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

On the Hausdorff dimension of Weighted Singular Vectors

Gaurav Aggarwal Gaurav Aggarwal
School of Mathematics, Tata Institute of Fundamental Research, Mumbai, India 400005
[email protected]
 and  Anish Ghosh Anish Ghosh
School of Mathematics, Tata Institute of Fundamental Research, Mumbai, India 400005
[email protected]
Abstract.

We prove a sharp upper bound on the Hausdorff dimension of weighted singular vectors in m\mathbb{R}^{m} using dynamics on homogeneous spaces, specifically the method of integral inequalities. Together with the lower bound proved recently by Kim and Park [KimPark2024], this determines the Hausdorff dimension of weighted singular vectors, thereby generalizing to arbitrary dimension, the work of Liao, Shi, Solan, and Tamam [LSST], who determined the Hausdorff dimension of weighted singular vectors in two dimensions. We also provide the first known bounds for the Hausdorff dimension of weighted singular vectors restricted to fractal subsets.

Key words and phrases:
Diophantine approximation, ergodic theory, Hausdorff dimension, flows on homogeneous spaces
2020 Mathematics Subject Classification:
11J13, 11J83, 37A17
A. Ghosh gratefully acknowledges support from a grant from the TNQ foundation under the “Numbers and Shapes” initiative and an endowment from the Infosys foundation. G. Aggarwal and A. Ghosh gratefully acknowledges a grant from the Department of Atomic Energy, Government of India, under project 12R&DTFR5.01050012-R\&D-TFR-5.01-0500.

1. Introduction

Singular vectors are a class of vectors for which Dirichlet’s theorem in Diophantine approximation can be ‘infinitely improved’. More precisely,

Definition 1.1.

Fix mm\in\mathbb{N} and w=(w1,,wm)mw=(w_{1},\ldots,w_{m})\in\mathbb{R}^{m} such that w1w2wm>0w_{1}\geq w_{2}\geq\ldots\geq w_{m}>0 and w1++wm=1w_{1}+\ldots+w_{m}=1. A vector θ=(θ1,,θm)m\theta=(\theta_{1},\ldots,\theta_{m})\in\mathbb{R}^{m} is called ww-singular if for any ε>0\varepsilon>0, there exists Tε>0T_{\varepsilon}>0 such that for all T>TεT>T_{\varepsilon} the following system of inequalities

|pi+qθi|\displaystyle|p_{i}+q\theta_{i}| εTwi for all i=1,,m\displaystyle\leq\frac{\varepsilon}{T^{w_{i}}}\quad\text{ for all }i=1,\ldots,m
0\displaystyle 0 <q<T,\displaystyle<q<T,

has an integer solution (p,q)m×(p,q)\in\mathbb{Z}^{m}\times\mathbb{N}. We denote by Sing(w)m\text{Sing}(w)\subset\mathbb{R}^{m}, the set of all ww-singular vectors in m\mathbb{R}^{m}.

While the study of the classical ‘non-weighted’ situation, i.e. the study of Sing(1/m,,1/m)\text{Sing}(1/m,\dots,1/m) is more ubiquitous, Diophantine approximation with weights has been extensively studied in recent times. We refer the reader to [CGGMS] for an introduction. The set of singular vectors is clearly non-empty and due to a classical result of Khintchine, it contains uncountably many vectors when the dimension is greater than 11. Moreover, a modification of a classical argument in [Casselsbook] shows that the set of singular vectors has zero Lebesgue measure. There is a closely related notion of Dirichlet improvable vectors, see [KleinbockRao] for the relation between singular vectors with Dirichlet improvable vectors with respect to different norms.

Two central questions about singular vectors have recently been the focus of several important works:

  1. (1)

    to estimate the measure of Sing(w)\text{Sing}(w) with respect to other natural measures. This problem encapsulates the themes of Diophantine approximation on manifolds as well as that on fractals,

  2. (2)

    to estimate the Hausdorff dimension of Sing(w)\text{Sing}(w), again one could ask for the Hausdorff dimension of Sing(w)\text{Sing}(w) restricted to a manifold or a fractal.

Following work of several authors, an approach to question 1 was developed by Kleinbock and Weiss in [KWsingular] where it was shown that if μ\mu belongs to a class of measures called friendly measures, then μ(Sing(w))=0\mu(\text{Sing}(w))=0. Friendly measures form a reasonably large class of measures and include some IFS’s as well as pushforwards of Lebesgue measure by smooth maps. There has been recent progress in showing the existence of weighted singular vectors on manifolds, cf. [KMWW, DattaTamam].

The second question, i.e. of estimating the Hausdorff dimension of Sing(w)\text{Sing}(w) turns out to be a difficult problem and has received much attention in the last two decades. In a landmark work [Cheung], Y. Cheung showed that the dimensionof Sing(1/2,1/2)2\text{Sing}(1/2,1/2)\subset\mathbb{R}^{2} is 4/34/3. This was subsequently generalized to n\mathbb{R}^{n} in an important work of Cheung and Chevallier [CheungChevallier]. Another important result was obtained by Kadyrov, Kleinbock, Lindenstrauss and Margulis in [KKLM] using methods from homogeneous dynamics. Namely, a sharp upper bound on the more general set of singular on average m×nm\times n matrices was obtained using integral inequalities as introduced in the famous work [EMM] on the Oppenheim conjecture. The complementary lower bound was obtained by Das, Fishman, Simmons and Urbanski in [DFSU] using methods from the parametric geometry of numbers, see also the recent paper [Solan] for more results in this direction. In [Khalilsing], Khalil upgraded the integral inequality approach of [KKLM] using a beautiful argument to deal with singular vectors on fractals. That is, an upper bound on the Hausdorff dimension of singular vectors lying on self-similar fractals in m\mathbb{R}^{m} that satisfied the open set condition was obtained. In [ShahYang], Shah and Yang obtained bounds for the dimension of certain singular vectors lying on affine subspaces.

All the impressive results towards question 2 discussed above are in the ‘unweighted’ setting. The literature in the more complicated weighted setting is sparser. In [LSST], the Hausdorff dimension of Sing(w)2\text{Sing}(w)\subset\mathbb{R}^{2} was shown to be equal to 211+w12-\frac{1}{1+w_{1}}. As far as higher dimensions are concerned, the authors (page 836 in [LSST]) write that “it is likely” that the dimension of weighted singular vectors in m\mathbb{R}^{m} is m1/λ1m-1/\lambda_{1} “where λ1\lambda_{1} is the top Lyapunov exponent for the adjoint action of the corresponding one-parameter semigroup on the corresponding unipotent group.”

This expectation turns out to be accurate. The lower bound was recently obtained by Kim and Park in [KimPark2024] using an appropriate generalization of the technique of [LSST] that involves the construction of a fractal set contained in Sing(w)\text{Sing}(w). A corollary of our main result provides the matching upper bound, and so together we have

Theorem 1.2.

Fix mm\in\mathbb{N} and w=(w1,,wm)mw=(w_{1},\ldots,w_{m})\in\mathbb{R}^{m} such that w1w2wm>0w_{1}\geq w_{2}\geq\ldots\geq w_{m}>0 and w1++wm=1w_{1}+\ldots+w_{m}=1. Then the Hausdorff dimension of Sing(w)\text{Sing}(w) satisfies

dimH(Sing(w))=m11+w1.\displaystyle\dim_{H}(\text{Sing}(w))=m-\frac{1}{1+w_{1}}. (1)

In fact, our methods also allow us to deal with Diophantine approximation on fractals; providing the very first dimension bounds for weighted singular vectors on fractals. Here is another corollary of our main theorem.

Theorem 1.3.

Fix mm\in\mathbb{N}. Fix w=(w1,,wm)mw=(w_{1},\ldots,w_{m})\in\mathbb{R}^{m} such that w1w2wm>0w_{1}\geq w_{2}\geq\ldots\geq w_{m}>0 and w1++wm=1w_{1}+\ldots+w_{m}=1. For each 1im1\leq i\leq m, let Φi\Phi_{i} be an IFS consisting of contractive similarities with equal contraction ratios on \mathbb{R} and satisfying the open set condition. Let 𝒦i\mathcal{K}_{i} be the limit set of Φi\Phi_{i} and 𝒦=i𝒦i\mathcal{K}=\prod_{i}\mathcal{K}_{i}. Then, the dimension of Sing(w)𝒦\text{Sing}(w)\cap\mathcal{K} satisfies

dimH(Sing(w)𝒦)dimH(𝒦)mini(dimH𝒦i)1+w1,\displaystyle\dim_{H}(\text{Sing}(w)\cap\mathcal{K})\leq\dim_{H}(\mathcal{K})-\frac{\min_{i}(\dim_{H}\mathcal{K}_{i})}{1+w_{1}}, (2)

where s=dim(𝒦)s=dim(\mathcal{K}).

We will now state our main theorem, postponing some notation for later. Our approach is dynamical in nature and owes its existence to an influential work of Dani [Dani-Crelle] where a connection was established between Diophantine properties of vectors and dynamical and topological properties of certain orbits of an associated unimodular lattice. Let 𝒳=SLm+1()/SLm+1()\mathcal{X}=\text{SL}_{m+1}(\mathbb{R})/\text{SL}_{m+1}(\mathbb{Z}) be the space of unimodular lattices in n\mathbb{R}^{n}. For t>0t>0 and θm\theta\in\mathbb{R}^{m}, define

gt=(tw1twmt1),u(θ)=(Imθ1).\displaystyle g_{t}=\begin{pmatrix}t^{w_{1}}\\ &\ddots\\ &&t^{w_{m}}\\ &&&t^{-1}\end{pmatrix},\quad u(\theta)=\begin{pmatrix}I_{m}&\theta\\ &1\end{pmatrix}. (3)

For x𝒳x\in\mathcal{X} define the set

Div(x,w):={θm:gtu(θ)x}.\text{Div}(x,w):=\{\theta\in\mathbb{R}^{m}~{}:~{}g_{t}u(\theta)x\rightarrow\infty\}.

By Dani’s results mentioned above (the famous “Dani correspondence”), singular vectors correspond exactly to the above set for x=eΓx=e\Gamma. The main result of the paper is the following theorem.

Theorem 1.4.

Fix mm\in\mathbb{N}. Fix w=(w1,,wm)mw=(w_{1},\ldots,w_{m})\in\mathbb{R}^{m} such that w1w2wm>0w_{1}\geq w_{2}\geq\ldots\geq w_{m}>0 and w1++wm=1w_{1}+\ldots+w_{m}=1. For each 1im1\leq i\leq m, let Φi\Phi_{i} be an IFS consisting of contractive similarities with equal contraction ratios on \mathbb{R} and satisfying the open set condition. Let 𝒦i\mathcal{K}_{i} be the limit set of Φi\Phi_{i} and 𝒦=i𝒦i\mathcal{K}=\prod_{i}\mathcal{K}_{i}. Then for every x𝒳x\in\mathcal{X},

dimH(Div(x,w)𝒦)dimH(𝒦)mini(dimH𝒦i)1+w1.\displaystyle\dim_{H}(\text{Div}(x,w)\cap\mathcal{K})\leq\dim_{H}(\mathcal{K})-\frac{\min_{i}(\dim_{H}\mathcal{K}_{i})}{1+w_{1}}. (4)

1.1. Outline of proof

The core strategy of the proof closely mirrors the approach developed in [KKLM]. The key idea is to construct a family of height functions that satisfy the contraction hypothesis. These height functions quantify the depth of orbits as they approach the cusp. Once such a family is constructed, it is then used iteratively to generate an open covering of Div(x,w)\text{Div}(x,w). Specifically, by applying the contraction hypothesis iteratively, we construct coverings of Div(x,v)\text{Div}(x,v) using balls of nearly uniform diameter, with the number of balls being bounded by a function of their common diameter. This estimate ensures the finiteness of the ll-dimensional Hausdorff measure of 𝒦\mathcal{K} for specific values of ll smaller than dd, thereby providing an upper bound on Div(x,w)\text{Div}(x,w). This bound corresponds to the one stated in Theorem 1.4.

Despite its apparent simplicity, the approach presents significant challenges. The biggest one is constructing a family of height functions that satisfy the optimal contraction conditions. The optimality is important because the bound for Div(x,w)\text{Div}(x,w) depends very sensitively on the contraction rate of the height functions. Existing techniques can create height functions, but they aren’t optimal unless all the weights are equal. As a result, the best possible bound for m\mathbb{R}^{m}, without the new tools in this paper, is mmwm1+w1m-\frac{mw_{m}}{1+w_{1}}, which only matches the actual bound when mwm=1mw_{m}=1, meaning all the weights are equal (where w1w2wmw_{1}\geq w_{2}\geq\dots\geq w_{m}).

To address this, we introduce some new concepts. First, we redefine the covolume of sublattices to depend on weights, as done in Section 3. Using this new definition of covolume .\|.\|, we as usual define the following proper functions on 𝒳\mathcal{X}:

φl:𝒳,Λmax{1Λl:Λl is a primitive sublattice of Λ of rank l}.\varphi_{l}:\mathcal{X}\rightarrow\mathbb{R},\quad\Lambda\mapsto\max\left\{\frac{1}{\|\Lambda_{l}\|}:\Lambda_{l}\text{ is a primitive sublattice of $\Lambda$ of rank $l$}\right\}.

In the unweighted case, the linear combination of these φl\varphi_{l} acts as a height function, and one of the essential ingredients in proving that the linear combination is indeed a height function is a result from [EMM], Lemma 5.6, which states that for any two sublattices Λ1,Λ2\Lambda_{1},\Lambda_{2} of Λ\Lambda and with usual definition of covolume .\|.\|, we have

Λ1Λ2Λ1+Λ2Λ1Λ2.\|\Lambda_{1}\cap\Lambda_{2}\|\|\Lambda_{1}+\Lambda_{2}\|\leq\|\Lambda_{1}\|\|\Lambda_{2}\|.

This inequality is used to link the average rate of expansion of vectors in certain representations to recurrence results on 𝒳\mathcal{X}, which leads to the height function. The most crucial step of the paper is to prove an analogue of this result using the redefined covolume. Although this may seem straightforward, the new covolume definition introduces complications. For instance, it is not even true that for any gGg\in G, one can find a constant Cg>0C_{g}>0 such that

gΛCgΛ,for all rank l sublattices Λ in m+1.\|g\Lambda\|\leq C_{g}\|\Lambda\|,\quad\text{for all rank }l\text{ sublattices }\Lambda\text{ in }\mathbb{R}^{m+1}.

The analogous key inequality is referred to as the “weighted Mother inequality” and is discussed in Section 3. Section 4 proves the optimal expansion rate of vectors in the exterior powers of the standard representation of SLm+1()\text{SL}_{m+1}(\mathbb{R}) with respect to the quasinorm based on weights. Section 5 constructs the height functions using results from Section 4 and the weighted Mother inequality.

The second challenge lies in deriving dimension bounds from the contraction hypothesis. This is done iteratively. In the equal weight case, for [0,1]m[0,1]^{m}, the process is as follows: Start with [0,1]m[0,1]^{m} and divide it into 2m2^{m} parts, i.e., [i1/2,(i1+1)/2]××[im/2,(im+1)/2][i_{1}/2,(i_{1}+1)/2]\times\dots\times[i_{m}/2,(i_{m}+1)/2]. Applying the contraction hypothesis with the base point xx, at most 2γ2^{\gamma} intervals are needed to cover Div(x,w)\text{Div}(x,w).

Now, pick one of the \ell_{\infty}-balls that intersect Div(x,w)\text{Div}(x,w), say [i1/2,(i1+1)/2]××[im/2,(im+1)/2][i_{1}/2,(i_{1}+1)/2]\times\dots\times[i_{m}/2,(i_{m}+1)/2], and divide it into 2m2^{m} smaller parts. By applying the contraction hypothesis again with base point g2m/(1+m)u(i1/2,,im/2)xg_{2^{m/(1+m)}}u(i_{1}/2,\dots,i_{m}/2)x, and using the fact that

{g2m/(1+m)u(θ)g2m/(1+m)u(i1/2,,im/2):θ[0,1]m}\displaystyle\{g_{2^{m/(1+m)}}u(\theta)g_{2^{m/(1+m)}}u(i_{1}/2,\dots,i_{m}/2):\theta\in[0,1]^{m}\}
={g22m/(1+m)u(θ):θ[i1/2,(i1+1)/2]××[im/2,(im+1)/2]}\displaystyle=\{g_{2^{2m/(1+m)}}u(\theta):\theta\in[i_{1}/2,(i_{1}+1)/2]\times\dots\times[i_{m}/2,(i_{m}+1)/2]\}

and that we know the point g22m/(1+m)u(i1/2,,im/2)xg_{2^{2m/(1+m)}}u(i_{1}/2,\dots,i_{m}/2)x is near the cusp (since Div(x,w)[i1/2,(i1+1)/2]××[im/2,(im+1)/2]\text{Div}(x,w)\cap[i_{1}/2,(i_{1}+1)/2]\times\dots\times[i_{m}/2,(i_{m}+1)/2]\neq\emptyset), we conclude that at most 2γ2^{\gamma} intervals are needed to cover Div(x,w)[i1/2,(i1+1)/2]××[im/2,(im+1)/2]\text{Div}(x,w)\cap[i_{1}/2,(i_{1}+1)/2]\times\dots\times[i_{m}/2,(i_{m}+1)/2]. Thus, 22γ2^{2\gamma} \ell_{\infty}-balls of diameter 1/221/2^{2} are needed to cover Div(x,w)\text{Div}(x,w). Repeating this process iteratively, we find that 2kγ2^{k\gamma} \ell_{\infty}-balls of diameter 1/2k1/2^{k} are needed to cover Div(x,w)\text{Div}(x,w), which gives the bound dimH(Div(x,w))γ\dim_{H}(\text{Div}(x,w))\leq\gamma.

In the equal weight case, this process is simpler because the action of gtg_{t} (as defined in (3) for w1==wmw_{1}=\dots=w_{m}) expands the interval [0,1]m[0,1]^{m} evenly. This allows us to easily subdivide it into cubes like [i1/2N,(i1+1)/2N]××[im/2N,(im+1)/2N][i_{1}/2^{N},(i_{1}+1)/2^{N}]\times\dots\times[i_{m}/2^{N},(i_{m}+1)/2^{N}] and conjugate them with g2m/(m+1)g_{2^{m/(m+1)}}, making the integral over these cubes behave like that over [0,1]m[0,1]^{m}. However, in the unequal expansion case, such simple divisions of [0,1]m[0,1]^{m} are no longer possible. Instead, the space must be divided into hypercuboids with unequal side lengths to ensure that they become [0,1]m[0,1]^{m} under conjugation. We must also ensure that these hypercuboids are disjoint, since we rely on the small measure of their union to conclude that the number of hypercuboids needed for covering is small. These two restrictions together are difficult to satisfy. This issue becomes even more challenging when dealing with fractals. Overcoming this obstacle is the second main contribution of this paper, discussed in Section 6.

2. Notation

The following notation will be used throughout the paper.

2.1. Homogeneous spaces

Fix mm\in\mathbb{N} and set G=SLm+1(),Γ=SLm+1()G=\text{SL}_{m+1}(\mathbb{R}),\Gamma=\text{SL}_{m+1}(\mathbb{Z}) and 𝒳=G/Γ\mathcal{X}=G/\Gamma. The space 𝒳\mathcal{X} can be naturally identified with the space of unimodular lattices in m+1\mathbb{R}^{m+1}, via the identification AΓAdA\Gamma\mapsto A\mathbb{Z}^{d}.

Fix w=(w1,,wm)mw=(w_{1},\ldots,w_{m})\in\mathbb{R}^{m} such that w1w2wm>0w_{1}\geq w_{2}\geq\ldots\geq w_{m}>0 and w1++wm=1w_{1}+\ldots+w_{m}=1. Define wm+1:=w1w_{m+1}:=w_{1} and set

wI=iIwi\displaystyle w_{I}=\sum_{i\in I}w_{i} (5)

for any subset II of {1,m+1}\{1,\ldots m+1\}. For t>0t>0, define

gt=(tw1twmt1),at=(etw1etwm+1).\displaystyle g_{t}=\begin{pmatrix}t^{w_{1}}\\ &\ddots\\ &&t^{w_{m}}\\ &&&t^{-1}\end{pmatrix},\quad a_{t}=\begin{pmatrix}e^{tw_{1}}\\ &\ddots\\ &&e^{tw_{m+1}}\end{pmatrix}. (6)

As in the Introduction, we define for θm\theta\in\mathbb{R}^{m}, the matrix

u(θ)=(Imθ1).\displaystyle u(\theta)=\begin{pmatrix}I_{m}&\theta\\ &1\end{pmatrix}. (7)

We define HH to be the following subgroup of GG,

H={(t1x1tmxm(t1tm)1):(t1,,tm)+m,(x1,,xm)m}.\displaystyle H=\left\{\begin{pmatrix}t_{1}&&&x_{1}\\ &\ddots&&\vdots\\ &&t_{m}&x_{m}\\ &&&(t_{1}\ldots t_{m})^{-1}\end{pmatrix}:(t_{1},\ldots,t_{m})\in\mathbb{R}_{+}^{m},(x_{1},\ldots,x_{m})\in\mathbb{R}^{m}\right\}. (8)

2.2. Iterated Function Systems

A contracting similarity is a map \mathbb{R}\rightarrow\mathbb{R} of the form xcx+yx\mapsto cx+y where c(0,1)c\in(0,1), and ydy\in\mathbb{R}^{d}. A finite similarity Iterated Function System with constant ratio (IFS) on \mathbb{R} is a collection of contracting similarities Φ=(ϕe:)eE\Phi=(\phi_{e}:\mathbb{R}\rightarrow\mathbb{R})_{e\in E} indexed by a finite set EE, called the alphabet, such that there exists a constant c(0,1)c\in(0,1) independent of ee so that

ϕe(x)=cx+we,\phi_{e}(x)=cx+w_{e},

for all eEe\in E.

Let B=EB=E^{\mathbb{N}}. The coding map of an IFS Φ\Phi is the map τ:B\tau:B\rightarrow\mathbb{R} defined by the formula

τ(b)=limlϕb1ϕbl(0).\displaystyle\tau(b)=\lim_{l\rightarrow\infty}\phi_{b_{1}}\circ\ldots\circ\phi_{b_{l}}(0). (9)

It is well known that the limit in (9)\eqref{eq:def tau} exists and that the coding map is continuous. The image of BB under the coding map called the limit set of Φ\Phi, is a compact subset of \mathbb{R}, which we denote by 𝒦=𝒦(Φ)\mathcal{K}=\mathcal{K}(\Phi). We define for e~=(e1,,el)Ep{\tilde{e}}=(e_{1},\ldots,e_{l})\in E^{p},

𝒦e~=ϕe1ϕel(𝒦),(l)={𝒦e~:e~El}.\displaystyle\mathcal{K}_{\tilde{e}}=\phi_{e_{1}}\circ\ldots\circ\phi_{e_{l}}(\mathcal{K}),\quad\mathcal{F}(l)=\{\mathcal{K}_{{\tilde{e}}^{\prime}}:{\tilde{e}}^{\prime}\in E^{l}\}. (10)

We will say that Φ\Phi satisfies an open set condition (OSC for short) if there exist a non-empty open subset UU\subset\mathbb{R} such that the following holds

ϕe(U)U\displaystyle\phi_{e}(U)\subset U for every eE\displaystyle\text{ for every }e\in E
ϕe(U)ϕe(U)=,\displaystyle\phi_{e}(U)\cap\phi_{e^{\prime}}(U)=\emptyset, for every eeE.\displaystyle\text{ for every }e\neq e^{\prime}\in E.

Let Prob(E)\text{Prob}(E) denote the space of probability measures on EE. For each νProb(E)\nu\in\text{Prob}(E) we can consider the measure ην\eta_{*}\nu^{\otimes\mathbb{N}} under the coding map. A measure of the form ην\eta_{*}\nu^{\otimes\mathbb{N}} is called a Bernoulli measure.

The following proposition is well known (see for eg [[Hutchinson], Thm. 5.3(1)] for a proof).

Proposition 2.1.

Suppose Φ={ϕe:eE}\Phi=\{\phi_{e}:e\in E\} is an IFS satisfying the open set condition with limit set 𝒦\mathcal{K}. Let cc denote the common contraction ratio of (ϕe)eE(\phi_{e})_{e\in E} and p=#Ep=\#E. Then the Hausdorff dimension ss of 𝒦\mathcal{K} is logp/logc-\log p/\log c. Moreover the ss-dimensional Hausdorff measure HsH^{s} satisfies 0<Hs(𝒦)<10<H^{s}(\mathcal{K})<1. If μ\mu denotes the normalised restriction of HsH^{s} to 𝒦\mathcal{K}, then μ\mu is a Bernoulli measure and equals ην\eta_{*}\nu^{\otimes\mathbb{N}}, where ν\nu is the uniform measure on EE, i.e, ν(F)=#F/#E\nu(F)=\#F/\#E for all FEF\subset E. Moreover, there exists a constant β>0\beta>0 such that for all xx\in\mathbb{R}, we have

μ([xr,x+r])βrs.\displaystyle\mu([x-r,x+r])\leq\beta r^{s}. (11)

For the rest of the paper, we fix for 1im1\leq i\leq m, the IFS Φi={ϕi,e:eEi}\Phi_{i}=\{\phi_{i,e}:e\in E_{i}\} with common contraction ratio cic_{i} and the limit set 𝒦i\mathcal{K}_{i}. Let pi=#Eip_{i}=\#E_{i}, si=logpi/logcis_{i}=-\log p_{i}/\log c_{i}, 𝒦=i𝒦i\mathcal{K}=\prod_{i}\mathcal{K}_{i} and s=isis=\sum_{i}s_{i}. Let μi\mu_{i} denote the normalised restriction of HsiH^{s_{i}} to 𝒦i\mathcal{K}_{i} and define the measure μ=iμi\mu=\otimes_{i}\mu_{i} on 𝒦\mathcal{K}. Let us define the constant η\eta as

η\displaystyle\eta =min{si:i{1,,m}}.\displaystyle=\min\{s_{i}:i\in\{1,\ldots,m\}\}. (12)

An observations that will be needed later in the proof is that

ηiIsiwI{i},\displaystyle\eta\geq\sum_{i\notin I}s_{i}w_{I\cup\{i\}}, (13)

for all I{1,,m}I\subsetneq\{1,\ldots,m\}.

3. Weighted Mother Inequality

Let us define

Vl=lm+1,V=l=1m+1Vl.V_{l}=\bigwedge^{l}\mathbb{R}^{m+1},\qquad V=\bigoplus_{l=1}^{m+1}V_{l}.

Define action of GG on VV via the map gl=1m+1lgg\mapsto\bigoplus_{l=1}^{m+1}\bigwedge^{l}g. Suppose {e1,,em+1}\{e_{1},\ldots,e_{m+1}\} denote the standard basis of m+1\mathbb{R}^{m+1}. For each index set I={i1<<il}{1,,m+1}I=\{i_{1}<\cdots<i_{l}\}\subset\{1,\dots,m+1\}, we define

eI:=ei1𝐞il.\displaystyle{e}_{I}:={e}_{i_{1}}\wedge\cdots\wedge\mathbf{e}_{i_{l}}. (14)

The collection of monomials 𝐞I\mathbf{e}_{I} gives a basis of Vl=lm+1V_{l}=\bigwedge^{l}\mathbb{R}^{m+1} for each 1lm+11\leq l\leq m+1. For vVv\in V and each index set II, we denote by vIv_{I}\in\mathbb{R}, the unique value so that v=JvJeJv=\sum_{J}v_{J}e_{J}, where the sum is taken over all index sets JJ.

For each ll, we define quasi-norm .\|.\| on each of VlV_{l} as

v=maxI|vI|1wI,\displaystyle\|v\|=\max_{I}|v_{I}|^{\frac{1}{w_{I}}}, (15)

where maximum is taken oven all index sets II of cardinality ll and wIw_{I} is defined as in (5). Note that for all tt\in\mathbb{R} and for all vVlv\in V_{l} (1lm+11\leq l\leq m+1), we have

atv=etv.\displaystyle\|a_{t}v\|=e^{t}\|v\|. (16)
Definition 3.1.

For a discrete subgroup Λ\Lambda of m+1\mathbb{R}^{m+1} of rank l1l\geq 1, we define vΛVl/{±1}v_{\Lambda}\in V_{l}/\{\pm 1\} as v1vlv_{1}\wedge\ldots\wedge v_{l}, where v1,,vlv_{1},\ldots,v_{l} is a \mathbb{Z}-basis of Λ\Lambda. Note that the definition of vΛv_{\Lambda} is independent of the choice of basis v1,,vlv_{1},\ldots,v_{l}. We define Λ\|\Lambda\| as

Λ=v,\displaystyle\|\Lambda\|=\|v\|, (17)

where .\|.\| on VlV_{l} is defined as in (15). We also define {0}=1\|\{0\}\|=1.

The main result of this section is the following inequality which originated in [EMM]. Part of this proof is motivated from [[BQ12], Proof of Prop. 3.1] where it is termed the ‘Mother inequality’.

Theorem 3.2.

There exist a constant D>0D>0 and a finite set {0}S[0,1)\{0\}\subset S\subset[0,1) such that the following holds. Fix a lattice Λ\Lambda in m+1\mathbb{R}^{m+1}. Then for any sublattices Λ1,Λ2\Lambda_{1},\Lambda_{2} of Λ\Lambda, we have

min{Λ1Λ2aΛ1+Λ21a:aS}Dmax{Λ1aΛ21a:0<a<1},\displaystyle\min\{\|\Lambda_{1}\cap\Lambda_{2}\|^{a}\|\Lambda_{1}+\Lambda_{2}\|^{1-a}:a\in S\}\leq D\max\{\|\Lambda_{1}\|^{a}\|\Lambda_{2}\|^{1-a}:0<a<1\}, (18)

where Λ1+Λ2\Lambda_{1}+\Lambda_{2} denotes the smallest discrete subgroup of m+1\mathbb{R}^{m+1} containing Λ1\Lambda_{1} and Λ2\Lambda_{2}.

Proof.

The proof is divided into two cases:

Case 11: Λ1Λ2={0}\Lambda_{1}\cap\Lambda_{2}=\{0\}. In this case, (18) follows if the following holds: For any pVip\in V_{i} and qVjq\in V_{j}, the following holds:

pq((i+j)!i!j!)mmax{paq1a:0<a<1}.\displaystyle\|p\wedge q\|\leq\left(\frac{(i+j)!}{i!j!}\right)^{m}\max\{\|p\|^{a}\|q\|^{1-a}:0<a<1\}. (19)

To prove (19), note that

pq\displaystyle\|p\wedge q\| =max#I=i+j|(pq)I|1wI\displaystyle=\max_{\#I=i+j}|(p\wedge q)_{I}|^{\frac{1}{w_{I}}} (20)
max#I=i+j(#J=i,#K=jJK=I|pJ||qK|)1wI\displaystyle\leq\max_{\#I=i+j}\left(\sum_{\begin{subarray}{c}\#J=i,\#K=j\\ J\cup K=I\end{subarray}}|p_{J}||q_{K}|\right)^{\frac{1}{w_{I}}} (21)
((i+j)!i!j!)mmax#I=i+j(max#J=i,#K=jJK=I|pJ||qK|)1wI\displaystyle\leq\left(\frac{(i+j)!}{i!j!}\right)^{m}\max_{\#I=i+j}\left(\max_{\begin{subarray}{c}\#J=i,\#K=j\\ J\cup K=I\end{subarray}}|p_{J}||q_{K}|\right)^{\frac{1}{w_{I}}} (22)
=((i+j)!i!j!)mmax#I=i+jmax#J=i,#K=jJK=I(|pJ|1wJ)wJwJ+wK(|wK|1wK)wKwJ+wKusing fact that wI=wJ+wK\displaystyle=\left(\frac{(i+j)!}{i!j!}\right)^{m}\max_{\#I=i+j}\max_{\begin{subarray}{c}\#J=i,\#K=j\\ J\cup K=I\end{subarray}}(|p_{J}|^{\frac{1}{w_{J}}})^{\frac{w_{J}}{w_{J}+w_{K}}}(|w_{K}|^{\frac{1}{w_{K}}})^{\frac{w_{K}}{w_{J}+w_{K}}}\quad\text{using fact that }w_{I}=w_{J}+w_{K} (23)
((i+j)!i!j!)mmax#I=i+jmax#J=i,#K=jJK=I(p)wJwJ+wKqwKwJ+wK\displaystyle\leq\left(\frac{(i+j)!}{i!j!}\right)^{m}\max_{\#I=i+j}\max_{\begin{subarray}{c}\#J=i,\#K=j\\ J\cup K=I\end{subarray}}(\|p\|)^{\frac{w_{J}}{w_{J}+w_{K}}}\|q\|^{\frac{w_{K}}{w_{J}+w_{K}}} (24)
((i+j)!i!j!)mmax{paq1a:0<a<1}.\displaystyle\leq\left(\frac{(i+j)!}{i!j!}\right)^{m}\max\{\|p\|^{a}\|q\|^{1-a}:0<a<1\}. (25)

Hence (19) follows.

Case 22: Λ1Λ2{0}\Lambda_{1}\cap\Lambda_{2}\neq\{0\}. In this case, (18) follows if the following holds: For any p=p1piVip=p_{1}\wedge\ldots\wedge p_{i}\in V_{i}, q=q1qjVjq=q_{1}\wedge\ldots\wedge q_{j}\in V_{j} and r=r1rkVkr=r_{1}\wedge\ldots\wedge r_{k}\in V_{k}, the following holds:

min{papqr1a:aS}Ci,j,k{pqapr1a:0<a<1}.\displaystyle\min\{\|p\|^{a}\|p\wedge q\wedge r\|^{1-a}:a\in S\}\leq C_{i,j,k}\{\|p\wedge q\|^{a}\|p\wedge r\|^{1-a}:0<a<1\}. (26)

To prove this, we define the linear map ψ:Vi+jVi+kViVi+j+k\psi:V_{i+j}\otimes V_{i+k}\rightarrow V_{i}\otimes V_{i+j+k} such that for any x=x1xi+jx=x_{1}\wedge\ldots\wedge x_{i+j} and y=y1yi+ly=y_{1}\wedge\ldots\wedge y_{i+l},

ψ(xy)=#I=iεIxIxIcy,\psi(x\otimes y)=\sum_{\#I=i}\varepsilon_{I}x_{I}\otimes x_{I^{c}}\wedge y,

where the sum is taken over all subsets I{1,,i+j}I\subset\{1,\ldots,i+j\} of size ii. Let us explain each term in the sum.

  • the element xIx_{I} is the exterior product xI=xl1xlrx_{I}=x_{l_{1}}\wedge\ldots\wedge x_{l_{r}}, when one writes I={l1,,lr}I=\{l_{1},...,l_{r}\} with i1<···<iri_{1}<\textperiodcentered\textperiodcentered\textperiodcentered<i_{r}.

  • the element xIcx_{I^{c}} is the exterior product xIc=xli+1xli+jx_{I^{c}}=x_{l_{i+1}}\wedge\ldots\wedge x_{l_{i+j}}, when one writes I={li+1,,li+j}I=\{l_{i+1},...,l_{i+j}\} with li+1<···<li+jl_{i+1}<\textperiodcentered\textperiodcentered\textperiodcentered<l_{i+j}.

  • the sign εI\varepsilon_{I} is the signature of the permutation of {1,,i+j}\{1,...,i+j\} sending zz to lzl_{z} for 1pi+j1\leq p\leq i+j.

For 1l,lm+11\leq l,l^{\prime}\leq m+1, we define quasi-norms .\|.\| on VlVlV_{l}\otimes V_{l^{\prime}} as

x=max#I=l,#J=l|xI,J|1wI+wJ,\displaystyle\|x\|=\max_{\#I=l,\#J=l^{\prime}}|x_{I,J}|^{\frac{1}{w_{I}+w_{J}}}, (27)

where x=#I=l,#J=lxI,JeIeJx=\sum_{\#I=l,\#J=l^{\prime}}x_{I,J}e_{I}\otimes e_{J}. Let

Sl,l={wIwI+wJ:#I=l,#J=l}.S_{l,l^{\prime}}=\{\frac{w_{I}}{w_{I}+w_{J}}:\#I=l,\#J=l^{\prime}\}.

Then it is easy to see that for every xVlx\in V_{l} and yVly\in V_{l^{\prime}}, we have aSl,la\in S_{l,l^{\prime}} such that

pq=paq1a.\displaystyle\|p\wedge q\|=\|p\|^{a}\|q\|^{1-a}. (28)

Let S=l,lSl,lS^{\prime}=\cup_{l,l^{\prime}}S_{l,l^{\prime}} and S=S{0}S=S^{\prime}\cup\{0\}. It is clear that SS is a finite set, not containing 11. Also note that 0S0\notin S^{\prime}.

Note that for 1l,lm+11\leq l,l^{\prime}\leq m+1, the action of ata_{t} (defined as in (6)) on VlVlV_{l}\otimes V_{l^{\prime}} via the map latlat\wedge^{l}a_{t}\otimes\wedge^{l^{\prime}}a_{t} satisfies

atw=etw.\displaystyle\|a_{t}w\|=e^{t}\|w\|. (29)

Also for all xVi+jx\in V_{i+j}, yVi+ky\in V_{i+k}, we have

ψ(at(xy))=atψ(xy).\displaystyle\psi(a_{t}(x\otimes y))=a_{t}\psi(x\otimes y). (30)

We define

Ci,j,k=sup{ψ(x):xVi+jVi+k,x=1},\displaystyle C_{i,j,k}=\sup\{\|\psi(x)\|:x\in V_{i+j}\otimes V_{i+k},\|x\|=1\}, (31)

which is a finite value since ψ\psi is continuous and the set {xVi+jVi+k:x=1}\{x\in V_{i+j}\otimes V_{i+k}:\|x\|=1\} is a compact set.

Fix p=p1piVip=p_{1}\wedge\ldots\wedge p_{i}\in V_{i}, q=q1qjVjq=q_{1}\wedge\ldots\wedge q_{j}\in V_{j} and r=r1rkVkr=r_{1}\wedge\ldots\wedge r_{k}\in V_{k}. Note that (26) hold trivially if pq=0\|p\wedge q\|=0 or pr=0\|p\wedge r\|=0. Hence, we may assume pqpr0\|p\wedge q\otimes p\wedge r\|\neq 0 and let s=logpqprs=\log\|p\wedge q\otimes p\wedge r\|. Then, we have using (29) that

as(pqpr)=1.\displaystyle\|a_{-s}(p\wedge q\otimes p\wedge r)\|=1. (32)

Note that

ψ(pqpr)=pqpr.\displaystyle\psi(p\wedge q\otimes p\wedge r)=p\otimes q\wedge p\wedge r. (33)

Hence, we have

min{papqr1a:aS}\displaystyle\min\{\|p\|^{a}\|p\wedge q\wedge r\|^{1-a}:a\in S\} ppqr\displaystyle\leq\|p\otimes p\wedge q\wedge r\| using (28)
=pqpr\displaystyle=\|p\otimes q\wedge p\wedge r\|
=ψ(pqpr)\displaystyle=\|\psi(p\wedge q\otimes p\wedge r)\| using (33)
=esasψ(pqpr)\displaystyle=e^{s}\|a_{-s}\psi(p\wedge q\otimes p\wedge r)\| using (29)
=esψ(as(pqpr))\displaystyle=e^{s}\|\psi(a_{-s}(p\wedge q\otimes p\wedge r))\| using (30)
esCi,j,k\displaystyle\leq e^{s}C_{i,j,k} using (31), (32)
=esCi,j,kas(pqpr)\displaystyle=e^{s}C_{i,j,k}\|a_{-s}(p\wedge q\otimes p\wedge r)\| using (32)
=Ci,j,kpqpr\displaystyle=C_{i,j,k}\|p\wedge q\otimes p\wedge r\| using (29)
Ci,j,kmax{paq1a:0<a<1}\displaystyle\leq C_{i,j,k}\max\{\|p\|^{a}\|q\|^{1-a}:0<a<1\} using (28).\displaystyle\text{using \eqref{eq: w 2}}.

Thus theorem now follows by taking any DD larger than all the constants appearing above. ∎

4. Critical Exponent

For gGg\in G, we define

gVl:=sup{gv:vVl,v=1}.\displaystyle\|g\|_{V_{l}}:=\sup\left\{\|gv\|:v\in V_{l},\|v\|=1\right\}. (34)

Also, for any compact subset QHQ\subset H (defined as in (8)), define the set Q~={ashas:s>0,hQ}\tilde{Q}=\{a_{s}ha_{-s}:s>0,h\in Q\}. It is easy to see that Q~\tilde{Q} is still a compact set. Define

Q=sup{gvVl,g1vVl:1lm,gQ~,vVl,vl=1},\displaystyle\|{Q}\|=\sup\{\|gv\|_{V_{l}},\|g^{-1}v\|_{V_{l}}:1\leq l\leq m,g\in\tilde{Q},v\in V_{l},\|v_{l}\|=1\}, (35)
Lemma 4.1.

Fix 1lm1\leq l\leq m. Then for all gHg\in H, vVlv\in V_{l} with v1\|v\|\leq 1, we have

gv{g}v\|gv\|\leq\|\{g\}\|\|v\|
Proof.

Let t=logv>0t=-\log\|v\|>0. Then by (16), we have atv=1\|a_{t}v\|=1. Thus we have

gv\displaystyle\|gv\| =etatgatatvusing (16)\displaystyle=e^{-t}\|a_{t}ga_{-t}a_{t}v\|\quad\text{using \eqref{eq: |||| is a t invariant}}
etatgatVl\displaystyle\leq e^{-t}\|a_{t}ga_{-t}\|_{V_{l}}
v{g}.\displaystyle\leq\|v\|\|\{g\}\|.

Recall the following well-known fact.

Lemma 4.2.

There exists c1c\geq 1 such that the following holds for all 1lm1\leq l\leq m and all v,wVlv,w\in V_{l}

v+wc(v+w)\|v+w\|\leq c(\|v\|+\|w\|)

The following lemma is a simple application of Fubini’s Theorem.

Lemma 4.3.

Let ν\nu be a Borel measure and ff a non-negative Borel function on a separable metric space XX. Then,

Xf𝑑ν=0ν({xX:f(x)R})𝑑R\int_{X}f\;d\nu=\int_{0}^{\infty}\nu\left(\{x\in X:f(x)\geq R\}\right)\;dR

The following Proposition studies the average rate of expansion of vectors in the exterior powers of m+1\mathbb{R}^{m+1}, and is the main result in this section.

Proposition 4.4.

Fix 1lm1\leq l\leq m. For all 0<ρ<10<\rho<1, there exists C=C(ρ)>1C=C(\rho)>1 such that the following holds for all vVl{0}v\in V_{l}\setminus\{0\},

𝒦gtu(x)vηρ𝑑μ(x)Ctηρvηρ.\int_{\mathcal{K}}\|g_{t}u(x)v\|^{-\eta\rho}\,d\mu(x)\leq Ct^{-\eta\rho}\|v\|^{-\eta\rho}.
Proof.

Fix 1lm1\leq l\leq m, 0<ρ<10<\rho<1 and vVlv\in V_{l}. Without loss of generality, we may assume that v=1\|v\|=1. Consider the vector space Vl+VlV_{l}^{+}\subset V_{l} spanned by {eI:I{1,,d}}\{e_{I}:I\subset\{1,\ldots,d\}\}. Also consider the vector space VlV_{l}^{-} spanned by {eI:I{1,,d}}\{e_{I}:I\nsubseteq\{1,\ldots,d\}\} We will denote by π+:Vl=Vl+VlVl+\pi_{+}:V_{l}=V_{l}^{+}\oplus V_{l}^{-}\rightarrow V_{l}^{+} the canonical projection. Similarly define π:VlVl\pi_{-}:V_{l}\rightarrow V_{l}^{-}. Note that for all vVlv\in V_{l}, we have

v=max{π+(v),π(v)}\|v\|=\max\{\|\pi_{+}(v)\|,\|\pi_{-}(v)\|\}

Then we have

𝒦gtu(x)vηρ𝑑μ(x)𝒦π+(gτ(x)u(x)v)ηρ𝑑μ(x)=𝒦tηρπ+(u(x)v)ηρ𝑑μ(x).\displaystyle\int_{\mathcal{K}}\|g_{t}u(x)v\|^{-\eta\rho}\,d\mu(x)\leq\int_{\mathcal{K}}\|\pi_{+}(g_{\tau(x)}u(x)v)\|^{-\eta\rho}\,d\mu(x)=\int_{\mathcal{K}}t^{-\eta\rho}\|\pi_{+}(u(x)v)\|^{-\eta\rho}\,d\mu(x).

Thus, it is enough to prove that the quantity 𝒦π+(u(x)v)ηρ𝑑μ(x)\int_{\mathcal{K}}\|\pi_{+}(u(x)v)\|^{-\eta\rho}\,d\mu(x) is bounded above by a constant independent of vv.

To do this, we define K={u(x):x𝒦}K=\|\{u(x):x\in\mathcal{K}\}\|, where .\|.\| is defined as in (35). Since μ\mu is compactly supported, we have that K<K<\infty.

Case 1 π(v)<1/3Kc)\|\pi_{-}(v)\|<1/3Kc). Then, since v=1\|v\|=1, we have π+(v)>2/3\|\pi_{+}(v)\|>2/3. This means that

π+(u(x)v)\displaystyle\|\pi_{+}(u(x)v)\| c1π+(u(x)π+(v))π+(u(x)π(v))using Lemma 4.2\displaystyle\geq c^{-1}\|\pi_{+}(u(x)\pi_{+}(v))\|-\|\pi_{+}(u(x)\pi_{-}(v))\|\quad\text{using Lemma \ref{lem: triangle inequality}}
c1π+(v)u(x)π(v)\displaystyle\geq c^{-1}\|\pi_{+}(v)\|-\|u(x)\pi_{-}(v)\|
23cK.13Kcusing Lemma 4.1 and fact that π(v)1\displaystyle\geq\frac{2}{3c}-K.\frac{1}{3Kc}\quad\text{using Lemma \ref{lem: operator norm make sense} and fact that $\|\pi_{-}(v)\|\leq 1$}
=13c.\displaystyle=\frac{1}{3c}.

Thus, 𝒦π+(u(x)v)ηρ𝑑μ(x)(3c)ηρ\int_{\mathcal{K}}\|\pi_{+}(u(x)v)\|^{-\eta\rho}\,d\mu(x)\leq(3c)^{-\eta\rho}.

Case 2 π(v)>1/3Kc\|\pi_{-}(v)\|>1/3Kc. In this case, there exists a subset {m+1}I{1,,m+1}\{m+1\}\subset I\subset\{1,\ldots,m+1\} of cardinality ll such that vI>1/Kv_{I}>1/K^{\prime}, where K=max{(3cK)wI:{m+1}I{1,,m+1},#I=l}K^{\prime}=\max\{(3cK)^{w_{I}}:\{m+1\}\subset I\subset\{1,\ldots,m+1\},\#I=l\}. Let us define for all iIi\notin I, the set Ii={I{i}{m+1}}I_{i}=\{I\cup\{i\}\setminus\{m+1\}\}

Let us define

E(v,R)\displaystyle E(v,R) ={xm:π+(u(x)v)<R},\displaystyle=\{x\in\mathbb{R}^{m}:\|\pi_{+}(u(x)v)\|<R\},
E(v,R)\displaystyle E^{\prime}(v,R) ={xm:for all iI, we have (π+(u(x)v))Ii<R}.\displaystyle=\{x\in\mathbb{R}^{m}:\text{for all $i\notin I$, we have }(\pi_{+}(u(x)v))_{I_{i}}<R\}.

It is clear that E(v,R)E(v,R)E(v,R)\subset E^{\prime}(v,R). Claim that for all R<1R<1, we have

μ(E(v,R))βm+1l(K)inIsiRη,\displaystyle\mu(E^{\prime}(v,R))\leq\beta^{m+1-l}(K^{\prime})^{\sum_{i\not inI}s_{i}}R^{\eta}, (36)

where β\beta is large enough so that (11) holds for all μi\mu_{i} (1im1\leq i\leq m). To see this, first of all, note that by explicit computation, we have for every J{1,,m}J\subset\{1,\ldots,m\}, the following holds

(π+(u(x)v))J=vJ+jJej,JvJ{m+1}{j}xj,(\pi_{+}(u(x)v))_{J}=v_{J}+\sum_{j\in J}e_{j,J}v_{J\cup\{m+1\}\setminus\{j\}}x_{j},

for some ej,J{±1}e_{j,J}\in\{\pm 1\}. Thus for all iIi\notin I, we have

(π+(u(x)v))Ii=eIi,ivIxi+(vIi+jI{m+1}ej,IivIi{m+1}{j}xj),(\pi_{+}(u(x)v))_{I_{i}}=e_{I_{i},i}v_{I}x_{i}+\left(v_{I_{i}}+\sum_{j\in I\setminus\{m+1\}}e_{j,I_{i}}v_{I_{i}\cup\{m+1\}\setminus\{j\}}x_{j}\right),

Thus if we fix xjx_{j} for jI{m+1}j\in I\setminus\{m+1\}, then for all iIi\notin I and R>0R>0, we have that the condition |(π+(u(x)v))Ii|1/wIi<R|(\pi_{+}(u(x)v))_{I_{i}}|^{1/{w_{I_{i}}}}<R is equivalent to the condition that xix_{i} belongs to an interval of size atmost RwIi/|vI|<KRwIiR^{w_{I_{i}}}/|v_{I}|<K^{\prime}R^{w_{I_{i}}}, which has μi\mu_{i} measure less than β(KRwIi)si\beta(K^{\prime}R^{w_{I_{i}}})^{s_{i}} (using (11)). Hence, by Fubini’s Theorem, we get that

μ(E(v,R))βm+1l(K)iIsiRiIsiwIi.\displaystyle\mu(E^{\prime}(v,R))\leq\beta^{m+1-l}(K^{\prime})^{\sum_{i\notin I}s_{i}}R^{\sum_{i\notin I}s_{i}w_{I_{i}}}. (37)

From (13), we know that iIsiwIiη\sum_{i\notin I}s_{i}w_{I_{i}}\geq\eta. Coming this with the fact that R<1R<1, we get (36) follows from (37).

Now using Lemma 4.3, we get that

𝒦π+(u(x)v)ηρ𝑑μ(x)\displaystyle\int_{\mathcal{K}}\|\pi_{+}(u(x)v)\|^{-\eta\rho}\,d\mu(x) =0μ(E(v,R1/ηρ))𝑑R\displaystyle=\int_{0}^{\infty}\mu(E(v,R^{-1/\eta\rho}))\,dR
01μ(E(v,R1/ηρ))𝑑R+1μ(E(v,R1/ηρ))𝑑R\displaystyle\leq\int_{0}^{1}\mu(E(v,R^{-1/\eta\rho}))\,dR+\int_{1}^{\infty}\mu(E(v,R^{-1/\eta\rho}))\,dR
1+βm+1l(K)inIsi1R1/ρdR=:C<,\displaystyle\leq 1+\beta^{m+1-l}(K^{\prime})^{\sum_{i\not inI}s_{i}}\int_{1}^{\infty}R^{-1/\rho}\,dR=:C^{\prime}<\infty,

where in last inequality we used fact that μ\mu is a probability measure and (36). Thus the proposition holds by taking C=max{(3c)etaρ,C}C=\max\{(3c)^{-eta\rho},C^{\prime}\}. ∎

5. Height Functions

We will need some notation before proceeding. For Λ𝒳\Lambda\in\mathcal{X}, let P(Λ)P(\Lambda) denote the set of all primitive subgroups of the lattice Λ\Lambda, i.e, the subgroups LL of the lattice Λ\Lambda satisfying L=Λspan(L)L=\Lambda\cap\mathrm{span}_{\mathbb{R}}(L), where span(L)\mathrm{span}_{\mathbb{R}}(L) is the smallest vector subspace containing LL. For every 0<lm+10<l\leq m+1, we define φl:X[1,)\varphi_{l}:X\rightarrow[1,\infty) as

φl(x)=max{L1:LP(x), rank(L)=l},\displaystyle\varphi_{l}(x)=\max\{\|L\|^{-1}:L\in P(x),\text{ rank}(L)=l\}, (38)

where .\|.\| is defined as in (17). For l=0l=0, we define φl1\varphi_{l}\equiv 1. It is easy to see that φm+11\varphi_{m+1}\equiv 1.

Also, let D>0D>0 and S[0,1)S\subset[0,1) be defined as in Theorem 3.2.

Lemma 5.1.

For every Λ𝒳\Lambda\in\mathcal{X} and 1lm1\leq l\leq m, we have φl(Λ)1\varphi_{l}(\Lambda)\geq 1.

Proof.

Fix 1lm1\leq l\leq m and Λ𝒳\Lambda\in\mathcal{X}. It is easy to see that there exists a primitive sublattice Λl\Lambda_{l} of rank ll of co-volume less than or equal to 11. This follows from a simple application of induction and Minkowski’s Convex Body Theorem. This means that the vector vΛl=IvΛl,IeIVlv_{\Lambda_{l}}=\sum_{I}v_{\Lambda_{l},I}e_{I}\in V_{l} (defined as in Def. 3.1) satisfies I|vΛl,I|1\sqrt{\sum_{I}|v_{\Lambda_{l},I}|}\leq 1, where the sum is taken over all index sets II of cardinality ll. Hence we have |vΛl,I|1|v_{\Lambda_{l},I}|\leq 1, which gives Λl1\|\Lambda_{l}\|\leq 1. Thus, φl(Λ)Λl11\varphi_{l}(\Lambda)\geq\|\Lambda_{l}\|^{-1}\geq 1.

Lemma 5.2.

For any compact subset QHQ\subset H, the following holds for every hQh\in Q, Λ𝒳\Lambda\in{\mathcal{X}},

Q1φl(Λ)φl(hΛ)Qφl(Λ),\displaystyle\|Q\|^{-1}{\varphi}_{l}(\Lambda)\leq{\varphi}_{l}(h\Lambda)\leq\|Q\|{\varphi}_{l}(\Lambda), (39)

for all 1lm1\leq l\leq m.

Proof.

Using Lemma 5.1, we have that there exists a ΛlP(Λ)\Lambda_{l}\in P(\Lambda) of rank ll such that Λl1\|\Lambda_{l}\|\leq 1 and φl(Λ)=Λl1\varphi_{l}(\Lambda)=\|\Lambda_{l}\|^{-1}. Let vΛlv_{\Lambda_{l}} be defined as in Def. 3.1. Then we have vΛl1\|v_{\Lambda_{l}}\|\leq 1. Hence we get

φl(Λ)\displaystyle\varphi_{l}(\Lambda) =1vΛlQhvΛlQφl(hΛ),\displaystyle=\frac{1}{\|v_{\Lambda_{l}}\|}\leq\frac{\|Q\|}{\|hv_{\Lambda_{l}}\|}\leq\|Q\|\varphi_{l}(h\Lambda),

where penultimate inequality follows by Lemma 4.1. Similarly, we have

φ(hΛ)Qφ(h1hΛ)=Qφ(Λ).\varphi(h\Lambda)\leq\|Q\|\varphi(h^{-1}h\Lambda)=\|Q\|\varphi(\Lambda).

Thus the lemma holds. ∎

The following Proposition introduces the method of ‘integral inequalities’ into the proof and is the place where weighted Mother inequality is used.

Proposition 5.3.

For every 0<ρ<10<\rho<1, there exists a constant C=C(ρ)C=C(\rho), such that for every tt\in\mathbb{R} and 1lm1\leq l\leq m., there exists ζ=ζ(t)1\zeta=\zeta(t)\geq 1, so that for all Λ𝒳\Lambda\in{\mathcal{X}}, we have

𝒦φlηρ(gtu(x)Λ)𝑑μ(x)\displaystyle\int_{\mathcal{K}}{\varphi}_{l}^{\eta\rho}(g_{t}u(x)\Lambda)\,d\mu({x}) Ctηρφlηρ(Λ)\displaystyle\leq Ct^{-\eta\rho}{\varphi}_{l}^{\eta\rho}(\Lambda) (40)
+(ζmax{φljaφl+j1a:aS,1jmin{l,m+1l}})ηρ.\displaystyle+\left(\zeta\max\{\varphi_{l-j}^{a}\varphi_{l+j}^{1-a}:a\in S,1\leq j\leq\min\{l,m+1-l\}\}\right)^{\eta\rho}. (41)
Proof.

Fix 1lm1\leq l\leq m, 0<ρ<10<\rho<1 and tt\in\mathbb{R}. Let us define

ζ={gtu(x):x𝒦},\zeta^{\prime}=\|\{g_{t}u(x):x\in\mathcal{K}\}\|,

and define ζ=max{D.ζ,1}\zeta=\max\{D.\zeta^{\prime},1\}. Also, define ΛlP(Λ)\Lambda_{l}\in P(\Lambda) be sublattice of rank ll such that φl(Λ)=Λl1\varphi_{l}(\Lambda)=\|\Lambda_{l}\|^{-1}. Let vΛlv_{\Lambda_{l}} be defined as in Def. 3.1. Claim that for all x𝒦x\in\mathcal{K}, we have

φl(gtu(x)Λ)max{1gtu(x)Λl,ζmax{φljaφl+j1a:aS,1jmin{l,m+1l}}}.\displaystyle\varphi_{l}(g_{t}u(x)\Lambda)\leq\max\left\{\frac{1}{\|g_{t}u(x)\Lambda_{l}\|},\zeta\max\{\varphi_{l-j}^{a}\varphi_{l+j}^{1-a}:a\in S,1\leq j\leq\min\{l,m+1-l\}\}\right\}. (42)

To see this, note that if φl(gtu(x)Λ)gtu(x)Λl1\varphi_{l}(g_{t}u(x)\Lambda)\neq\|g_{t}u(x)\Lambda_{l}\|^{-1}, then the following holds. Using Lemma 5.1, we get that in this case, there exists Λl,xP(Λ)\Lambda_{l,x}\in P(\Lambda) with Λl,xΛl\Lambda_{l,x}\neq\Lambda_{l} and gtu(x)Λl,x1\|g_{t}u(x)\Lambda_{l,x}\|\leq 1 such that

φl(gtu(x)Λ)\displaystyle\varphi_{l}(g_{t}u(x)\Lambda) =gtu(x)Λl,x1\displaystyle=\|g_{t}u(x)\Lambda_{l,x}\|^{-1}
ζ(gtu(x))1gtu(x)Λl,xusing Lemma 4.1\displaystyle\leq\frac{\zeta^{\prime}}{\|(g_{t}u(x))^{-1}g_{t}u(x)\Lambda_{l,x}\|}\quad\text{using Lemma \ref{lem: operator norm make sense}}
=ζΛl,x\displaystyle=\frac{\zeta^{\prime}}{\|\Lambda_{l,x}\|}
ζmin{1Λl,xaΛl1a:0<a<1}\displaystyle\leq\zeta^{\prime}\min\{\frac{1}{\|\Lambda_{l,x}\|^{a}\|\Lambda_{l}\|^{1-a}}:0<a<1\}
Dζmax{1ΛlΛl,xaΛl+Λl,x1a:aS}using Theorem 3.2\displaystyle\leq D\zeta^{\prime}\max\{\frac{1}{\|\Lambda_{l}\cap\Lambda_{l,x}\|^{a}\|\Lambda_{l}+\Lambda_{l,x}\|^{1-a}}:a\in S\}\quad\text{using Theorem \ref{thm: Mother Ineq}}
ζmax{φljaφl+j1a:aS,1jmin{l,m+1l}}\displaystyle\leq\zeta\max\{\varphi_{l-j}^{a}\varphi_{l+j}^{1-a}:a\in S,1\leq j\leq\min\{l,m+1-l\}\}

From (42), we have

𝒦φlρη(gtu(x)Λ)𝑑μ(x)\displaystyle\int_{\mathcal{K}}{\varphi}_{l}^{\rho\eta}(g_{t}u(x)\Lambda)\,d\mu({x})
𝒦1gtu(x)vΛlηρ𝑑μ(x)+(ζmax{φljaφl+j1a:aS,1jmin{l,m+1l}})ηρ\displaystyle\leq\int_{\mathcal{K}}\frac{1}{\|g_{t}u(x)v_{\Lambda_{l}}\|^{\eta\rho}}\,d\mu({x})+\left(\zeta\max\{\varphi_{l-j}^{a}\varphi_{l+j}^{1-a}:a\in S,1\leq j\leq\min\{l,m+1-l\}\}\right)^{\eta\rho}
C(ρ,l)tρη1vΛlηρ+(ζmax{φljaφl+j1a:aS,1jmin{l,m+1l}})ηρ,\displaystyle\leq C(\rho,l)t^{-\rho\eta}\frac{1}{\|v_{\Lambda_{l}}\|^{\eta\rho}}+\left(\zeta\max\{\varphi_{l-j}^{a}\varphi_{l+j}^{1-a}:a\in S,1\leq j\leq\min\{l,m+1-l\}\}\right)^{\eta\rho},

where the last inequality follows from Proposition 4.4. Hence, the proposition follows with C(ρ)=maxlC(ρ,l)C(\rho)=\max_{l}C(\rho,l). ∎

Lemma 5.4.

There exists constants α0>α1>>αm+1\alpha_{0}>\alpha_{1}>\ldots>\alpha_{m+1} such that

αi1+max{aαij+(1a)αi+j:aS},\displaystyle\alpha_{i}\geq 1+\max\{a\alpha_{i-j}+(1-a)\alpha_{i+j}:a\in S\}, (43)

for all 1im+11\leq i\leq m+1 and 1jmin{i,m+1i}1\leq j\leq\min\{i,m+1-i\}.

Proof.

Let c=supS<1c=\sup S<1. We will define αl\alpha_{l} by induction. Let α0=1\alpha_{0}=1 and suppose that αk\alpha_{k} is already defined for 0k<l0\leq k<l satisfying α0>α1>>αl1\alpha_{0}>\alpha_{1}>\ldots>\alpha_{l-1} and (43) for i,ji,j satisfying i+j<li+j<l. Then define αl\alpha_{l} inductively as

αl=11c+min{αl1,11cmin1jl/2{αljcαl2j}}.\alpha_{l}=-\frac{1}{1-c}+\min\{\alpha_{l-1},\frac{1}{1-c}\min_{1\leq j\leq l/2}\{\alpha_{l-j}-c\alpha_{l-2j}\}\}.

Then, it is clear that αl<αl1\alpha_{l}<\alpha_{l-1}. Also for 1il1\leq i\leq l, 1min{i,m+1i}j1\leq\min\{i,m+1-i\}j\leq such that i+jli+j\leq l, we have

αimax{aαij+(1a)αi+j:aS}\displaystyle\alpha_{i}-\max\{a\alpha_{i-j}+(1-a)\alpha_{i+j}:a\in S\}
αi(cαij+(1c)αi+j)\displaystyle\geq\alpha_{i}-(c\alpha_{i-j}+(1-c)\alpha_{i+j}) using αij<αi+j\alpha_{i-j}<\alpha_{i+j} and the definition of cc
=(1c)(11c(αlj(cαl2j)αl)\displaystyle=(1-c)\left(\frac{1}{1-c}\left(\alpha_{l-j}-(c\alpha_{l-2j}\right)-\alpha_{l}\right)
1\displaystyle\geq 1 using the definition of αl.\displaystyle\text{using the definition of }\alpha_{l}.

Thus (43) holds for i,ji,j satisfying i+jli+j\leq l. Hence, by induction, the sequence exists. ∎

Fix α0,,αm+1\alpha_{0},\ldots,\alpha_{m+1} as in Lemma 5.4. For 0<ε<10<\varepsilon<1 and 0<ρ<10<\rho<1, we define the function

fε,ρ(Λ)=l=0m+1εαlφlηρ(Λ)\displaystyle f_{\varepsilon,\rho}(\Lambda)=\sum_{l=0}^{m+1}\varepsilon^{\alpha_{l}}\varphi_{l}^{\eta\rho}(\Lambda) (44)
Proposition 5.5.

For every 0<ρ<10<\rho<1, there exists a constant Cρ>0C_{\rho}>0, depending only on ρ\rho such that for every t>1t>1, there exists b=b(t)0b=b(t)\geq 0 and 0<ε=ε(t)<10<\varepsilon=\varepsilon(t)<1 so that for all Λ𝒳\Lambda\in\mathcal{X}, the following holds

𝒦fε,ρ(gtu(x)Λ)𝑑μ(x)Cρtηρfε,ρ(Λ)+b.\displaystyle\int_{\mathcal{K}}f_{\varepsilon,\rho}(g_{t}u(x)\Lambda)\,d\mu(x)\leq C_{\rho}t^{-\eta\rho}f_{\varepsilon,\rho}(\Lambda)+b. (45)
Proof.

Fix t>1t>1 and 0<ρ<10<\rho<1. Let C,ζC,\zeta be the constants provided by Proposition 5.3. Let 0<ε<10<\varepsilon<1 be a constant to be determined. Suppose Λ𝒳\Lambda\in\mathcal{X}. Then using Proposition 5.3, we get that

𝒦fε,ρ(gtu(x)Λ)𝑑μ(x)\displaystyle\int_{\mathcal{K}}f_{\varepsilon,\rho}(g_{t}u(x)\Lambda)\,d\mu(x)
=εα0+εαm+1+𝒦l=1mεαlφlηρ(gtu(x)Λ)dμ(x)\displaystyle=\varepsilon^{\alpha_{0}}+\varepsilon^{\alpha_{m+1}}+\int_{\mathcal{K}}\sum_{l=1}^{m}\varepsilon^{\alpha_{l}}\varphi_{l}^{\eta\rho}(g_{t}u(x)\Lambda)\,d\mu(x)
εα0+εαm+1+Ctηρl=1mεαlφlηρ(Λ)\displaystyle\leq\varepsilon^{\alpha_{0}}+\varepsilon^{\alpha_{m+1}}+Ct^{-\eta\rho}\sum_{l=1}^{m}\varepsilon^{\alpha_{l}}\varphi_{l}^{\eta\rho}(\Lambda)
+l=1mεαlζηρmax{(φljηρ)a(φl+jηρ)1a:aS,1jmin{l,m+1l}}\displaystyle+\sum_{l=1}^{m}\varepsilon^{\alpha_{l}}\zeta^{\eta\rho}\max\{(\varphi_{l-j}^{\eta\rho})^{a}(\varphi_{l+j}^{\eta\rho})^{1-a}:a\in S,1\leq j\leq\min\{l,m+1-l\}\} (46)

Note that

φijηρ(x)\displaystyle\varphi_{i-j}^{\eta\rho}(x) εαijηρfε,ρ(x),\displaystyle\leq\varepsilon^{-\alpha_{i-j}\eta\rho}f_{\varepsilon,\rho}(x),
φi+jηρ(x)\displaystyle\varphi_{i+j}^{\eta\rho}(x) εαi+jηρfε,ρ.\displaystyle\leq\varepsilon^{-\alpha_{i+j}\eta\rho}f_{\varepsilon,\rho}.

Thus, we have

εαimax{(φijηρ(x))a(φi+jηρ(x))1a:aS,1jmin{l,m+1l}}\displaystyle\varepsilon^{\alpha_{i}}\max\{(\varphi_{i-j}^{\eta\rho}(x))^{a}(\varphi_{i+j}^{\eta\rho}(x))^{1-a}:a\in S,1\leq j\leq\min\{l,m+1-l\}\}
εαimax{ε(aαij+(1a)αi+j)ηρfε,ρ(x):aS,1jmin{l,m+1l}}\displaystyle\leq\varepsilon^{\alpha_{i}}\max\{\varepsilon^{-(a\alpha_{i-j}+(1-a)\alpha_{i+j})\eta\rho}f_{\varepsilon,\rho}(x):a\in S,1\leq j\leq\min\{l,m+1-l\}\}
=fε,ρ(x)max{εαi(aαij+(1a)αi+j)ηρ:aS,1jmin{l,m+1l}}\displaystyle=f_{\varepsilon,\rho}(x)\max\{\varepsilon^{\alpha_{i}-(a\alpha_{i-j}+(1-a)\alpha_{i+j})\eta\rho}:a\in S,1\leq j\leq\min\{l,m+1-l\}\}
εfε,ρ(x)using (43) and fact that ηρ1.\displaystyle\leq\varepsilon f_{\varepsilon,\rho}(x)\quad\text{using \eqref{eq: lem alpha i condition} and fact that $\eta\rho\leq 1$.} (47)

Thus, we get from (46) and (47) that

𝒦fε,ρ(gtu(x)Λ)𝑑μ(x)Ctηρfε,ρ(Λ)+(εα0+εαm)(1Ctηρ)+mεζηρfε,ρ(Λ)\displaystyle\int_{\mathcal{K}}f_{\varepsilon,\rho}(g_{t}u(x)\Lambda)\,d\mu(x)\leq Ct^{-\eta\rho}f_{\varepsilon,\rho}(\Lambda)+(\varepsilon^{\alpha_{0}}+\varepsilon^{\alpha_{m}})(1-Ct^{-\eta\rho})+m\varepsilon\zeta^{\eta\rho}f_{\varepsilon,\rho}(\Lambda)

Choosing ε=Ctηρmζ\varepsilon=\frac{Ct^{-\eta\rho}}{m\zeta} and b=(εα0+εαm)(1Ctηρ)b=(\varepsilon^{\alpha_{0}}+\varepsilon^{\alpha_{m}})(1-Ct^{-\eta\rho}), and Cρ=2CC_{\rho}=2C, we get (45). Thus the proposition follows. ∎

6. The Contraction Hypothesis

Definition 6.1 (The Contraction Hypothesis).

Suppose YY is a metric space equipped with an action of GG. Given a collection of functions ft:Y(0,]:tS{f_{t}:Y\rightarrow(0,\infty]:t\in S} for some unbounded set S(0,)S\subset(0,\infty) and real numbers 0<β0<\beta, we say that μ\mu satisfies the ((ft)t,β)((f_{t})_{t},\beta)-contraction hypothesis on XX if the following properties hold:

  1. (1)

    The set Z={ft=}Z=\{f_{t}=\infty\} is independent of tt and is GG-invariant.

  2. (2)

    For every tSt\in S, ftf_{t} is uniformly log Lipschitz with respect to the HH (defined in (8)) action. That is for every bounded neighborhood 𝒪\mathcal{O} of identity in HH, there exists a constant C𝒪1C_{\mathcal{O}}\geq 1 such that for every g𝒪g\in\mathcal{O}, yYy\in Y and tt\in\mathbb{N},

    C𝒪1ft(y)ft(gy)C𝒪ft(y).\displaystyle C_{\mathcal{O}}^{-1}f_{t}(y)\leq f_{t}(gy)\leq C_{\mathcal{O}}f_{t}(y). (48)
  3. (3)

    There exists a constant c1c\geq 1 such that the following holds: for every tSt\in S, there exists T>0T>0 such that for all yYy\in Y with ft(y)>Tf_{t}(y)>T,

    𝒦ft(gtu(x)y)𝑑μ(x)cft(y)tβ.\displaystyle\int_{\mathcal{K}}f_{t}(g_{t}u(x)y)\,d\mu(x)\leq cf_{t}(y)t^{-\beta}. (49)

The functions ftf_{t} will be referred to as height functions.

Remark 6.2.

The above definition of Contraction Hypothesis is motivated from the corresponding definition in [Khalilsing].

Theorem 6.3.

Let YY be a metric space equipped with an action by GG. Assume that μ\mu satisfies the ({ft}tS,β)(\{f_{t}\}_{t\in S},\beta)-contraction hypothesis on YY for a real numbers 0<βs0<\beta\leq s. Then for all yX\{ft=}y\in X\backslash\{f_{t}=\infty\},

dimH(x𝒦:fs(atu(x)y)t for all sS)sβ1+w1.\displaystyle\dim_{H}\left(x\in\mathcal{K}:f_{s}(a_{t}u(x)y)\rightarrow_{t\rightarrow\infty}\infty\text{ for all }s\in S\right)\leq s-\frac{\beta}{1+w_{1}}.
Proof.

Without loss of generality, we may assume that 0𝒦0\in\mathcal{K}. Now fix α>0\alpha>0 large enough so that 𝒦[α,α]m\mathcal{K}\subset[-\alpha,\alpha]^{m}. Let us define OHO\subset H as

O={1(t1tm)1/(m+1)(t1tm1)u(x):x[2α,2α]m,citici1 for all 1im}.\displaystyle O=\left\{\frac{1}{(t_{1}\ldots t_{m})^{1/(m+1)}}\begin{pmatrix}t_{1}\\ &\ddots\\ &&t_{m}\\ &&&1\end{pmatrix}u(x):x\in[-2\alpha,2\alpha]^{m},c_{i}\leq t_{i}\leq c_{i}^{-1}\text{ for all }1\leq i\leq m\right\}. (50)

Let A>1A>1 be the constant so that (48) holds for all tSt\in S, yYy\in Y and hOh\in O with CO=AC_{O}=A.

Fix yYy\in Y. Fix tSt\in S large enough so that t>ci1t>c_{i}^{-1} for all ii. For the sake of simplicity, let g=gtg=g_{t}. For every kk\in\mathbb{N} and 1jm1\leq j\leq m, we define Nk(j)N_{k}(j) as the unique integer such that cjNk(j)tk(1+wj)<cjNk(j)1c_{j}^{N_{k}(j)}\leq t^{-k(1+w_{j})}<c_{j}^{N_{k}(j)-1}. We define for all kk\in\mathbb{N}, the element hkGh_{k}\in G as

hk=1(c1Nk(1)cmNk(m))1/(m+1)(c1Nk(1)cmNk(m)1).\displaystyle h_{k}=\frac{1}{(c_{1}^{-N_{k}(1)}\ldots c_{m}^{-N_{k}(m)})^{1/(m+1)}}\begin{pmatrix}c_{1}^{-N_{k}(1)}\\ &\ddots\\ &&c_{m}^{-N_{k}(m)}\\ &&&1\end{pmatrix}. (51)

Let us quickly recall some notation. For 1im1\leq i\leq m, 𝒦i\mathcal{K}_{i} is the limit set of the IFS Φi={ϕi,e:eEi}\Phi_{i}=\{\phi_{i,e}:e\in E_{i}\}, with common contraction ratio cic_{i} and pi=#Eip_{i}=\#E_{i}. The dimension of 𝒦i\mathcal{K}_{i} equals si=logpi/logcis_{i}=-\log p_{i}/\log c_{i}. The set 𝒦=i𝒦i\mathcal{K}=\prod_{i}\mathcal{K}_{i} has dimension s=isis=\sum_{i}s_{i}. The measure μi\mu_{i} denote the normalised restriction of HsiH^{s_{i}} to 𝒦i\mathcal{K}_{i} and the measure μ\mu on 𝒦\mathcal{K} is defined as μ=iμi\mu=\otimes_{i}\mu_{i}. For 1jm1\leq j\leq m and kk\in\mathbb{N}, we define j(l)\mathcal{F}_{j}(l) as in (10) corresponding to Φj={ϕj,e:eEj}\Phi_{j}=\{\phi_{j,e}:e\in E_{j}\}. We also define ηj:Ej\eta_{j}:E_{j}^{\mathbb{N}}\rightarrow\mathbb{R} as in (12). We define κj:lj(l)\kappa_{j}:\cup_{l}\mathcal{F}_{j}(l)\rightarrow\mathbb{R} as κj(ϕj,e1ϕj,el(𝒦j))=ϕj,e1ϕj,el(0)\kappa_{j}(\phi_{j,e_{1}}\circ\ldots\circ\phi_{j,e_{l}}(\mathcal{K}_{j}))=\phi_{j,e_{1}}\circ\ldots\circ\phi_{j,e_{l}}(0). Let us define (k)=jj(Nk(j))\mathcal{F}(k)=\prod_{j}\mathcal{F}_{j}(N_{k}(j)). Also define κ:k(k)m\kappa:\cup_{k}\mathcal{F}(k)\rightarrow\mathbb{R}^{m} as restriction of κ1×κm\kappa_{1}\times\ldots\kappa_{m}. Note that since 0𝒦0\in\mathcal{K}, so κ(R)\kappa(R) belongs to RR for all Rk(k)R\in\cup_{k}\mathcal{F}(k).

Let T>0T^{\prime}>0 and c>0c>0 be so that (49) holds for ftf_{t} and yYy\in Y with ft(y)>Tf_{t}(y)>T^{\prime}. Let T=A2TT=A^{2}T.

Then we have that

{x𝒦:fτ(asu(x)y)s for all τS}NZ(N),\displaystyle\left\{x\in\mathcal{K}:f_{\tau}(a_{s}u(x)y)\rightarrow_{s\rightarrow\infty}\infty\text{ for all }\tau\in S\right\}\subset\bigcup_{N\in\mathbb{N}}Z(N), (52)

where Z(N)={x𝒦:ft(gnu(x)y)>T for all nN}Z(N)=\{x\in\mathcal{K}:f_{t}(g^{n}u(x)y)>T\text{ for all }n\geq N\}. Fix NN\in\mathbb{N}.

Let us make some easy observations before proceeding:

Observation 1 For all kk\in\mathbb{N}, we have hkgkh_{k}g^{-k} and gkhk1g_{k}h_{k}^{-1} belongs to OO (defined as in (50)).
Explanation: This follows from the definition of hkh_{k} along with fact that

gk=1(tk(1+w1)tk(1+wm))1/(m+1)(tk(1+w1)tk(1+wm)1).g^{k}=\frac{1}{(t^{k(1+w_{1})}\ldots t^{k(1+w_{m})})^{1/(m+1)}}\begin{pmatrix}t^{k(1+w_{1})}\\ &\ddots\\ &&t^{k(1+w_{m})}\\ &&&1\end{pmatrix}.

Observation 2 For M>NM>N, if R(M)R\in\mathcal{F}(M) is such that RZ(N)R\cap Z(N)\neq\emptyset, then ft(gMu(x)y)>T/Af_{t}(g^{M}u(x)y)>T/A for all xRx\in R.
Explanation: Suppose xRZ(N)x\in R\cap Z(N). Then as xZ(N)x\in Z(N) and M>NM>N, we have ft(gMu(x)y)>Tf_{t}(g^{M}u(x)y)>T. Now if xRx^{\prime}\in R, then the jj-th co-ordinate xxx^{\prime}-x is less than 2cjNM(j)α2tM(1+wj)α2c_{j}^{N_{M}(j)}\alpha\leq 2t^{-M(1+w_{j})}\alpha. If define x′′mx^{\prime\prime}\in\mathbb{R}^{m} as the vector whose jj-th entry is tM(1+wj)t^{-M(1+w_{j})} times the jj-th entry of xxx^{\prime}-x, then u(x′′)Ou(x^{\prime\prime})\in O. Thus we have

ft(gMu(x)y)=ft(gMu(xx)gMgMu(x)y)=ft(u(x′′)gMu(x)y)ft(gMu(x)y)A=TA.f_{t}(g^{M}u(x^{\prime})y)=f_{t}(g^{M}u(x^{\prime}-x)g^{-M}g^{M}u(x)y)=f_{t}(u(x^{\prime\prime})g^{M}u(x)y)\geq\frac{f_{t}(g^{M}u(x)y)}{A}=\frac{T}{A}.

Observation 3 For any measurable function ψ:m+\psi:\mathbb{R}^{m}\rightarrow\mathbb{R}_{+}, measurable set XmX\subset\mathbb{R}^{m} and MM\in\mathbb{N}, we have

R(M)RXRψ(x)𝑑μ(x)R(M1)RXRψ(x)𝑑μ(x).\sum_{\begin{subarray}{c}R\in\mathcal{F}(M)\\ R\cap X\neq\emptyset\end{subarray}}\int_{R}\psi(x)\,d\mu(x)\leq\sum_{\begin{subarray}{c}R\in\mathcal{F}(M-1)\\ R\cap X\neq\emptyset\end{subarray}}\int_{R}\psi(x)\,d\mu(x).

Explanation: Using the fact that each (Ψj)(\Psi_{j}) satisfies open set condition, we get that for any R(M1)R\in\mathcal{F}(M-1), the following holds

Rψ(x)𝑑μ(x)=R(M)RRRψ(x)𝑑μ(x).\int_{R}\psi(x)\,d\mu(x)=\sum_{\begin{subarray}{c}R^{\prime}\in\mathcal{F}(M)\\ R^{\prime}\subset R\end{subarray}}\int_{R^{\prime}}\psi(x)\,d\mu(x).

The observation now follows immediately.

Observation 4 For any measurable function ψ:Y+\psi:Y\rightarrow\mathbb{R}_{+}, MM\in\mathbb{N} and R(M)R\in\mathcal{F}(M), yYy\in Y, we have

Rψ(hMu(x)y)𝑑μ(x)=μ(R)𝒦ψ(u(x)hMu(κ(R))y)𝑑μ(x).\int_{R}\psi(h_{M}u(x)y)\,d\mu(x)=\mu(R)\int_{\mathcal{K}}\psi(u(x)h_{M}u(\kappa(R))y)\,d\mu(x).

Explanation: Let us define the matrix hMh_{M}^{\prime} as

hM=(c1NM(1)cmNM(m)).\displaystyle h_{M}^{\prime}=\begin{pmatrix}c_{1}^{N_{M}(1)}\\ &\ddots\\ &&c_{m}^{N_{M}(m)}\end{pmatrix}. (53)

Using the definition of (M)\mathcal{F}(M) and κ\kappa, it is easy to see that R=hM𝒦+κ(R)R=h_{M}^{\prime}\mathcal{K}+\kappa(R). Since μ=iμi\mu=\otimes_{i}\mu_{i} and μi\mu_{i} is a Bernoulli measure using the Proposition 2.1 corresponding to uniform measure on EiE_{i}, we get that 1μ(R)μ|R\frac{1}{\mu(R)}\mu|_{R} equals pushforward of μ\mu under the map xhMx+κ(R)x\mapsto h_{M}^{\prime}x+\kappa(R). Thus, we have

Rψ(hMu(x)y)𝑑μ(x)\displaystyle\int_{R}\psi(h_{M}u(x)y)\,d\mu(x) =μ(R)𝒦ψ(hMu(hMx+κ(R))dμ(x)\displaystyle=\mu(R)\int_{\mathcal{K}}\psi(h_{M}u(h_{M}^{\prime}x+\kappa(R))\,d\mu(x)
=μ(R)𝒦ψ(hMu(hMx)hM1hMu(κ(R))dμ(x)\displaystyle=\mu(R)\int_{\mathcal{K}}\psi(h_{M}u(h_{M}^{\prime}x)h_{M}^{-1}h_{M}u(\kappa(R))\,d\mu(x)
=μ(R)𝒦ψ(u((hM)1hMx)hMu(κ(R))dμ(x)\displaystyle=\mu(R)\int_{\mathcal{K}}\psi(u((h_{M}^{\prime})^{-1}h_{M}^{\prime}x)h_{M}u(\kappa(R))\,d\mu(x)
=μ(R)𝒦ψ(u(x)hMu(κ(R))y)𝑑μ(x).\displaystyle=\mu(R)\int_{\mathcal{K}}\psi(u(x)h_{M}u(\kappa(R))y)\,d\mu(x).

Observation 5 For any measurable function ψ:Y+\psi:Y\rightarrow\mathbb{R}_{+}, MM\in\mathbb{N} and R(M)R\in\mathcal{F}(M), yYy\in Y, we have

μ(R)ψ(hMu(κ(R)y)ARψ(hMu(x)y)dμ(x).\mu(R)\psi(h_{M}u(\kappa(R)y)\leq A\int_{R}\psi(h_{M}u(x)y)\,d\mu(x).

Explanation: Note that for all x𝒦x\in\mathcal{K}, we have

μ(R)ψ(hMu(κ(R))y)\displaystyle\mu(R)\psi(h_{M}u(\kappa(R))y) μ(R)A𝒦ψ((gMhM1)u(x)hMu(κ(R))y)𝑑μ(x)using Observation 1\displaystyle\leq\mu(R)A\int_{\mathcal{K}}\psi((g^{M}h_{M}^{-1})u(x)h_{M}u(\kappa(R))y)\,d\mu(x)\quad\text{using Observation 1}
=ARψ((gMhM1)hMu(x)y)𝑑μ(x) using Observation 4 for yψ((gMhM1)y)\displaystyle=A\int_{R}\psi((g^{M}h_{M}^{-1})h_{M}u(x)y)\,d\mu(x)\quad\text{ using Observation 4 for $y\mapsto\psi((g^{M}h_{M}^{-1})y)$}
=ARψ(gMu(x)y)𝑑μ(x)\displaystyle=A\int_{R}\psi(g^{M}u(x)y)\,d\mu(x)

Observation 6 For M>NM>N and R(M1)R\in\mathcal{F}(M-1) such that RZ(N)R\cap Z(N)\neq\emptyset, we have

𝒦ft(gu(x)hM1u(κ(R))y)𝑑μ(x)tβft(hM1u(κ(R))y).\int_{\mathcal{K}}f_{t}(gu(x)h_{M-1}u(\kappa(R))y)\,d\mu(x)\leq t^{-\beta}f_{t}(h_{M-1}u(\kappa(R))y).

Explanation: Since RZ(N)R\cap Z(N)\neq\emptyset, we get from Observation 2 and fact that κ(R)R\kappa(R)\in R that ft(gM1u(κ(R))y)>T/Af_{t}(g^{M-1}u(\kappa(R))y)>T/A. Combining this with Observation 1, we get that ft(gM1u(κ(R))y)>T/A2=Tf_{t}(g^{M-1}u(\kappa(R))y)>T/A^{2}=T^{\prime}. Now above observation follows by using (49).

Note that for any M>NM>N, we have using Observation 2 that

R(M)RZ(N)μ(R)\displaystyle\sum_{\begin{subarray}{c}R\in\mathcal{F}(M)\\ R\cap Z(N)\neq\emptyset\end{subarray}}\mu(R) ATR(M)RZ(N)Rft(gMu(x)y)𝑑μ(x).\displaystyle\leq\frac{A}{T}\sum_{\begin{subarray}{c}R\in\mathcal{F}(M)\\ R\cap Z(N)\neq\emptyset\end{subarray}}\int_{R}f_{t}(g^{M}u(x)y)\,d\mu(x). (54)

Also, we have for L>NL>N

R(L)RZ(N)Rft(gLu(x)y)𝑑μ(x)\displaystyle\sum_{\begin{subarray}{c}R\in\mathcal{F}(L)\\ R\cap Z(N)\neq\emptyset\end{subarray}}\int_{R}f_{t}(g^{L}u(x)y)\,d\mu(x) R(L1)RZ(N)Rft(gLu(x)y)𝑑μ(x) using Observation 3\displaystyle\leq\sum_{\begin{subarray}{c}R\in\mathcal{F}(L-1)\\ R\cap Z(N)\neq\emptyset\end{subarray}}\int_{R}f_{t}(g^{L}u(x)y)\,d\mu(x)\text{ using Observation 3}
AR(L1)RZ(N)Rft(ghL1u(x)y)𝑑μ(x) using Observation 1\displaystyle\leq A\sum_{\begin{subarray}{c}R\in\mathcal{F}(L-1)\\ R\cap Z(N)\neq\emptyset\end{subarray}}\int_{R}f_{t}(gh_{L-1}u(x)y)\,d\mu(x)\text{ using Observation 1}
=AR(L1)RZ(N)μ(R)𝒦ft(gu(x)hL1u(κ(R))y)𝑑μ(x) using Observation 4\displaystyle=A\sum_{\begin{subarray}{c}R\in\mathcal{F}(L-1)\\ R\cap Z(N)\neq\emptyset\end{subarray}}\mu(R)\int_{\mathcal{K}}f_{t}(gu(x)h_{L-1}u(\kappa(R))y)\,d\mu(x)\text{ using Observation 4}
AtβR(L1)RZ(N)μ(R)ft(hL1u(κ(R))y)dμ(x) using Observation 6\displaystyle\leq At^{-\beta}\sum_{\begin{subarray}{c}R\in\mathcal{F}(L-1)\\ R\cap Z(N)\neq\emptyset\end{subarray}}\mu(R)f_{t}(h_{L-1}u(\kappa(R))y)\,d\mu(x)\text{ using Observation 6}
A2tβR(L1)RZ(N)Rft(gL1u(x)y)𝑑μ(x) using Observation 5.\displaystyle\leq A^{2}t^{-\beta}\sum_{\begin{subarray}{c}R\in\mathcal{F}(L-1)\\ R\cap Z(N)\neq\emptyset\end{subarray}}\int_{R}f_{t}(g^{L-1}u(x)y)\,d\mu(x)\text{ using Observation 5}. (55)

Using (54) and iteratively using (55) for L=M,M1,,N+1L=M,M-1,\ldots,N+1, we get that for any M>NM>N the following holds

R(M)RZ(N)μ(R)\displaystyle\sum_{\begin{subarray}{c}R\in\mathcal{F}(M)\\ R\cap Z(N)\neq\emptyset\end{subarray}}\mu(R) CNA2MtMβ,\displaystyle\leq C_{N}A^{2M}t^{-M\beta}, (56)

where CN=A12NtN𝒦ft(gNu(x)y)𝑑μ(x)/TC_{N}=A^{1-2N}t^{N}\int_{\mathcal{K}}f_{t}(g^{N}u(x)y)\,d\mu(x)/T. This

For 1jm1\leq j\leq m and ll\in\mathbb{N}, we define Pl(j)P_{l}(j) as the unique integer satisfying cjPl(j)t(1+w1)l<cjPl(j)1c_{j}^{P_{l}(j)}\leq t^{-(1+w_{1})l}<c_{j}^{P_{l}(j)-1}. Let us try to cover Z(N)Z(N) by balls of size less than or equal to tM(1+w1)t^{-M(1+w_{1})}. We do this by selecting sets in jj(PM(j))\prod_{j}\mathcal{F}_{j}(P_{M}(j)) which intersect Z(N)Z(N). Note that there are jpjPM(j)=jcjsjPl(j)(c1cm)1ts(1+w1)M\prod_{j}p_{j}^{P_{M}(j)}=\prod_{j}c_{j}^{-s_{j}P_{l}(j)}\leq(c_{1}\ldots c_{m})^{-1}t^{s(1+w_{1})M} many elements in jj(PM(j))\prod_{j}\mathcal{F}_{j}(P_{M}(j)), each of which has equal μ\mu-measure. Thus, we get that

#{Rjj(PM(j)):RZ(N)}\displaystyle\#\{R\in\prod_{j}\mathcal{F}_{j}(P_{M}(j)):R\cap Z(N)\neq\emptyset\} =(#jj(PM(j))).(Rjj(PM(j))RZ(N)μ(R))\displaystyle=(\#\prod_{j}\mathcal{F}_{j}(P_{M}(j))).(\sum_{\begin{subarray}{c}R\in\prod_{j}\mathcal{F}_{j}(P_{M}(j))\\ R\cap Z(N)\neq\emptyset\end{subarray}}\mu(R))
(c1cm)1ts(1+w1)M.(R(M)RZ(N)μ(R))\displaystyle\leq(c_{1}\ldots c_{m})^{-1}t^{s(1+w_{1})M}.(\sum_{\begin{subarray}{c}R\in\mathcal{F}(M)\\ R\cap Z(N)\neq\emptyset\end{subarray}}\mu(R))
CN(A2tβts(1+w1))M,\displaystyle\leq C_{N}^{\prime}\left(A^{2}t^{-\beta}t^{s(1+w_{1})}\right)^{M},

where CN=CN(c1cm)1C_{N}^{\prime}=C_{N}(c_{1}\ldots c_{m})^{-1}. Thus for every t>0t>0, we can cover Z(N)Z(N) by CN(A2tβts(1+w1))MC_{N}^{\prime}\left(A^{2}t^{-\beta}t^{s(1+w_{1})}\right)^{M}-many hypercuboid of diameter less than or equal to max{cjPM(j)diam(𝒦j)}2α.tM(1+w1)\max\{c_{j}^{P_{M}(j)}\text{diam}(\mathcal{K}_{j})\}\leq 2\alpha.t^{-M(1+w_{1})}. Thus,

Hsγ(Z(N))limMCN(A2tβts(1+w1))M(2α)sγtM(1+w1)(sγ),\displaystyle H^{s-\gamma}(Z(N))\leq\lim_{M\rightarrow\infty}C_{N}^{\prime}\left(A^{2}t^{-\beta}t^{-s(1+w_{1})}\right)^{M}(2\alpha)^{s-\gamma}t^{-M(1+w_{1})(s-\gamma)},

which is finite for all γ>0\gamma>0 satisfying A2tβts(1+w1)t(1+w1)(sγ)1A^{2}t^{-\beta}t^{s(1+w_{1})}t^{-(1+w_{1})(s-\gamma)}\leq 1, i.e,

γ11+w1(β2logAlogt).\gamma\leq\frac{1}{1+w_{1}}\left(\beta-\frac{2\log A}{\log t}\right).

This gives that

dimH(Z(N))s11+w1(β2logAlogt).\displaystyle\dim_{H}(Z(N))\leq s-\frac{1}{1+w_{1}}\left(\beta-\frac{2\log A}{\log t}\right). (57)

Using fact that dimH(nIn)=supndimH(In)\dim_{H}(\cup_{n}I_{n})=\sup_{n}\dim_{H}(I_{n}) for any countable collection of borel sets InI_{n} and (52), (57), we get that

dimH({x𝒦:fτ(asu(x)y)s for all τS})s11+w1(β2logAlogt).\dim_{H}(\left\{x\in\mathcal{K}:f_{\tau}(a_{s}u(x)y)\rightarrow_{s\rightarrow\infty}\infty\text{ for all }\tau\in S\right\})\leq s-\frac{1}{1+w_{1}}\left(\beta-\frac{2\log A}{\log t}\right).

Since AA is independent of tt, letting tt\rightarrow\infty (which exists as SS is unbounded), we get that the theorem holds. Hence proved. ∎

7. Final Proof

Proof of Theorem 1.4.

Let 0<ρ<10<\rho<1 be given. Using Proposition 5.5, for every t>1t>1, choose ε(t)\varepsilon(t) and define the collection of height functions

{ft:=fε(t),ρ:t}.\{f_{t}:=f_{\varepsilon(t),\rho}:t\in\mathbb{R}\}.

Now it is easy to see that the action of GG on 𝒳{\mathcal{X}} satisfies ({ft},η)(\{f_{t}\},\eta)-contraction hypothesis with respect to measure μ\mu. Indeed the first properties of Definition 6.1 follow immediately from the definition of fkf_{k}. The second condition follows from Lemma 5.2 and the fact that for all tt, ftf_{t} is a linear combination of φl\varphi_{l}. The third property follows from Proposition 5.5 and for c=2Cρc=2C_{\rho} and T=btηρ/CρT=bt^{\eta\rho}/C_{\rho} corresponding to each tt (Note that the value of bb also depends on tt).

Thus by Theorem 6.3, we get that for all Λ𝒳\Lambda\in\mathcal{X} the following holds

dimH(x𝒦:fs(atu(x)y)t for all s>1)sηρ1+w1.\displaystyle\dim_{H}\left(x\in\mathcal{K}:f_{s}(a_{t}u(x)y)\rightarrow_{t\rightarrow\infty}\infty\text{ for all }s>1\right)\leq s-\frac{\eta\rho}{1+w_{1}}.

Since {x𝒦:fs(atu(x)Λ)t for all s>1}=Div(Λ,w)\{x\in\mathcal{K}:f_{s}(a_{t}u(x)\Lambda)\rightarrow_{t\rightarrow\infty}\infty\text{ for all }s>1\}=Div(\Lambda,w), we get that

dimH(Div(Λ,w))sηρ1+w1.\dim_{H}(Div(\Lambda,w))\leq s-\frac{\eta\rho}{1+w_{1}}.

Since 0<ρ<10<\rho<1 is arbitrary, the theorem follows. ∎

8. Concluding Remarks

More generally, one could ask for the dimension of weighted singular matrices. Our approach is applicable in this setting. It is also possible to deal with the product of more general fractals rather than the same contraction ratios. The main changes needed in this case involve the contraction hypothesis. These generalizations are ongoing work in progress.

References