0t>0 large enough, by Lemma 6.10, with 0<ν<1r(dr)0<\nu<\frac{1}{r(d-r)},

I02(𝐬)α𝐦ν(a(n+1)tu𝝋(𝐬+𝐬~ednt)Γ)𝑑𝐬~\displaystyle\int_{I_{02}(\mathbf{s})}\alpha_{\mathbf{m}}^{\nu}(a_{(n+1)t}u_{\bm{\varphi}}(\mathbf{s}+\tilde{\mathbf{s}}e^{-dnt})\Gamma)d\tilde{\mathbf{s}} I01w(n+1,𝐬+ednt𝐬~,𝐯)ν𝑑𝐬~\displaystyle\leq\int_{I_{0}}\frac{1}{\left\lVert w(n+1,\mathbf{s}+e^{-dnt}\tilde{\mathbf{s}},\mathbf{v})\right\rVert^{\nu}}d\tilde{\mathbf{s}}
rνI01|S(𝐬~)|ν𝑑𝐬~\displaystyle\leq r^{\nu}\cdot\int_{I_{0}}\frac{1}{|S(\tilde{\mathbf{s}})|^{\nu}}d\tilde{\mathbf{s}}
rνcνC~νr(dr)eν(dr)tw(n,𝐬,𝐯)ν\displaystyle\leq r^{\nu}c_{\nu}\tilde{C}^{\nu r(d-r)}\cdot e^{-\nu(d-r)t}\left\lVert w(n,\mathbf{s},\mathbf{v})\right\rVert^{-\nu}
14α𝐦ν(antu𝝋(𝐬)Γ).\displaystyle\leq\frac{1}{4}\alpha_{\mathbf{m}}^{\nu}(a_{nt}u_{\bm{\varphi}}(\mathbf{s})\Gamma).

This prove the lemma. Therefore, it remains to achieve (6.9). Consider the function Ψ=q=1rp=1kmp𝝋qp\Psi=\sum_{q=1}^{r}\sum_{p=1}^{k}m_{p}\bm{\varphi}_{qp}, where 𝐦=(m1,,mk)t\mathbf{m}=(m_{1},\cdots,m_{k})^{t}. We have

S(𝐬~)=e(dr)(n+1)t\displaystyle S(\tilde{\mathbf{s}})=e^{(d-r)(n+1)t} [Ψ(𝐬+ednt𝐬~)Ψ(𝐬)i=1rj=1drednts~ij<(𝐯)>r𝐦,𝐞j>\displaystyle[\Psi(\mathbf{s}+e^{-dnt}\tilde{\mathbf{s}})-\Psi(\mathbf{s})-\sum_{i=1}^{r}\sum_{j=1}^{d-r}e^{-dnt}\tilde{s}_{ij}<(\mathbf{v})_{>r}\cdot\mathbf{m},\mathbf{e}_{j}>
+e(dr)nti=1rwi(n,𝐬,𝐯)].\displaystyle+e^{-(d-r)nt}\sum_{i=1}^{r}w_{i}(n,\mathbf{s},\mathbf{v})]. (6.10)

Let N3=4r(dr)N_{3}=4r(d-r), define the partition {1(𝐬),2(𝐬),3(𝐬)}\{\mathcal{I}_{1}(\mathbf{s}),\mathcal{I}_{2}(\mathbf{s}),\mathcal{I}_{3}(\mathbf{s})\} of {(i,j):1ir,1jdr}\{(i,j):1\leq i\leq r,1\leq j\leq d-r\} by

1(𝐬):={(i,j):|<(𝐯)>r𝐦,𝐞j>|=(𝐯)>r𝐦};\displaystyle\mathcal{I}_{1}(\mathbf{s}):=\{(i,j):\left|<(\mathbf{v})_{>r}\cdot\mathbf{m},\mathbf{e}_{j}>\right|=\left\lVert(\mathbf{v})_{>r}\cdot\mathbf{m}\right\rVert_{\infty}\};
2(𝐬):={(i,j):(𝐯)>r𝐦>|<(𝐯)>r𝐦,𝐞j>|1N3(𝐯)>r𝐦};\displaystyle{\mathcal{I}_{2}(\mathbf{s}):=\{(i,j):\left\lVert(\mathbf{v})_{>r}\cdot\mathbf{m}\right\rVert_{\infty}>\left|<(\mathbf{v})_{>r}\cdot\mathbf{m},\mathbf{e}_{j}>\right|\geq\frac{1}{N_{3}}\left\lVert(\mathbf{v})_{>r}\cdot\mathbf{m}\right\rVert_{\infty}\};}
3(𝐬):={(i,j):|<(𝐯)>r𝐦,𝐞j>|<1N3(𝐯)>r𝐦}.\displaystyle\mathcal{I}_{3}(\mathbf{s}):=\{(i,j):\left|<(\mathbf{v})_{>r}\cdot\mathbf{m},\mathbf{e}_{j}>\right|<\frac{1}{N_{3}}\left\lVert(\mathbf{v})_{>r}\cdot\mathbf{m}\right\rVert_{\infty}\}.

Note that by definition, 1(𝐬)\mathcal{I}_{1}(\mathbf{s})\neq\emptyset. Using (6.10), the choice of N3N_{3}, and the estimate

|Ψs~ij(𝐬+ednt𝐬~)|\displaystyle{|\frac{\partial\Psi}{\partial\tilde{s}_{ij}}(\mathbf{s}+e^{-dnt}\tilde{\mathbf{s}})|} (p=1kmp2)12(p=1k(q=1rφqpsij)2))12𝐦k12rM1N1,𝐬~I0,\displaystyle\leq(\sum_{p=1}^{k}m_{p}^{2})^{\frac{1}{2}}\cdot(\sum_{p=1}^{k}(\sum_{q=1}^{r}\frac{\partial\varphi_{qp}}{\partial s_{ij}})^{2}))^{\frac{1}{2}}\leq\left\lVert\mathbf{m}\right\rVert\frac{k^{\frac{1}{2}}\cdot rM_{1}}{N_{1}},\forall\tilde{\mathbf{s}}\in I_{0}^{\prime},

the following holds for any 𝐬~I0\tilde{\mathbf{s}}\in I_{0}^{\prime}:

  • For any (i,j)(i,j), where 1ir1\leq i\leq r, 1jdr1\leq j\leq d-r,

    |Ss~ij(𝐬~)|C1:=2e(dr)t(w(n,𝐬,𝐯))>r;\displaystyle|\frac{\partial S}{\partial\tilde{s}_{ij}}(\tilde{\mathbf{s}})|\leq C_{1}:=2e^{(d-r)t}\left\lVert(w(n,\mathbf{s},\mathbf{v}))_{>r}\right\rVert_{\infty};
  • For any (i,j)1(𝐬)(i,j)\in\mathcal{I}_{1}(\mathbf{s}),

    |Ss~ij(𝐬~)|c1:=e(dr)t(1k1/2rN1)(w(n,𝐬,𝐯))>r;\displaystyle|\frac{\partial S}{\partial\tilde{s}_{ij}}(\tilde{\mathbf{s}})|\geq c_{1}:=e^{(d-r)t}(1-\frac{k^{1/2}r}{N_{1}})\left\lVert(w(n,\mathbf{s},\mathbf{v}))_{>r}\right\rVert_{\infty};
  • For any (i,j)2(𝐬)(i,j)\in\mathcal{I}_{2}(\mathbf{s}),

    |Ss~ij(𝐬~)|>C2:=e(dr)t(1N3k1/2rN1)(w(n,𝐬,𝐯))>r;\displaystyle|\frac{\partial S}{\partial\tilde{s}_{ij}}(\tilde{\mathbf{s}})|>C_{2}:=e^{(d-r)t}(\frac{1}{N_{3}}-\frac{k^{1/2}r}{N_{1}})\left\lVert(w(n,\mathbf{s},\mathbf{v}))_{>r}\right\rVert_{\infty};
  • For any (i,j)3(𝐬)(i,j)\in\mathcal{I}_{3}(\mathbf{s}),

    |Ss~ij(𝐬~)|<c2:=e(dr)t(1N3+k1/2rN1)(w(n,𝐬,𝐯))>r.\displaystyle|\frac{\partial S}{\partial\tilde{s}_{ij}}(\tilde{\mathbf{s}})|<c_{2}:=e^{(d-r)t}(\frac{1}{N_{3}}+\frac{k^{1/2}r}{N_{1}})\left\lVert(w(n,\mathbf{s},\mathbf{v}))_{>r}\right\rVert_{\infty}.

Applying Lemma 6.13 with N=r(dr)N=r(d-r), and C1,c1,C2,c2C_{1},c_{1},C_{2},c_{2} given as above, we obtain 𝐁=𝐁(𝐬)SOr(dr)()\mathbf{B}=\mathbf{B}(\mathbf{s})\in SO_{r(d-r)}(\mathbb{R}) (Since the partition depends only on 𝐬\mathbf{s}, 𝐁\mathbf{B} depends only on 𝐬\mathbf{s}), such that after the change of basis 𝐬~=𝐁𝐬~\mathbf{\tilde{s}^{\prime}}=\mathbf{B}\mathbf{\tilde{s}}, the vector 𝐯(𝐬~)=(Ss~ij(𝐬~))ij\mathbf{v}(\tilde{\mathbf{s}}^{\prime})=(\frac{\partial S}{\partial\tilde{s}_{ij}^{\prime}}(\tilde{\mathbf{s}}^{\prime}))_{ij} satisfies the following: For any 1ir,1jdr1\leq i\leq r,1\leq j\leq d-r,

min{c1NNc2,C2}|Ss~ij(𝐬~)|C1+Nc2,𝐬~I0.\displaystyle\min\{\frac{c_{1}}{\sqrt{N}}-\sqrt{N}c_{2},C_{2}\}\leq|\frac{\partial S}{\partial\tilde{s}_{ij}^{\prime}}(\tilde{\mathbf{s}}^{\prime})|\leq C_{1}+\sqrt{N}c_{2},\forall\tilde{\mathbf{s}}^{\prime}\in I_{0}^{\prime}. (6.11)

By the assumption of Case 2, and the choice of N2N_{2}, it is elementary to verify that

min{c1NNc2,C2}e(dr)t140r3d(dr)w(n,𝐬,𝐯), and\displaystyle\min\{\frac{c_{1}}{\sqrt{N}}-\sqrt{N}c_{2},C_{2}\}\geq e^{(d-r)t}\frac{1}{40r^{3}d(d-r)}\left\lVert w(n,\mathbf{s},\mathbf{v})\right\rVert,\text{ and }
C1+Nc24r(dr)e(dr)tw(n,𝐬,𝐯).\displaystyle C_{1}+\sqrt{N}c_{2}\leq 4\sqrt{r(d-r)}e^{(d-r)t}\left\lVert w(n,\mathbf{s},\mathbf{v})\right\rVert. (6.12)

Moreover, using the expression (6.10), we obtain

SI0e(dr)t(2(r(dr))3/2+r)w(n,𝐬,𝐯).\displaystyle\left\lVert S\right\rVert_{I_{0}^{\prime}}\leq e^{(d-r)t}(2(r(d-r))^{3/2}+r)\left\lVert w(n,\mathbf{s},\mathbf{v})\right\rVert. (6.13)

Also, note that

SI0\displaystyle\left\lVert S\right\rVert_{I_{0}^{\prime}} inf𝐬~I0,(i,j)1|Ss~ij(𝐬~)|e(dr)t110r2dw(n,𝐬,𝐯).\displaystyle\geq\inf_{\tilde{\mathbf{s}}\in I_{0},(i,j)\in\mathcal{I}_{1}}|\frac{\partial S}{\partial\tilde{s}_{ij}}(\tilde{\mathbf{s}})|\geq e^{(d-r)t}\frac{1}{10r^{2}d}\left\lVert w(n,\mathbf{s},\mathbf{v})\right\rVert. (6.14)

Now we choose A1,A2A_{1},A_{2} as in (6), by (6.11)(6)(6.13)(6.14), (6.9) is achieved. This finishes the proof of the lemma. ∎

7. Proof of Proposition 2.1

Following a general strategy developed in [8, Section 6.6], we derive Proposition 2.1 from Proposition 6.1.

Let YY be a locally compact, second countable Hausdorff topological space. Let BB be a compact box in Matr×(dr)()Mat_{r\times(d-r)}(\mathbb{R}). Let ϕ:Matr×(dr)()Y\phi:Mat_{r\times(d-r)}(\mathbb{R})\to Y be a continuous map. Let f:×YYf:\mathbb{R}\times Y\to Y be a continuous map and we write f(t,y)f(t,y) as ft(y)f^{t}(y) for (t,y)×Y(t,y)\in\mathbb{R}\times Y.

Let 0={B}\mathcal{F}_{0}=\{B\}. For every nn\in\mathbb{N}, let n\mathcal{F}_{n} be a partition of elements in n1\mathcal{F}_{n-1} into countably many subboxes with positive Lebesgue measure. By construction, {n}n0\{\mathcal{F}_{n}\}_{n\in\mathbb{Z}_{\geq 0}} is a filtration. For any 𝐬B\mathbf{s}\in B, let In(𝐬)I_{n}(\mathbf{s}) denote the atom in n\mathcal{F}_{n} containing 𝐬\mathbf{s}.

Let β:Y[1,]\beta:Y\to[1,\infty] be a measurable map. Assume that β\beta satisfies the following conditions:

(1) β\beta satisfies contraction hypothesis: There exist 0<a<10<a<1 and b>0b>0 such that for any n0n\in\mathbb{Z}_{\geq 0} and any atom InI_{n} in n\mathcal{F}_{n},

Inβ(fn+1ϕ(𝐬))𝑑𝐬<aInβ(fnϕ(𝐬))𝑑𝐬+b|In|;\displaystyle\int_{I_{n}}\beta(f^{n+1}\phi(\mathbf{s}))d\mathbf{s}<a\int_{I_{n}}\beta(f^{n}\phi(\mathbf{s}))d\mathbf{s}+b|I_{n}|; (7.1)

(2) β\beta satisfies Lipschitz property: There exists a constant M>0M>0 such that for any 𝐬B\mathbf{s}\in B, any n0n\in\mathbb{Z}_{\geq 0}, and any 𝐬~In(𝐬)\tilde{\mathbf{s}}\in I_{n}(\mathbf{s}),

β(fnϕ(𝐬~))Mβ(fnϕ(𝐬)),\displaystyle\beta(f^{n}\phi(\tilde{\mathbf{s}}))\leq M\beta(f^{n}\phi(\mathbf{s})),
β(fn+1ϕ(𝐬))Mβ(fnϕ(𝐬));\displaystyle\beta(f^{n+1}\phi(\mathbf{s}))\leq M\beta(f^{n}\phi(\mathbf{s})); (7.2)

(3) β\beta is bounded on ϕ(B)\phi(B), that is, there exists l>0l>0 such that

{ϕ(𝐬):𝐬B}Yl={yY:β(y)<l}.\displaystyle\{\phi(\mathbf{s}):\mathbf{s}\in B\}\subset Y_{l}=\{y\in Y:\beta(y)<l\}. (7.3)

For any T>0T>0 and a measurable subset KK of YY, define

𝒜KT(𝐬):=1T0TχK(ftϕ(𝐬))𝑑t,\displaystyle\mathcal{A}_{K}^{T}(\mathbf{s}):=\frac{1}{T}\int_{0}^{T}\chi_{K}(f^{t}\phi(\mathbf{s}))dt,

where χK\chi_{K} is the indicator function of KK.

Lemma 7.1.

[8, Lemma 6.20] For any ϵ>0\epsilon>0, there exist 0<l1<0<l_{1}<\infty and 0<c1<10<c_{1}<1 such that for K=Yl1K=Y_{l_{1}}, and any T>1T>1,

|{𝐬B:𝒜KT(𝐬)1ϵ}|c1T|B|.\displaystyle|\{\mathbf{s}\in B:\mathcal{A}_{K}^{T}(\mathbf{s})\leq 1-\epsilon\}|\leq c_{1}^{T}|B|.
proof of Proposition 2.1.

We will apply Lemma 7.1 to Y=XY=X, B=IB=I, β=β𝐦\beta=\beta_{\mathbf{m}}, ϕ(𝐬)=u𝝋(𝐬)Γ\phi(\mathbf{s})=u_{\bm{\varphi}}(\mathbf{s})\Gamma and ft=atf^{t}=a_{t} for t>0t>0 sufficiently large so that Proposition 6.1 holds.

Recall that II is a closed cube in Matr×(dr)()Mat_{r\times(d-r)}(\mathbb{R}). We may assume that t>0t>0 is large enough such that edte^{-dt} is less than the length of each side of II.

We construct a filtration {n}n\{\mathcal{F}_{n}\}_{n\in\mathbb{N}} on II as follows. Let 0={,I}\mathcal{F}_{0}=\{\emptyset,I\}. Suppose that we have already constructed n1\mathcal{F}_{n-1}. We divide each box JJ of n1\mathcal{F}_{n-1} consecutively into cubes and boxes such that cubes have side length ednte^{-dnt} and boxes have side length between ednte^{-dnt} and 2ednt2e^{-dnt}. Then conditions (7.1)(7.2)(7.3) follows from Proposition 6.1.

Therefore, applying Lemma 7.1, we obtain l1>0l_{1}>0 such that the set KK defined by

K:={xX:β𝐦(x)<l1}\displaystyle K:=\{x\in X:\beta_{\mathbf{m}}(x)<l_{1}\}

is a compact subset of XX𝐦X\setminus X_{\mathbf{m}}, and (2.4) holds. ∎

8. Proof of variants of Theorem 1.3

Given 𝐌SLd()\mathbf{M}\in SL_{d}(\mathbb{R}), we may write

𝐌=[𝐀𝐁𝐂𝐃],\displaystyle\mathbf{M}=\begin{bmatrix}\mathbf{A}&\mathbf{B}\\ \mathbf{C}&\mathbf{D}\end{bmatrix}, (8.1)

where 𝐀Matr×r()\mathbf{A}\in Mat_{r\times r}(\mathbb{R}), 𝐁Matr×(dr)()\mathbf{B}\in Mat_{r\times(d-r)}(\mathbb{R}), 𝐂Mat(dr)×r()\mathbf{C}\in Mat_{(d-r)\times r}(\mathbb{R}), 𝐃Mat(dr)×(dr)()\mathbf{D}\in Mat_{(d-r)\times(d-r)}(\mathbb{R}). For 𝐬𝒰\mathbf{s}\in\mathcal{U}, we can write

u(𝐬)𝐌=[𝐀+𝐬𝐂𝐁+𝐬𝐃𝐂𝐃]=[𝐀(𝐬)𝐁(𝐬)𝐂𝐃].\displaystyle u(\mathbf{s})\mathbf{M}=\begin{bmatrix}\mathbf{A}+\mathbf{s}\cdot\mathbf{C}&\mathbf{B}+\mathbf{s}\cdot\mathbf{D}\\ \mathbf{C}&\mathbf{D}\end{bmatrix}=\begin{bmatrix}\mathbf{A}(\mathbf{s})&\mathbf{B}(\mathbf{s})\\ \mathbf{C}&\mathbf{D}\end{bmatrix}. (8.2)

Since 𝐌SLd()\mathbf{M}\in SL_{d}(\mathbb{R}), it is clear that the set of 𝐬𝒰\mathbf{s}\in\mathcal{U} such that det𝐀(𝐬)=0\det\mathbf{A}(\mathbf{s})=0 is a proper algebraic subvariety of 𝒰\mathcal{U} and hence, it has Lebesgue measure zero.

Therefore we can assume that det𝐀(𝐬)0\det\mathbf{A}(\mathbf{s})\neq 0, and

u(𝐬)𝐌=[𝐀(𝐬)𝟎𝐂𝐃𝐂𝐀(𝐬)1𝐁(𝐬)][𝟏r𝐀(𝐬)1𝐁(𝐬)𝟎dr,r𝟏dr].\displaystyle u(\mathbf{s})\mathbf{M}=\begin{bmatrix}\mathbf{A}(\mathbf{s})&\mathbf{0}\\ \mathbf{C}&\mathbf{D}-\mathbf{C}\mathbf{A}(\mathbf{s})^{-1}\mathbf{B}(\mathbf{s})\end{bmatrix}\cdot\begin{bmatrix}\mathbf{1}_{r}&\mathbf{A}(\mathbf{s})^{-1}\mathbf{B}(\mathbf{s})\\ \mathbf{0}_{d-r,r}&\mathbf{1}_{d-r}\end{bmatrix}.

We may write

u(𝐬)𝐌(Id,𝝋(𝐬))=(u(𝐬)𝐌(Id,𝝋(𝐬)))(u(𝐬)𝐌(Id,𝝋(𝐬)))+,\displaystyle u(\mathbf{s})\mathbf{M}(Id,\bm{\varphi}(\mathbf{s}))=(u(\mathbf{s})\mathbf{M}(Id,\bm{\varphi}(\mathbf{s})))_{-}\cdot(u(\mathbf{s})\mathbf{M}(Id,\bm{\varphi}(\mathbf{s})))_{+},

where

(u(𝐬)𝐌(Id,𝝋(𝐬)))=[𝐀(𝐬)𝟎𝐂𝐃𝐂𝐀1(𝐬)𝐁(𝐬)](Id,[𝟎(𝝋(𝐬))>r]),\displaystyle(u(\mathbf{s})\mathbf{M}(Id,\bm{\varphi}(\mathbf{s})))_{-}=\begin{bmatrix}\mathbf{A}(\mathbf{s})&\mathbf{0}\\ \mathbf{C}&\mathbf{D}-\mathbf{C}\mathbf{A}^{-1}(\mathbf{s})\mathbf{B}(\mathbf{s})\end{bmatrix}\cdot\left(Id,\begin{bmatrix}\mathbf{0}\\ (\bm{\varphi}(\mathbf{s}))_{>r}\end{bmatrix}\right),
(u(𝐬)𝐌(Id,𝝋(𝐬)))+=([𝟏r𝐀(𝐬)1𝐁(𝐬)𝟎dr,r𝟏dr],[(𝝋(𝐬))r+𝐀(𝐬)1𝐁(𝐬)(𝝋(𝐬))>r𝟎]).\displaystyle(u(\mathbf{s})\mathbf{M}(Id,\bm{\varphi}(\mathbf{s})))_{+}=\left(\begin{bmatrix}\mathbf{1}_{r}&\mathbf{A}(\mathbf{s})^{-1}\mathbf{B}(\mathbf{s})\\ \mathbf{0}_{d-r,r}&\mathbf{1}_{d-r}\end{bmatrix},\begin{bmatrix}(\bm{\varphi}(\mathbf{s}))_{\leq r}+\mathbf{A}(\mathbf{s})^{-1}\mathbf{B}(\mathbf{s})\cdot(\bm{\varphi}(\mathbf{s}))_{>r}\\ \mathbf{0}\end{bmatrix}\right).
Lemma 8.1.

For a.e. 𝐬0𝒰\mathbf{s}_{0}\in\mathcal{U}, there is an open neighborhood 𝒱\mathcal{V} of 𝐬0\mathbf{s}_{0} contained in 𝒰\mathcal{U} and an open subset 𝒱~\tilde{\mathcal{V}} of Matr×(dr)()Mat_{r\times(d-r)}(\mathbb{R}), such that the map ϕ:𝒱𝒱~\phi:\mathcal{V}\to\tilde{\mathcal{V}} defined by

ϕ(𝐬)=𝐀(𝐬)1𝐁(𝐬)\displaystyle\phi(\mathbf{s})=\mathbf{A}(\mathbf{s})^{-1}\mathbf{B}(\mathbf{s}) (8.3)

is a diffeomorphism.

Proof.

For any 𝐬0𝒰\mathbf{s}_{0}\in\mathcal{U} such that det𝐀(𝐬0)0\det\mathbf{A}(\mathbf{s}_{0})\neq 0, there is a neighborhood 𝒱\mathcal{V} of 𝐬0\mathbf{s}_{0}, for any 𝐬𝒱\mathbf{s}\in\mathcal{V}, the map ϕ(𝐬)\phi(\mathbf{s}) is well defined and differentiable. Therefore 𝒱~=ϕ(𝒱)Matr×(dr)()\tilde{\mathcal{V}}=\phi(\mathcal{V})\subset Mat_{r\times(d-r)}(\mathbb{R}) is an open subset. Let

ϕ1(𝐬~)=(𝐀𝐬~𝐁)(𝐃𝐂𝐬~)1\displaystyle\phi^{-1}(\tilde{\mathbf{s}})=(\mathbf{A}\tilde{\mathbf{s}}-\mathbf{B})(\mathbf{D}-\mathbf{C}\tilde{\mathbf{s}})^{-1}

for any 𝐬~\tilde{\mathbf{s}} such that det(𝐃𝐂𝐬~)0\det(\mathbf{D}-\mathbf{C}\tilde{\mathbf{s}})\neq 0. Now we verify that ϕ1\phi^{-1} is the inverse of ϕ\phi, that is, we need to verify that for 𝐬~=ϕ(𝐬)\tilde{\mathbf{s}}=\phi(\mathbf{s}),

𝐀𝐬~𝐁=𝐬(𝐃𝐂𝐬~).\displaystyle\mathbf{A}\tilde{\mathbf{s}}-\mathbf{B}=\mathbf{s}(\mathbf{D}-\mathbf{C}\tilde{\mathbf{s}}). (8.4)

Note that as 𝐬~=ϕ(𝐬)\tilde{\mathbf{s}}=\phi(\mathbf{s}), we have (𝐀+𝐬𝐂)𝐬~=𝐁+𝐬𝐃.(\mathbf{A}+\mathbf{s}\mathbf{C})\tilde{\mathbf{s}}=\mathbf{B}+\mathbf{s}\mathbf{D}. Therefore, left hand side of (8.4) is 𝐀𝐬~𝐁=(𝐀+𝐬𝐂𝐬𝐂)𝐬~𝐁=𝐬𝐃𝐬𝐂𝐬~\mathbf{A}\tilde{\mathbf{s}}-\mathbf{B}=(\mathbf{A}+\mathbf{s}\mathbf{C}-\mathbf{s}\mathbf{C})\tilde{\mathbf{s}}-\mathbf{B}=\mathbf{s}\mathbf{D}-\mathbf{s}\mathbf{C}\tilde{\mathbf{s}}, which is equal to the right hand side of (8.4). ∎

Proof of Theorem 1.5.

Choose 𝐬0,𝒱,𝒱~\mathbf{s}_{0},\mathcal{V},\tilde{\mathcal{V}} satisfying Lemma 8.1. By Lemma 2.3, it suffices to prove that for a.e. 𝐬𝒱\mathbf{s}\in\mathcal{V}, the point

(u(𝐬)𝐌(Id,𝝋(𝐬)))+Γ\displaystyle(u(\mathbf{s})\mathbf{M}(Id,\bm{\varphi}(\mathbf{s})))_{+}\Gamma (8.5)

is Birkhoff generic with respect to (X,μX,at)(X,\mu_{X},a_{t}). For 𝐬~𝒱~\tilde{\mathbf{s}}\in\tilde{\mathcal{V}}, define

𝝋~(𝐬~):=[(𝝋(ϕ1(𝐬~)))r+𝐬~(𝝋(ϕ1(𝐬~)))>r𝟎].\displaystyle\tilde{\bm{\varphi}}(\tilde{\mathbf{s}}):=\begin{bmatrix}(\bm{\varphi}(\phi^{-1}(\tilde{\mathbf{s}})))_{\leq r}+\tilde{\mathbf{s}}\cdot(\bm{\varphi}(\phi^{-1}(\tilde{\mathbf{s}})))_{>r}\\ \mathbf{0}\end{bmatrix}.

Applying Corollary 1.4 to u𝝋~(𝐬~)Γu_{\tilde{\bm{\varphi}}}(\tilde{\mathbf{s}})\Gamma for 𝐬~𝒱~\tilde{\mathbf{s}}\in\tilde{\mathcal{V}}, we obtain that if for any 𝐦k{𝟎}\mathbf{m}\in\mathbb{Z}^{k}\setminus\{\mathbf{0}\},

|{𝐬~𝒱~:(𝝋~(𝐬~))r𝐦𝐬~dr+r}|=0,\displaystyle|\{\tilde{\mathbf{s}}\in\tilde{\mathcal{V}}:(\tilde{\bm{\varphi}}(\tilde{\mathbf{s}}))_{\leq r}\cdot\mathbf{m}\in\tilde{\mathbf{s}}\cdot\mathbb{Z}^{d-r}+\mathbb{Z}^{r}\}|=0, (8.6)

then for a.e. 𝐬~𝒱~\tilde{\mathbf{s}}\in\tilde{\mathcal{V}}, u𝝋~(𝐬~)Γu_{\tilde{\bm{\varphi}}}(\tilde{\mathbf{s}})\Gamma is Birkhoff generic with respect to (X,μX,at)(X,\mu_{X},a_{t}).

Suppose for some 𝐬~𝒱~\tilde{\mathbf{s}}\in\tilde{\mathcal{V}}, and some 𝐦k{𝟎}\mathbf{m}\in\mathbb{Z}^{k}\setminus\{\mathbf{0}\},

(𝝋~(𝐬~))r𝐦𝐬~dr+r.\displaystyle(\tilde{\bm{\varphi}}(\tilde{\mathbf{s}}))_{\leq r}\cdot\mathbf{m}\in\tilde{\mathbf{s}}\cdot\mathbb{Z}^{d-r}+\mathbb{Z}^{r}. (8.7)

Let 𝐬=ϕ1(𝐬~)\mathbf{s}=\phi^{-1}(\tilde{\mathbf{s}}). By definition of ϕ\phi and 𝝋~\tilde{\bm{\varphi}}, (8.7) implies that there exist 𝐚dr\mathbf{a}\in\mathbb{Z}^{d-r} and 𝐛r\mathbf{b}\in\mathbb{Z}^{r} such that

(𝐀(𝐬)(𝝋(𝐬))r+𝐁(𝐬)(𝝋(𝐬))>r)𝐦=𝐁(𝐬)𝐚+𝐀(𝐬)𝐛,\displaystyle(\mathbf{A}(\mathbf{s})\cdot(\bm{\varphi}(\mathbf{s}))_{\leq r}+\mathbf{B}(\mathbf{s})\cdot(\bm{\varphi}(\mathbf{s}))_{>r})\cdot\mathbf{m}=\mathbf{B}(\mathbf{s})\cdot\mathbf{a}+\mathbf{A}(\mathbf{s})\cdot\mathbf{b},

then (8.7) implies that

[𝐀(𝐬)𝐁(𝐬)𝐂𝐃]𝝋(𝐬)𝐦=[𝐁(𝐬)𝐚+𝐀(𝐬)𝐛(𝐂(𝝋(𝐬))r+𝐃(𝝋(𝐬))>r)𝐦].\displaystyle\begin{bmatrix}\mathbf{A}(\mathbf{s})&\mathbf{B}(\mathbf{s})\\ \mathbf{C}&\mathbf{D}\end{bmatrix}\cdot\bm{\varphi}(\mathbf{s})\cdot\mathbf{m}=\begin{bmatrix}\mathbf{B}(\mathbf{s})\cdot\mathbf{a}+\mathbf{A}(\mathbf{s})\cdot\mathbf{b}\\ (\mathbf{C}\cdot(\bm{\varphi}(\mathbf{s}))_{\leq r}+\mathbf{D}\cdot(\bm{\varphi}(\mathbf{s}))_{>r})\cdot\mathbf{m}\end{bmatrix}. (8.8)

By (8.2), (8.8) further implies that

𝝋(𝐬)𝐦\displaystyle\bm{\varphi}(\mathbf{s})\cdot\mathbf{m} =𝐌1u(𝐬)[𝐁(𝐬)𝐚+𝐀(𝐬)𝐛(𝐂(𝝋(𝐬))r+𝐃(𝝋(𝐬))>r)𝐦]\displaystyle=\mathbf{M}^{-1}u(-\mathbf{s})\cdot\begin{bmatrix}\mathbf{B}(\mathbf{s})\cdot\mathbf{a}+\mathbf{A}(\mathbf{s})\cdot\mathbf{b}\\ (\mathbf{C}\cdot(\bm{\varphi}(\mathbf{s}))_{\leq r}+\mathbf{D}\cdot(\bm{\varphi}(\mathbf{s}))_{>r})\cdot\mathbf{m}\end{bmatrix}
d+𝐌1u(𝐬)[𝟎dr].\displaystyle\in\mathbb{Z}^{d}+\mathbf{M}^{-1}u(-\mathbf{s})\cdot\begin{bmatrix}\mathbf{0}\\ \mathbb{R}^{d-r}\end{bmatrix}.

Therefore, the condition (1.9) implies (8.6). By definition, for any 𝐬𝒱\mathbf{s}\in\mathcal{V},

(u(𝐬)𝐌(Id,𝝋(𝐬)))+=u𝝋~(𝐬~), where 𝐬~=ϕ(𝐬).\displaystyle(u(\mathbf{s})\mathbf{M}(Id,\bm{\varphi}(\mathbf{s})))_{+}=u_{\tilde{\bm{\varphi}}}(\tilde{\mathbf{s}}),\text{ where }\tilde{\mathbf{s}}=\phi(\mathbf{s}).

This finishes the proof. ∎

Proof of Corollary 1.8.

Note that u𝝋(𝐬)(𝐌,𝐯)=u(𝐬)𝐌(Id,𝝋~(𝐬))u_{\bm{\varphi}}(\mathbf{s})\cdot(\mathbf{M},\mathbf{v})=u(\mathbf{s})\mathbf{M}(Id,\tilde{\bm{\varphi}}(\mathbf{s})), where 𝝋~(𝐬)=𝐌𝟏(𝝋(𝐬)+𝐯).\tilde{\bm{\varphi}}(\mathbf{s})=\mathbf{M^{-1}}(\bm{\varphi}(\mathbf{s})+\mathbf{v}). Applying Theorem 1.5 to u(𝐬)𝐌(Id,𝝋~(𝐬))Γu(\mathbf{s})\mathbf{M}(Id,\tilde{\bm{\varphi}}(\mathbf{s}))\Gamma, we obtain that if for any 𝐦{𝟎}\mathbf{m}\in\mathbb{Z}\setminus\{\mathbf{0}\},

|{𝐬𝒰:𝝋~(𝐬)𝐦𝐌1u(𝐬)[𝟎dr]+d}|=0,\displaystyle|\{\mathbf{s}\in\mathcal{U}:\tilde{\bm{\varphi}}(\mathbf{s})\cdot\mathbf{m}\in\mathbf{M}^{-1}u(-\mathbf{s})\cdot\begin{bmatrix}\mathbf{0}\\ \mathbb{R}^{d-r}\end{bmatrix}+\mathbb{Z}^{d}\}|=0, (8.9)

then for Lebesgue a.e. 𝐬𝒰\mathbf{s}\in\mathcal{U}, u(𝐬)𝐌(Id,𝝋~(𝐬))Γu(\mathbf{s})\mathbf{M}(Id,\tilde{\bm{\varphi}}(\mathbf{s}))\Gamma is Birkhoff generic with respect to (X,μX,at)(X,\mu_{X},a_{t}). By definition of 𝝋~\tilde{\bm{\varphi}}, (8.9) is equivalent to

|{𝐬𝒰:(𝝋(𝐬)+𝐯)𝐦u(𝐬)[𝟎dr]+𝐌d}|=0.\displaystyle|\{\mathbf{s}\in\mathcal{U}:(\bm{\varphi}(\mathbf{s})+\mathbf{v})\cdot\mathbf{m}\in u(-\mathbf{s})\cdot\begin{bmatrix}\mathbf{0}\\ \mathbb{R}^{d-r}\end{bmatrix}+\mathbf{M}\cdot\mathbb{Z}^{d}\}|=0.

The corollary is proven, ∎

Proof of Corollary 1.9.

Fix an 𝐬0𝒰\mathbf{s}_{0}\in\mathcal{U} at which the map 𝐬𝐄𝟏(𝐬)1[𝐞r+1,,𝐞d]\mathbf{s}\mapsto\mathbf{E_{1}}(\mathbf{s})^{-1}\cdot[\mathbf{e}_{r+1},\cdots,\mathbf{e}_{d}] has a nonsingular differential. It is enough to prove the corollary for a.e. 𝐬\mathbf{s} in a neighborhood of 𝐬0\mathbf{s}_{0}. Choose a neighborhood 𝒱\mathcal{V} of 𝐬0\mathbf{s}_{0} such that for any 𝐬𝒱\mathbf{s}\in\mathcal{V}, as in (8.1) we can write

𝐄2(𝐬)=𝐄1(𝐬)𝐄1(𝐬0)1=[𝐀(𝐬)𝐁(𝐬)𝐂(𝐬)𝐃(𝐬)],\displaystyle\mathbf{E}_{2}(\mathbf{s})=\mathbf{E}_{1}(\mathbf{s})\cdot\mathbf{E}_{1}(\mathbf{s}_{0})^{-1}=\begin{bmatrix}\mathbf{A}(\mathbf{s})&\mathbf{B}(\mathbf{s})\\ \mathbf{C}(\mathbf{s})&\mathbf{D}(\mathbf{s})\end{bmatrix},

where det𝐀(𝐬)0\det\mathbf{A}(\mathbf{s})\neq 0 and det𝐃(𝐬)0\det\mathbf{D}(\mathbf{s})\neq 0. This can be done by smoothness of 𝐄1\mathbf{E}_{1}.

Since 𝐄2(𝐬)SOd()\mathbf{E}_{2}(\mathbf{s})\in SO_{d}(\mathbb{R}), 𝐄2(𝐬)𝐄2(𝐬)t=Id\mathbf{E}_{2}(\mathbf{s})\cdot\mathbf{E}_{2}(\mathbf{s})^{t}=Id, that is,

[𝐀(𝐬)𝐁(𝐬)𝐂(𝐬)𝐃(𝐬).][𝐀(𝐬)t𝐂(𝐬)t𝐁(𝐬)t𝐃(𝐬)t]\displaystyle\begin{bmatrix}\mathbf{A}(\mathbf{s})&\mathbf{B}(\mathbf{s})\\ \mathbf{C}(\mathbf{s})&\mathbf{D}(\mathbf{s}).\end{bmatrix}\cdot\begin{bmatrix}\mathbf{A}(\mathbf{s})^{t}&\mathbf{C}(\mathbf{s})^{t}\\ \mathbf{B}(\mathbf{s})^{t}&\mathbf{D}(\mathbf{s})^{t}\end{bmatrix} =Id.\displaystyle=Id.

In particular, we have

𝐀(𝐬)𝐂(𝐬)t+𝐁(𝐬)𝐃(𝐬)t=𝟎.\displaystyle\mathbf{A}(\mathbf{s})\cdot\mathbf{C}(\mathbf{s})^{t}+\mathbf{B}(\mathbf{s})\cdot\mathbf{D}(\mathbf{s})^{t}=\mathbf{0}.

We may write

𝐄2(𝐬)=𝐄2(𝐬)u(𝐂(𝐬)t(𝐃(𝐬)t)1),\displaystyle\mathbf{E}_{2}(\mathbf{s})=\mathbf{E}_{2}(\mathbf{s})_{-}\cdot u(-\mathbf{C}(\mathbf{s})^{t}\cdot(\mathbf{D}(\mathbf{s})^{t})^{-1}),

where

𝐄2(𝐬)=[𝐀(𝐬)𝟎𝐂(𝐬)𝐃(𝐬)𝐂𝐀(𝐬)1𝐁(𝐬)].\displaystyle\mathbf{E}_{2}(\mathbf{s})_{-}=\begin{bmatrix}\mathbf{A}(\mathbf{s})&\mathbf{0}\\ \mathbf{C}(\mathbf{s})&\mathbf{D}(\mathbf{s})-\mathbf{C}\mathbf{A}(\mathbf{s})^{-1}\mathbf{B}(\mathbf{s})\end{bmatrix}.

By Lemma 2.3, for any 𝐬𝒰\mathbf{s}\in\mathcal{U}, 𝐄1(𝐬)(Id,𝝋(𝐬))Γ\mathbf{E}_{1}(\mathbf{s})(Id,\bm{\varphi}(\mathbf{s}))\Gamma is Birkhoff generic with respect to (X,μX,at)(X,\mu_{X},a_{t}) if and only if

u(𝐂(𝐬)t(𝐃(𝐬)t)1)𝐄1(𝐬0)(Id,𝝋(𝐬))Γ\displaystyle u(-\mathbf{C}(\mathbf{s})^{t}\cdot(\mathbf{D}(\mathbf{s})^{t})^{-1})\cdot\mathbf{E}_{1}(\mathbf{s}_{0})(Id,\bm{\varphi}(\mathbf{s}))\Gamma

is Birkhoff generic with respect to (X,μX,at)(X,\mu_{X},a_{t}).

By assumption, the map

𝐬𝐄2(𝐬)1[𝐞r+1,,𝐞d]=[𝐂(𝐬)t𝐃(𝐬)t]\displaystyle\mathbf{s}\mapsto\mathbf{E}_{2}(\mathbf{s})^{-1}\cdot[\mathbf{e}_{r+1},\cdots,\mathbf{e}_{d}]=\begin{bmatrix}\mathbf{C}(\mathbf{s})^{t}\\ \mathbf{D}(\mathbf{s})^{t}\end{bmatrix}

has nonsingular differential at 𝐬0\mathbf{s}_{0}. Thus the map

ϕ:𝐬𝐂(𝐬)t(𝐃(𝐬)t)1\displaystyle\phi:\mathbf{s}\mapsto-\mathbf{C}(\mathbf{s})^{t}\cdot(\mathbf{D}(\mathbf{s})^{t})^{-1} (8.10)

also has nonsingular differential at 𝐬0\mathbf{s}_{0}.

Shrink the neighborhood 𝒱\mathcal{V} of 𝐬0\mathbf{s}_{0} if necessary, we can assume that there exists an open subset 𝒱~\tilde{\mathcal{V}} of Matr×(dr)()Mat_{r\times(d-r)}(\mathbb{R}) such that ϕ:𝒱𝒱~\phi:\mathcal{V}\to\tilde{\mathcal{V}} is a diffeomorphism. Denote ϕ1\phi^{-1} the inverse of ϕ\phi.

Let 𝝋~(𝐬~)=𝝋(ϕ1(𝐬~))\tilde{\bm{\varphi}}(\tilde{\mathbf{s}})=\bm{\varphi}(\phi^{-1}(\tilde{\mathbf{s}})). Applying Theorem 1.5 to

{u(𝐬~)𝐄1(𝐬0)(Id,𝝋~(𝐬~))Γ:𝐬~𝒱~},\displaystyle\{u(\tilde{\mathbf{s}})\mathbf{E}_{1}(\mathbf{s}_{0})(Id,\tilde{\bm{\varphi}}(\tilde{\mathbf{s}}))\Gamma:\tilde{\mathbf{s}}\in\tilde{\mathcal{V}}\},

we obtain that if for any 𝐦k{𝟎}\mathbf{m}\in\mathbb{Z}^{k}\setminus\{\mathbf{0}\},

|{𝐬~𝒱~:𝝋~(𝐬~)𝐦𝐄1(𝐬0)1u(𝐬~)[𝟎dr]+d}|=0,\displaystyle|\{\tilde{\mathbf{s}}\in\tilde{\mathcal{V}}:\tilde{\bm{\varphi}}(\tilde{\mathbf{s}})\cdot\mathbf{m}\in\mathbf{E}_{1}(\mathbf{s}_{0})^{-1}\cdot u(-\tilde{\mathbf{s}})\cdot\begin{bmatrix}\mathbf{0}\\ \mathbb{R}^{d-r}\end{bmatrix}+\mathbb{Z}^{d}\}|=0, (8.11)

then for a.e. 𝐬~𝒱~\tilde{\mathbf{s}}\in\tilde{\mathcal{V}}, u(𝐬~)𝐄1(𝐬0)(Id,𝝋~(𝐬~))Γu(\tilde{\mathbf{s}})\mathbf{E}_{1}(\mathbf{s}_{0})(Id,\tilde{\bm{\varphi}}(\tilde{\mathbf{s}}))\Gamma is Birkhoff generic with respect to (X,μX,at)(X,\mu_{X},a_{t}). Since ϕ\phi is a diffeomorphism, (8.11) is equivalent to

|{𝐬𝒱:𝝋(𝐬)𝐦𝐄1(𝐬)1[𝟎dr]+d}|=0.\displaystyle|\{\mathbf{s}\in\mathcal{V}:\bm{\varphi}(\mathbf{s})\cdot\mathbf{m}\in\mathbf{E}_{1}(\mathbf{s})^{-1}\cdot\begin{bmatrix}\mathbf{0}\\ \mathbb{R}^{d-r}\end{bmatrix}+\mathbb{Z}^{d}\}|=0.

This completes the proof. ∎

9. Application to universal hitting time statistics for integrable flows

9.1. An adapted form of Corollary 1.9

Following notations of [3], for l>0l>0, let

D(el)=diag[e(d1)l,el,,el].\displaystyle D(e^{-l})=diag[e^{-(d-1)l},e^{l},\cdots,e^{l}].
Theorem 9.1.

Let 𝒰\mathcal{U} be a bounded open subset of d1\mathbb{R}^{d-1} and φ:𝒰(d)k\varphi:\mathcal{U}\to(\mathbb{R}^{d})^{k} be a C1C^{1} map. Let 𝐄𝟏:𝒰SOd()\mathbf{E_{1}}:\mathcal{U}\to SO_{d}(\mathbb{R}) be a smooth map such that the map 𝐬𝐄𝟏(𝐬)1𝐞1\mathbf{s}\mapsto\mathbf{E_{1}}(\mathbf{s})^{-1}\cdot\mathbf{e}_{1} has a nonsingular differential at Lebesgue almost every 𝐬𝒰\mathbf{s}\in\mathcal{U}. Assume that for any 𝐦k{𝟎}\mathbf{m}\in\mathbb{Z}^{k}\setminus\{\mathbf{0}\},

|{𝐬𝒰:φ(𝐬)𝐦𝐄1(𝐬)1𝐞1+d}|=0.\displaystyle|\{\mathbf{s}\in\mathcal{U}:\varphi(\mathbf{s})\cdot\mathbf{m}\in\mathbb{R}\mathbf{E}_{1}(\mathbf{s})^{-1}\cdot\mathbf{e}_{1}+\mathbb{Z}^{d}\}|=0.

Then for Lebesgue a.e. 𝐬𝒰\mathbf{s}\in\mathcal{U}, 𝐄1(𝐬)(Id,φ(𝐬))Γ\mathbf{E}_{1}(\mathbf{s})(Id,\varphi(\mathbf{s}))\Gamma is Birkhoff generic with respect to (X,μX,D(el))(X,\mu_{X},D(e^{-l})).

Proof.

For any l>0l>0, denote al=diag[el,,el,e(d1)l]a_{l}=diag[e^{l},\cdots,e^{l},e^{-(d-1)l}]. Choose ωSOd()SLd()\omega\in SO_{d}(\mathbb{R})\cap SL_{d}(\mathbb{Z}) such that for any

ω1D(el)ω=al.\displaystyle\omega^{-1}\cdot D(e^{-l})\cdot\omega=a_{l}.

Note that since ωSLd()\omega\in SL_{d}(\mathbb{Z}),

ω1D(el)E1(𝐬)(Id,φ(𝐬))Γ=alω1E1(𝐬)ω(Id,ω1φ(𝐬))Γ.\displaystyle\omega^{-1}D(e^{-l})E_{1}(\mathbf{s})(Id,\varphi_{(}\mathbf{s}))\Gamma=a_{l}\omega^{-1}E_{1}(\mathbf{s})\omega(Id,\omega^{-1}\varphi(\mathbf{s}))\Gamma.

Applying Corollary 1.9, we obtain that for a.e. 𝐬𝒰\mathbf{s}\in\mathcal{U}, ω1E1(𝐬)ω(Id,ω1φ(𝐬))Γ\omega^{-1}E_{1}(\mathbf{s})\omega(Id,\omega^{-1}\varphi(\mathbf{s}))\Gamma is Birkhoff generic with respect to (X,μX,al)(X,\mu_{X},a_{l}), and the theorem follows. ∎

9.2. Universal hitting time

Let (,,ν)(\mathcal{M},\mathcal{B},\nu) be a measurable space with probability measure ν\nu, and φt:\varphi^{t}:\mathcal{M}\to\mathcal{M} be a measure-preserving dynamical system. Given some target set 𝒟\mathcal{D}\subset\mathcal{M}, it is natural to study how often a φ\varphi-trajectory along random initial data xx\in\mathcal{M} intersects this target set. On the other hand, another question is to consider a sequence of randomized target sets whose "size" shrink to zero, and study the distribution of intersection times of a random φ\varphi-trajectory with these shrinking targets. The interested reader is referred to [3] and the references therein for a survey of the history of aforementioned questions.

In the setting of universal hitting time statistics for integrable flows (cf. [3]), the above question is studied when the measurable space is a dd dimensional torus 𝕋d\mathbb{T}^{d} and φt\varphi^{t} is a linear flow on the torus. Now let d2d\geq 2 and k1k\geq 1 be fixed integers. In this article, the sequence of target sets we consider is a sequence of union of kk many bounded codimensional one balls in 𝕋d\mathbb{T}^{d}, whose radius shrink to zero. More precisely, let 𝒰\mathcal{U} be a bounded open subset of d1\mathbb{R}^{d-1}. Consider the following smooth functions:

𝜽,ϕj:𝒰𝕋d,1jk,\displaystyle\bm{\theta},\bm{\phi}_{j}:\mathcal{U}\to\mathbb{T}^{d},1\leq j\leq k,
𝐟,𝐮j:𝒰𝐒1d1,1jk,\displaystyle\mathbf{f},\mathbf{u}_{j}:\mathcal{U}\to\mathbf{S}_{1}^{d-1},1\leq j\leq k,

where 𝐒1d1\mathbf{S}_{1}^{d-1} is the unit one sphere in d\mathbb{R}^{d}. For the functions above, we assign to any 𝐬𝒰\mathbf{s}\in\mathcal{U} the following:

  • 𝜽(𝐬)=\bm{\theta}(\mathbf{s})=initial position of the flow;

  • 𝐟(𝐬)=\mathbf{f}(\mathbf{s})=direction of the flow;

  • 𝐮j(𝐬)=\mathbf{u}_{j}(\mathbf{s})=direction of the jj-th target ball;

  • ϕj(𝐬)=\bm{\phi}_{j}(\mathbf{s})=center of the jj-th target ball.

With these functions, we define the flow

φt:𝒰𝕋d×𝒰,𝐬(𝜽(𝐬)+t𝐟(𝐬),𝐬).\displaystyle\varphi^{t}:\mathcal{U}\to\mathbb{T}^{d}\times\mathcal{U},\mathbf{s}\mapsto(\bm{\theta}(\mathbf{s})+t\mathbf{f}(\mathbf{s}),\mathbf{s}). (9.1)

From now on, we fix a map 𝐯𝐑𝐯\mathbf{v}\mapsto\mathbf{R}_{\mathbf{v}} from 𝐒1d1\mathbf{S}_{1}^{d-1} to SOd()SO_{d}(\mathbb{R}) such that for all 𝐯𝐒1d1\mathbf{v}\in\mathbf{S}_{1}^{d-1},

𝐑𝐯𝐯=𝐞1,\displaystyle\mathbf{R}_{\mathbf{v}}\cdot\mathbf{v}=\mathbf{e}_{1}, (9.2)

and 𝐯𝐑𝐯\mathbf{v}\mapsto\mathbf{R}_{\mathbf{v}} is smooth on 𝐒1d1{𝐯0}\mathbf{S}_{1}^{d-1}\setminus\{\mathbf{v}_{0}\} for a singular point 𝐯0𝐒1d1\mathbf{v}_{0}\in\mathbf{S}_{1}^{d-1}. For 1jk1\leq j\leq k, fix a bounded open subset Ωjd1×𝒰\Omega_{j}\subset\mathbb{R}^{d-1}\times\mathcal{U}. For any l>0l>0, denote the ll-level target set with to be

𝒟l=j=1k𝒟l(𝐮j,ϕj,Ωj),\displaystyle\mathcal{D}_{l}=\bigcup_{j=1}^{k}\mathcal{D}_{l}(\mathbf{u}_{j},\bm{\phi}_{j},\Omega_{j}),

where

𝒟l(𝐮j,ϕj,Ωj)={(ϕj(𝐬)+el𝐑𝐮j(𝐬)1[0𝐱],𝐬)𝕋d×𝒰:(𝐱,𝐬)Ωj}.\displaystyle\mathcal{D}_{l}(\mathbf{u}_{j},\bm{\phi}_{j},\Omega_{j})=\{(\bm{\phi}_{j}(\mathbf{s})+e^{-l}\mathbf{R}^{-1}_{\mathbf{u}_{j}(\mathbf{s})}\cdot\begin{bmatrix}0\\ \mathbf{x}\end{bmatrix},\mathbf{s})\in\mathbb{T}^{d}\times\mathcal{U}:(\mathbf{x},\mathbf{s})\in\Omega_{j}\}.

Note that l>0l>0 parameterizes the size of the target. For any 𝐬𝒰\mathbf{s}\in\mathcal{U}, let 𝒯(𝐬,𝒟l)\mathcal{T}(\mathbf{s},\mathcal{D}_{l}) be the set of hitting times defined by

𝒯(𝐬,𝒟l):={t>0:φt(𝐬)𝒟l}.\displaystyle\mathcal{T}(\mathbf{s},\mathcal{D}_{l}):=\{t>0:\varphi^{t}(\mathbf{s})\in\mathcal{D}_{l}\}.

This is a discrete subset of >0\mathbb{R}_{>0}, and we label its elements by

0<t1(𝐬,𝒟l)<t2(𝐬,𝒟l)<.\displaystyle 0<t_{1}(\mathbf{s},\mathcal{D}_{l})<t_{2}(\mathbf{s},\mathcal{D}_{l})<\cdots.

By Santalo’s formula (cf. [2]), if 𝐬𝒰\mathbf{s}\in\mathcal{U} is such that the components of 𝐟(𝐬)\mathbf{f}(\mathbf{s}) are not rationally related, then for any nn\in\mathbb{N}, the normalized nn-th return time to target 𝒟l\mathcal{D}_{l} is

tn(𝐬,𝒟l)e(d1)lσ¯(𝐬),\displaystyle\frac{t_{n}(\mathbf{s},\mathcal{D}_{l})}{e^{(d-1)l}\cdot\overline{\sigma}(\mathbf{s})},

where e(d1)lσ¯(𝐬)e^{(d-1)l}\cdot\overline{\sigma}(\mathbf{s}) is the mean return time (cf. [3, Section 2]), and

σ¯(𝐬)=1j=1k|Ωj(𝐬)|𝐮j(𝐬)𝐟(𝐬), Ωj(𝐬)={𝐱d1:(𝐱,𝐬)Ωj}.\displaystyle\overline{\sigma}(\mathbf{s})=\frac{1}{\sum_{j=1}^{k}|\Omega_{j}(\mathbf{s})|\mathbf{u}_{j}(\mathbf{s})\cdot\mathbf{f}(\mathbf{s})},\textbf{ }\Omega_{j}(\mathbf{s})=\{\mathbf{x}\in\mathbb{R}^{d-1}:(\mathbf{x},\mathbf{s})\in\Omega_{j}\}.
Definition 9.2.

The smooth map 𝐟:𝒰𝐒1d1\mathbf{f}:\mathcal{U}\to\mathbf{S}_{1}^{d-1} is regular if the push forward of Lebesgue measure on 𝒰\mathcal{U} under 𝐟\mathbf{f} is absolutely continuous with respect to the Haar measure on 𝐒1d1\mathbf{S}_{1}^{d-1}.

The following is a corollary of Theorem 9.1:

Corollary 9.3.

Let 𝒰\mathcal{U} be a bounded open subset of d1\mathbb{R}^{d-1}. Let 𝐟:𝒰𝐒1d1\mathbf{f}:\mathcal{U}\to\mathbf{S}_{1}^{d-1} be a regular smooth map. Let 𝛗:𝒰(d)k\bm{\varphi}:\mathcal{U}\to(\mathbb{R}^{d})^{k} be a C1C^{1} map. If for any 𝐦k{𝟎}\mathbf{m}\in\mathbb{Z}^{k}\setminus\{\mathbf{0}\},

|{𝐬𝒰:𝝋(𝐬)𝐦𝐟(𝐬)+d}|=0,\displaystyle|\{\mathbf{s}\in\mathcal{U}:\bm{\varphi}(\mathbf{s})\cdot\mathbf{m}\in\mathbb{R}\mathbf{f}(\mathbf{s})+\mathbb{Z}^{d}\}|=0, (9.3)

then for Lebesgue a.e. 𝐬𝒰\mathbf{s}\in\mathcal{U}, 𝐑𝐟(𝐬)(Id,𝛗(𝐬))Γ\mathbf{R}_{\mathbf{f}(\mathbf{s})}(Id,\bm{\varphi}(\mathbf{s}))\Gamma is Birkhoff generic with respect to (X,μX,D(el))(X,\mu_{X},D(e^{-l})).

Proof.

By assumption, 𝐟\mathbf{f} is a regular smooth map from 𝒰\mathcal{U} to 𝐒1d1\mathbf{S}_{1}^{d-1}. By Sard’s theorem, the set of critical points of 𝐟\mathbf{f} has Lebesgue measure 0. For any 𝐬0𝒰\mathbf{s}_{0}\in\mathcal{U} which is not a critical point of 𝐟\mathbf{f}, there exists an open neighborhood 𝒱\mathcal{V} of 𝐬0\mathbf{s}_{0} such that 𝐟\mathbf{f} is a diffeomorphism of 𝒱\mathcal{V} to some open subset of 𝐒1d1\mathbf{S}_{1}^{d-1}.

Therefore, the map 𝐬𝐑𝐟(𝐬)1𝐞1=𝐟(𝐬)\mathbf{s}\mapsto\mathbf{R}_{\mathbf{f}(\mathbf{s})}^{-1}\cdot\mathbf{e}_{1}=\mathbf{f}(\mathbf{s}) has nonsingular differentials for all 𝐬𝒱\mathbf{s}\in\mathcal{V}. Then we apply Theorem 9.1 to 𝐄1(𝐬)=𝐑𝐟(𝐬)\mathbf{E}_{1}(\mathbf{s})=\mathbf{R}_{\mathbf{f}(\mathbf{s})} and the corollary follows. ∎

Definition 9.4.

The kk-tuple of smooth functions ϕ1,ϕk:𝒰𝕋d\bm{\phi}_{1},\cdots\bm{\phi}_{k}:\mathcal{U}\to\mathbb{T}^{d} is 𝜽\bm{\theta}-generic if for any 𝐦=(m1,,mk)k{𝟎}\mathbf{m}=(m_{1},\cdots,m_{k})\in\mathbb{Z}^{k}\setminus\{\mathbf{0}\}, we have

|{𝐬𝒰:j=1kmj(ϕj(𝐬)𝜽(𝐬))𝐟(𝐬)+d}|=0.\displaystyle|\{\mathbf{s}\in\mathcal{U}:\sum_{j=1}^{k}m_{j}(\bm{\phi}_{j}(\mathbf{s})-\bm{\theta}(\mathbf{s}))\in\mathbb{R}\mathbf{f}(\mathbf{s})+\mathbb{Z}^{d}\}|=0.

To state our theorem precisely, we need some preparations. Our notations follows from [3, Section 6]. Given NN\in\mathbb{N}, denote N¯={1,,N}\overline{N}=\{1,\cdots,N\}. For j{1,,k}j\in\{1,\cdots,k\} and 𝐬𝒰\mathbf{s}\in\mathcal{U}, define j(𝐬)=𝐑𝐟(𝐬)𝐑𝐮j(𝐬)1\mathfrak{R}_{j}(\mathbf{s})=\mathbf{R}_{\mathbf{f}(\mathbf{s})}\mathbf{R}^{-1}_{\mathbf{u}_{j}(\mathbf{s})}. Let ~j(𝐬)\tilde{\mathfrak{R}}_{j}(\mathbf{s}) be the matrix of the linear transformation

𝐱(j(𝐬)[0𝐱])d1,\displaystyle\mathbf{x}\mapsto(\mathfrak{R}_{j}(\mathbf{s})\begin{bmatrix}0\\ \mathbf{x}\end{bmatrix})_{\perp}\in\mathbb{R}^{d-1}, (9.4)

where 𝐮=(u2,,ud)trd1\mathbf{u}_{\perp}=(u_{2},\cdots,u_{d})^{tr}\in\mathbb{R}^{d-1} for 𝐮=(u1,,ud)trd\mathbf{u}=(u_{1},\cdots,u_{d})^{tr}\in\mathbb{R}^{d}.

For any 𝐬𝒰\mathbf{s}\in\mathcal{U}, we define

Ω~j(𝐬):=σ¯(𝐬)1d1~j(𝐬)Ωj(𝐬)d1.\displaystyle\tilde{\Omega}_{j}(\mathbf{s}):=\overline{\sigma}(\mathbf{s})^{\frac{1}{d-1}}\tilde{\mathfrak{R}}_{j}(\mathbf{s})\Omega_{j}(\mathbf{s})\subset\mathbb{R}^{d-1}. (9.5)

Let G1=SLd()dG_{1}=SL_{d}(\mathbb{R})\ltimes\mathbb{R}^{d}. For g=(g,(𝝃1,,𝝃k))Gg=(g^{\prime},(\bm{\xi}_{1},\cdots,\bm{\xi}_{k}))\in G and j{1,,k}j\in\{1,\cdots,k\}, write g[j]=(g,𝝃j)G1g^{[j]}=(g^{\prime},\bm{\xi}_{j})\in G_{1}. Our main theorem of this section is the following.

Theorem 9.5.

Let 𝒰\mathcal{U} be a bounded open subset of d1\mathbb{R}^{d-1}. For 1jk1\leq j\leq k, let 𝐟,𝛉,𝐮j,ϕj\mathbf{f},\bm{\theta},\mathbf{u}_{j},\bm{\phi}_{j} be given as in the beginning of this section. Let Ωj\Omega_{j} be a bounded open subset of d1×𝒰\mathbb{R}^{d-1}\times\mathcal{U}. For each j=1,,kj=1,\cdots,k, assume that

(1) |𝐮j1({𝐯0})|=0|\mathbf{u}_{j}^{-1}(\{\mathbf{v}_{0}\})|=0,

(2) 𝐮j(𝐬)𝐟(𝐬)>0\mathbf{u}_{j}(\mathbf{s})\cdot\mathbf{f}(\mathbf{s})>0 for all 𝐬𝒰\mathbf{s}\in\mathcal{U},

(3) for a.e. 𝐬𝒰\mathbf{s}\in\mathcal{U}, the boundary Ωj(𝐬)\partial\Omega_{j}(\mathbf{s}) has Lebesgue measure 0.

(4) |Ωj(𝐬)||\Omega_{j}(\mathbf{s})| is a smooth positive function of 𝐬𝒰\mathbf{s}\in\mathcal{U}.

Also assume that 𝐟\mathbf{f} is regular and (ϕ1,,ϕk)(\bm{\phi}_{1},\cdots,\bm{\phi}_{k}) is 𝛉\bm{\theta}-generic. Then for any NN\in\mathbb{N}, any Tn>0T_{n}>0 for nN¯n\in\overline{N}, the following holds: For a.e. 𝐬𝒰\mathbf{s}\in\mathcal{U},

limL1L|{l[0,L]:tn(𝐬,𝒟l)e(d1)lσ¯(𝐬)Tn,nN¯}|\displaystyle\lim_{L\to\infty}\frac{1}{L}|\{l\in[0,L]:\frac{t_{n}(\mathbf{s},\mathcal{D}_{l})}{e^{(d-1)l}\cdot\overline{\sigma}(\mathbf{s})}\leq T_{n},\forall n\in\overline{N}\}|
=μX({gΓX:j=1k#{[t𝐱]g[j]d:0<t<Tn,𝐱Ω~j(𝐬)}n,\displaystyle=\mu_{X}(\{g\Gamma\in X:\sum_{j=1}^{k}\#\left\{\begin{bmatrix}t\\ \mathbf{x}\end{bmatrix}\in g^{[j]}\mathbb{Z}^{d}:0<t<T_{n},\mathbf{x}\in-\tilde{\Omega}_{j}(\mathbf{s})\right\}\geq n,
nN¯}).\displaystyle\forall n\in\overline{N}\}).

Note that in [3, Theorem 2], the authors proved that for each nn\in\mathbb{N}, there is some random variable τn\tau_{n} in >0\mathbb{R}_{>0} such that the nn-th normalized hitting time tn(,𝒟l)/(e(d1)lσ¯())t_{n}(\cdot,\mathcal{D}_{l})/(e^{(d-1)l}\cdot\overline{\sigma}(\cdot)) converges to τn\tau_{n} in distribution as ll\to\infty. Unlike [3], in Theorem 9.5 we are interested in the question that given a fixed initial 𝐬\mathbf{s}, when the target is shrinking, how often the nn-th normalized hitting time is bounded by some given constant.

For any j{1,,k}j\in\{1,\cdots,k\}, any real numbers Y<ZY<Z, following [3, Eq. (8.9)], we define

𝐀~j,Y,Z={([t~j(𝐬)𝐱],𝐬):(𝐱,𝐬)Ωj,σ¯(𝐬)Y<tσ¯(𝐬)Z}.\displaystyle\tilde{\mathbf{A}}_{j,Y,Z}=\{(\begin{bmatrix}t\\ -\tilde{\mathfrak{R}}_{j}(\mathbf{s})\mathbf{x}\end{bmatrix},\mathbf{s}):(\mathbf{x},\mathbf{s})\in\Omega_{j},\overline{\sigma}(\mathbf{s})Y<t\leq\overline{\sigma}(\mathbf{s})Z\}.

Given any real numbers Yn<ZnY_{n}<Z_{n} for nN¯n\in\overline{N}, following [3, Eq.(8.10)], we define

B[(Yn),(Zn)]={(gΓ,𝐬)G/Γ×𝒰:j=1k#(A~j,Yn,Zn(𝐬)g[j]d)n,nN¯}.\displaystyle B[(Y_{n}),(Z_{n})]=\{(g\Gamma,\mathbf{s})\in G/\Gamma\times\mathcal{U}:\sum_{j=1}^{k}\#(\tilde{A}_{j,Y_{n},Z_{n}}(\mathbf{s})\cap g^{[j]}\cdot\mathbb{Z}^{d})\geq n,\forall n\in\overline{N}\}.

where

A~j,Yn,Zn(𝐬)={𝐱d:(𝐱,𝐬)A~j,Yn,Zn}.\displaystyle\tilde{A}_{j,Y_{n},Z_{n}}(\mathbf{s})=\{\mathbf{x}\in\mathbb{R}^{d}:(\mathbf{x},\mathbf{s})\in\tilde{A}_{j,Y_{n},Z_{n}}\}.

For any 𝐬𝒰\mathbf{s}\in\mathcal{U}, denote

B[(Yn),(Zn)](𝐬):={gΓG/Γ:(gΓ,𝐬)B[(Yn),(Zn)]}.\displaystyle B[(Y_{n}),(Z_{n})](\mathbf{s}):=\{g\Gamma\in G/\Gamma:(g\Gamma,\mathbf{s})\in B[(Y_{n}),(Z_{n})]\}.
Lemma 9.6.

[3, Lemma 17] For every 𝐬𝒰\mathbf{s}\in\mathcal{U}, and B=B[(Yn),(Zn)]B=B[(Y_{n}),(Z_{n})], μX(B(𝐬))=0\mu_{X}(\partial B(\mathbf{s}))=0.

Proof.

By [3, Lemma 14,16], it suffices to prove that for every j{1,,k}j\in\{1,\cdots,k\} and n{1,,N}n\in\{1,\cdots,N\}, A~j,Yn,Zn(𝐬)\partial\tilde{A}_{j,Y_{n},Z_{n}}(\mathbf{s}) has Lebesgue measure zero. Now since for any Y<ZY<Z, we have

A~j,Y,Z(𝐬)\displaystyle\partial\tilde{A}_{j,Y,Z}(\mathbf{s}) ={[t~j(𝐬)x]:𝐱Ωj(𝐬),σ¯(𝐬)Y<tσ¯(𝐬)Z}\displaystyle=\{\begin{bmatrix}t\\ -\tilde{\mathfrak{R}}_{j}(\mathbf{s})x\end{bmatrix}:\mathbf{x}\in\partial\Omega_{j}(\mathbf{s}),\overline{\sigma}(\mathbf{s})Y<t\leq\overline{\sigma}(\mathbf{s})Z\}
{[t~j(𝐬)x]:𝐱Ωj(𝐬)¯,t{σ¯(𝐬)Y,σ¯(𝐬)Z}}.\displaystyle\bigcup\{\begin{bmatrix}t\\ -\tilde{\mathfrak{R}}_{j}(\mathbf{s})x\end{bmatrix}:\mathbf{x}\in\overline{\Omega_{j}(\mathbf{s})},t\in\{\overline{\sigma}(\mathbf{s})Y,\overline{\sigma}(\mathbf{s})Z\}\}.

By assumption, for a.e. 𝐬𝒰\mathbf{s}\in\mathcal{U}, |Ωj(𝐬)|=0|\partial\Omega_{j}(\mathbf{s})|=0, the lemma follows. ∎

We are now ready to prove Theorem 9.5.

Proof of Theorem 9.5.

The proof of Theorem 9.5 is almost the same as the proof of [3, Theorem 2], except that here we use the equidistribution result for the average along ata_{t} trajectory, while in [3], the equidistribution of ata_{t} translation of the average over a bounded open subset in horospherical subgroup is used.

Let 𝝋~:𝒰(d)k\tilde{\bm{\varphi}}:\mathcal{U}\to(\mathbb{R}^{d})^{k} be a map given by

𝝋~(𝐬)=(ϕ1(𝐬)𝜽(𝐬),,ϕk(𝐬)𝜽(𝐬)).\displaystyle\tilde{\bm{\varphi}}(\mathbf{s})=(\bm{\phi}_{1}(\mathbf{s})-\bm{\theta}(\mathbf{s}),\cdots,\bm{\phi}_{k}(\mathbf{s})-\bm{\theta}(\mathbf{s})).

Since (ϕ1,,ϕk)(\bm{\phi}_{1},\cdots,\bm{\phi}_{k}) is 𝜽\bm{\theta}-generic, 𝐟\mathbf{f} is regular, assumption (9.3) in Corollary 9.3 are satisfied for the maps 𝝋~\tilde{\bm{\varphi}} and 𝐟\mathbf{f}, thus Corollary 9.3 applies.

Let B=B[(Yn),(Zn)]B=B[(Y_{n}),(Z_{n})] for Yn,ZnY_{n},Z_{n}\in\mathbb{R} and nN¯n\in\overline{N}. Since by Lemma 9.6, μX(B(𝐬))=0\mu_{X}(\partial B(\mathbf{s}))=0, for all 𝐬𝒰\mathbf{s}\in\mathcal{U}, we have for a.e. 𝐬𝒰\mathbf{s}\in\mathcal{U},

limL1L0LχB(𝐬)(D(el)𝐑𝐟(𝐬)(Id,φ~(𝐬)))𝑑l=μX(B(𝐬)).\displaystyle\lim_{L\to\infty}\frac{1}{L}\int_{0}^{L}\chi_{B(\mathbf{s})}(D(e^{-l})\mathbf{R}_{\mathbf{f}(\mathbf{s})}(Id,\tilde{\varphi}(\mathbf{s})))dl=\mu_{X}(B(\mathbf{s})).

Then the rest of the proof follows from the proof of [3, Theorem 2]. ∎

Acknowledgment. I would like to thank Yitwah Cheung, Nimish Shah, Ronggang Shi and Barak Weiss for helpful discussions and comments. I thank Jim Cogdell for pointing out some inaccuracies in the draft. I also thank anonymous referee for pointing out many inaccuracies of the paper, as well as providing a lot of helpful suggestions to improve this article.

References

  • [1] Yves Benoist and Jean-Francois Quint. Random walks on finite volume homogeneous spaces. Invent. Math., 187(1):37–59, 2012. MR 2874934
  • [2] N. Chernov. Entropy, Lyapunov exponents, and mean free path for billiards. J. Statist. Phys., 88(1-2):1–29, 1997. MR MR1468377
  • [3] Carl P. Dettmann, Jens Marklof, and Andreas Strömbergsson. Universal hitting time statistics for integrable flows. J. Stat. Phys., 166(3-4):714–749, 2017. MR MR3607587
  • [4] Manfred Einsiedler and Thomas Ward. Ergodic theory with a view towards number theory, volume 259 of Graduate Texts in Mathematics. Springer-Verlag London, Ltd., London, 2011. MR 2723325
  • [5] Alex Eskin and Gregory Margulis. Recurrence properties of random walks on finite volume homogeneous manifolds. In Random walks and geometry, pages 431–444. Walter de Gruyter, Berlin, 2004. MR 2087794
  • [6] Alex Eskin, Gregory Margulis, and Shahar Mozes. Upper bounds and asymptotics in a quantitative version of the Oppenheim conjecture. Ann. of Math. (2), 147(1):93–141, 1998. MR 1609447
  • [7] Alex Eskin, Gregory Margulis, and Shahar Mozes. Quadratic forms of signature (2,2)(2,2) and eigenvalue spacings on rectangular 2-tori. Ann. of Math. (2), 161(2):679–725, 2005. MR 2153398
  • [8] Krzysztof Frączek, Ronggang Shi, and Corinna Ulcigrai. Genericity on curves and applications: pseudo-integrable billiards, Eaton lenses and gap distributions. J. Mod. Dyn., 12:55–122, 2018. MR 3808209
  • [9] Michael Hochman and Pablo Shmerkin. Equidistribution from fractal measures. Invent. Math., 202(1):427–479, 2015. MR 3402802
  • [10] Bernard Host. Nombres normaux, entropie, translations. Israel J. Math., 91(1-3):419–428, 1995. MR 1348326
  • [11] D. Y. Kleinbock and G. A. Margulis. Flows on homogeneous spaces and Diophantine approximation on manifolds. Ann. of Math. (2), 148(1):339–360, 1998. MR 1652916
  • [12] Dmitry Kleinbock, Nicolas de Saxcé, Nimish A Shah, and Pengyu Yang. Equidistribution in the space of 3-lattices and dirichlet-improvable vectors on planar lines. arXiv preprint arXiv:2106.08860, 2021.
  • [13] Dmitry Kleinbock, Ronggang Shi, and Barak Weiss. Pointwise equidistribution with an error rate and with respect to unbounded functions. Math. Ann., 367(1-2):857–879, 2017. MR 3606456
  • [14] A. Malcev. On the representation of an algebra as a direct sum of the radical and a semi-simple subalgebra. C. R. (Doklady) Acad. Sci. URSS (N.S.), 36:42–45, 1942. MR 0007397
  • [15] Gregory Margulis. Problems and conjectures in rigidity theory. In Mathematics: frontiers and perspectives, pages 161–174. Amer. Math. Soc., Providence, RI, 2000. MR 1754775
  • [16] Gregory Margulis and Amir Mohammadi. Quantitative version of the Oppenheim conjecture for inhomogeneous quadratic forms. Duke Math. J., 158(1):121–160, 2011. MR 2794370
  • [17] Shahar Mozes. Epimorphic subgroups and invariant measures. Ergodic Theory Dynam. Systems, 15(6):1207–1210, 1995. MR 1366316
  • [18] M. S. Raghunathan. Discrete subgroups of Lie groups. Springer-Verlag, New York-Heidelberg, 1972. Ergebnisse der Mathematik und ihrer Grenzgebiete, Band 68. MR 0507234
  • [19] Marina Ratner. On Raghunathan’s measure conjecture. Ann. of Math. (2), 134(3):545–607, 1991. MR 1135878
  • [20] Marina Ratner. Raghunathan’s topological conjecture and distributions of unipotent flows. Duke Math. J., 63(1):235–280, 1991. MR 1106945
  • [21] Nimish A. Shah. Limit distributions of expanding translates of certain orbits on homogeneous spaces. Proc. Indian Acad. Sci. Math. Sci., 106(2):105–125, 1996. MR 1403756
  • [22] Nimish A. Shah. Equidistribution of expanding translates of curves and Dirichlet’s theorem on Diophantine approximation. Invent. Math., 177(3):509–532, 2009. MR 2534098
  • [23] Nimish A. Shah. Limiting distributions of curves under geodesic flow on hyperbolic manifolds. Duke Math. J., 148(2):251–279, 2009. MR 2524496
  • [24] Ronggang Shi. Expanding cone and applications to homogeneous dynamics. International Mathematics Research Notices, 2015. MR 4105521
  • [25] Ronggang Shi. Pointwise equidistribution for one parameter diagonalizable group action on homogeneous space. Trans. Amer. Math. Soc., 373(6):4189–4221, 2020. MR 4251297
  • [26] Ronggang Shi and Barak Weiss. Invariant measures for solvable groups and Diophantine approximation. Israel J. Math., 219(1):479–505, 2017. MR 3642031
  • [27] George Tomanov. Actions of maximal tori on homogeneous spaces. In Rigidity in dynamics and geometry (Cambridge, 2000), pages 407–424. Springer, Berlin, 2002. MR 1919414
  • [28] Pengyu Yang. Equidistribution of expanding translates of curves and Diophantine approximation on matrices. Invent. Math., 220(3):909–948, 2020. MR 4094972
  • [29] Jon Chaika and Alex Eskin. Every flat surface is Birkhoff and Oseledets generic in almost every direction. J. Mod. Dyn., 9:1–23, 2015.