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

Semisimple random walks on the torus

Weikun He Institute of Mathematics, Academy of Mathematics and System Science, CAS, Zhongguancun East Road 55, Beijing 100190, P.R. China heweikun@amss.ac.cn  and  Nicolas de Saxcé CNRS – Université Sorbonne Paris Nord, LAGA, 93430 Villetaneuse, France. desaxce@math.univ-paris13.fr
Abstract.

We study linear random walks on the torus and show a quantitative equidistribution statement, under the assumption that the Zariski closure of the acting group is semisimple.

Key words and phrases:
Equidistribution, sum-product, Lyapunov exponent, Fourier decay
2010 Mathematics Subject Classification:
Primary 37A17, 11B75; Secondary 37A45, 11L07, 20G30.

1. Introduction

Let d2d\geq 2 be an integer and 𝕋d=d/d\mathbb{T}^{d}=\mathbb{R}^{d}/\mathbb{Z}^{d} the torus of dimension dd. We study a random walk (xn)n0(x_{n})_{n\geq 0} on 𝕋d\mathbb{T}^{d} given by

n0,xn=gng1x0\forall n\geq 0,\quad x_{n}=g_{n}\dots g_{1}x_{0}

where (gn)n1(g_{n})_{n\geq 1} is a sequence of independent identically distributed random variables with law μ\mu on GLd()\operatorname{GL}_{d}(\mathbb{Z}). Let Γ\Gamma denote the group generated by the support of μ\mu, and GG be the Zariski closure of Γ\Gamma in GLd()\operatorname{GL}_{d}(\mathbb{R}). In [11], Bourgain, Furman, Lindenstrauss and Mozes showed that if GG acts strongly irreducibly and proximally on d\mathbb{R}^{d}, then the random walk (xn)n0(x_{n})_{n\geq 0} equidistributes in law to the Haar probability measure m𝕋d\operatorname{m}_{\mathbb{T}^{d}} as soon as x0x_{0} is irrational, i.e.

x0d/d,μnδx0n+m𝕋d.\forall x_{0}\not\in\mathbb{Q}^{d}/\mathbb{Z}^{d},\quad\mu^{*n}*\delta_{x_{0}}\rightharpoonup^{*}_{n\to+\infty}\operatorname{m}_{\mathbb{T}^{d}}.

Moreover, this result is quantitative : an explicit rate of convergence is obtained in terms of the distance from x0x_{0} to rational points of small denominator. Following their strategy, we showed in [25] that their theorem is still valid without the proximality assumption, as long as the action of GG on d\mathbb{R}^{d} is strongly irreducible. On the other hand, the theory developed by Benoist and Quint in their series of articles [3, 5, 6, 4] made it clear that when studying random walks on homogeneous spaces, it is most natural to only assume that the acting algebraic group GG is semisimple. Indeed, under this assumption [5, Theorem 1.1] gives a full classification of stationary measures, which in turn implies the very general equidistribution results of [6]. It is therefore desirable to obtain quantitative convergence results similar to those of [11] or [25] in this more general setting, and this is the goal of the present article.


Of course, without the irreducibility assumption, there can exist some proper closed Γ\Gamma-invariant subsets of 𝕋d\mathbb{T}^{d}, and the random walk may not equidistribute to the Haar measure, even if the starting point x0x_{0} is irrational. So in order to state our main result, we need to set up some notation. Let GG^{\circ} denote the identity component of GG for the Zariski topology. The subalgebra of d()\operatorname{\mathcal{M}}_{d}(\mathbb{R}) generated by GG^{\circ} is denoted by EE. Since GG is semisimple, we may write d=V0V1Vr\mathbb{R}^{d}=V_{0}\oplus V_{1}\oplus\dots\oplus V_{r} where for i=0,,ri=0,\dotsc,r, ViV_{i} is a maximal sum of simple GG-modules having the same top Lyapunov exponent for the action of μ\mu. Reordering the subspaces ViV_{i}, we may assume in addition that

λ1(μ,V1)>>λ1(μ,Vr)>λ1(μ,V0)=0.\lambda_{1}(\mu,V_{1})>\dots>\lambda_{1}(\mu,V_{r})>\lambda_{1}(\mu,V_{0})=0.

The space V0V_{0} will play a special role in our analysis of the random walk behavior. By a result of Furstenberg, the image of GG in GL(V0)\operatorname{GL}(V_{0}) is compact, and for that reason, we shall say that V0V_{0} is the sum of all compact factors of GG in d\mathbb{R}^{d}. We define a quasi-norm on d\mathbb{R}^{d} by

|v|=max0irvi1λ1(μ,Vi)\lvert v\rvert=\max_{0\leq i\leq r}\lVert v_{i}\rVert^{\frac{1}{\lambda_{1}(\mu,V_{i})}}

where v=v0++vrv=v_{0}+\dotsc+v_{r} is the decomposition of vv according to the direct sum d=i=1rVi\mathbb{R}^{d}=\oplus_{i=1}^{r}V_{i}. By convention, we set 10=+\frac{1}{0}=+\infty and

v0+={0if v01,+otherwise.\lVert v_{0}\rVert^{+\infty}=\begin{cases}0&\text{if }\lVert v_{0}\rVert\leq 1,\\ +\infty&\text{otherwise.}\end{cases}

This quasi-norm induces a quasi-distance on d\mathbb{R}^{d} given by d~(x,y)=|xy|\tilde{d}(x,y)=\lvert x-y\rvert, which projects to a quasi-distance on 𝕋d\mathbb{T}^{d}, still denoted by d~\tilde{d}. A finite measure μ\mu on GLd()\operatorname{GL}_{d}(\mathbb{Z}) is said to have a finite exponential moment if there exists τ>0\tau>0 such that

gτdμ(g)<+,\int\lVert g\rVert^{\tau}\,\mathrm{d}\mu(g)<+\infty,

where \lVert\cdot\rVert denotes a norm on the space Md()M_{d}(\mathbb{R}) of d×dd\times d matrices; this definition does not depend on the choice of the norm. Our goal is the following theorem.

Theorem 1.1 (Equidistribution of semisimple linear random walks on 𝕋d\mathbb{T}^{d}).

Let μ\mu be a probability measure on GLd()\operatorname{GL}_{d}(\mathbb{Z}) having a finite exponential moment. Denote by GGLd()G\subset\operatorname{GL}_{d}(\mathbb{R}) the algebraic group generated by μ\mu, by GG^{\circ} its identity component, and let EE be the subalgebra of d()\operatorname{\mathcal{M}}_{d}(\mathbb{R}) generated by GG^{\circ}. As above, we write V0V_{0} for the sum of all compact factors of GG in d\mathbb{R}^{d}. If the algebraic group GG is semisimple, then for every λ(0,1)\lambda\in(0,1), there exists C=C(μ,λ)0C=C(\mu,\lambda)\geq 0 such that the following holds.

Given x0𝕋dx_{0}\in\mathbb{T}^{d}, assume that for some t(0,12)t\in(0,\frac{1}{2}), a0d{0}a_{0}\in\mathbb{Z}^{d}\setminus\{0\}, and nCloga0tn\geq C\log\frac{\lVert a_{0}\rVert}{t},

|(μnδx0)^(a0)|t.\lvert\widehat{(\mu^{*n}*\delta_{x_{0}})}(a_{0})\rvert\geq t.

Then, there exists γG/G\gamma\in G/G^{\circ} such that, denoting W0=(a0γE)W_{0}=(a_{0}\gamma E)^{\perp}, one has

d~(x0pqv,W0)enλ\tilde{d}\bigl{(}x_{0}-\frac{p}{q}-v,W_{0}\bigr{)}\leq e^{-n\lambda}

for some vV0v\in V_{0}, pdp\in\mathbb{Z}^{d} and q{0}q\in\mathbb{Z}\setminus\{0\} such that max(v,|q|)(a0t)C.\max(\lVert v\rVert,\lvert q\rvert)\leq\left(\frac{\lVert a_{0}\rVert}{t}\right)^{C}.

In the above, of course, pq\frac{p}{q}, vv and W0W_{0} are identified with their projection to the torus 𝕋d\mathbb{T}^{d}. A slightly more precise version of Theorem 1.1 is stated below as Theorem 6.1.

Remark.

If GG is connected, i.e. G=GG=G^{\circ}, then W0=(a0E)W_{0}=(a_{0}E)^{\perp} is entirely determined by a0a_{0}. Existence of a large Fourier coefficient (μnδx0)^(a0)\widehat{(\mu^{*n}*\delta_{x_{0}})}(a_{0}) implies that the starting point of the random walk is close to a rational translate of an invariant closed subset of the form W0+BV0(0,R)moddW_{0}+\operatorname{B}_{V_{0}}(0,R)\mod\mathbb{Z}^{d}, where BV0(0,R)\operatorname{B}_{V_{0}}(0,R) denotes the centered closed ball of radius RR in VV with respect to some GG-invariant Euclidean norm on VV and RR is controlled in terms of a0\lVert a_{0}\rVert and |(μnδx0)^(a0)|\lvert\widehat{(\mu^{*n}*\delta_{x_{0}})}(a_{0})\rvert.

Example (Reducible random walk).

Fix a probability measure μ0\mu_{0} on SL2()\operatorname{SL}_{2}(\mathbb{Z}) such that suppμ0\operatorname{supp}\mu_{0} generates SL2()\operatorname{SL}_{2}(\mathbb{Z}), and let μ=μ0μ0\mu=\mu_{0}\otimes\mu_{0}. Using the block diagonal embedding SL2×SL2SL4\operatorname{SL}_{2}\times\operatorname{SL}_{2}\hookrightarrow\operatorname{SL}_{4}, we view μ\mu as a probability measure on SL4()\operatorname{SL}_{4}(\mathbb{Z}). In that setting, E=M2()×M2()E=M_{2}(\mathbb{R})\times M_{2}(\mathbb{R}) and V0={0}V_{0}=\{0\}.
Assume as in the theorem that (μnδx0)^(a0)\widehat{(\mu^{*n}*\delta_{x_{0}})}(a_{0}) is large. Write a0=(a1,a2,a3,a4)a_{0}=(a_{1},a_{2},a_{3},a_{4}).

  • If (a1,a2)(a_{1},a_{2}) and (a3,a4)(a_{3},a_{4}) are both nonzero, then a0E=4a_{0}E=\mathbb{R}^{4} and therefore W0={0}W_{0}=\{0\}. Thus, the starting point x0x_{0} must be close to a rational point with small denominator.

  • If (a1,a2)=0(a_{1},a_{2})=0 and (a3,a4)0(a_{3},a_{4})\neq 0, then a0E={0}2a_{0}E=\{0\}\oplus\mathbb{R}^{2} so that W0=2{0}W_{0}=\mathbb{R}^{2}\oplus\{0\}. The theorem only allows us to conclude that x0x_{0} is close to a rational translate of the invariant subtorus 𝕋2×{0}=W0mod4\mathbb{T}^{2}\times\{0\}=W_{0}\mod\mathbb{Z}^{4}. In particular, we can conclude nothing about the first two coordinates of x0x_{0}. And indeed, if one starts from a point x0x_{0} whose third and fourth coordinates are zero, then the random walk is trapped in the proper invariant subset 𝕋2×{0}\mathbb{T}^{2}\times\{0\}. For every frequency of the form a0=(0,0,a3,a4)a_{0}=(0,0,a_{3},a_{4}) and for all nn, one has (μnδx0)^(a0)=1\widehat{(\mu^{*n}*\delta_{x_{0}})}(a_{0})=1.

Example (Compact factors and satellite measures).

Consider the quadratic form Q(x,y,z)=x2+y22z2Q(x,y,z)=x^{2}+y^{2}-\sqrt{2}z^{2} on 3\mathbb{R}^{3}, and SOQSL3\operatorname{SO}_{Q}\subset\operatorname{SL}_{3} its special orthogonal group:

SOQ={gSL3|tgJQg=JQ},whereJQ=diag(1,1,2).\operatorname{SO}_{Q}=\{g\in\operatorname{SL}_{3}\ |\ ^{t}\!gJ_{Q}g=J_{Q}\},\quad\mbox{where}\ J_{Q}=\mathrm{diag}(1,1,-\sqrt{2}).

If gg is an element of the group Γ=SOQ([2])\Gamma=\operatorname{SO}_{Q}(\mathbb{Z}[\sqrt{2}]) of elements of SOQ\operatorname{SO}_{Q} with entries in the ring [2]\mathbb{Z}[\sqrt{2}], one can write g=A+2Bg=A+\sqrt{2}B, with A,BA,B in M3()M_{3}(\mathbb{Z}). The map

g=A+2B(A2BBA)g=A+\sqrt{2}B\mapsto\begin{pmatrix}A&2B\\ B&A\end{pmatrix}

embeds Γ\Gamma into SL6()\operatorname{SL}_{6}(\mathbb{Z}). Let μ\mu be a probability measure on SL6()\operatorname{SL}_{6}(\mathbb{Z}) whose support generates the group Γ\Gamma.
Since (A2BBA)\begin{pmatrix}A&2B\\ B&A\end{pmatrix} is conjugate to diag(A+2B,A2B)\mathrm{diag}(A+\sqrt{2}B,A-\sqrt{2}B), so that Γ\Gamma preserves a direct sum decomposition 6=33\mathbb{R}^{6}=\mathbb{R}^{3}\oplus\mathbb{R}^{3}. The group Γ\Gamma acts on the second factor as a subgroup of SOQ¯\operatorname{SO}_{\bar{Q}}, where Q¯(x,y,z)=x2+y2+2z2\bar{Q}(x,y,z)=x^{2}+y^{2}+\sqrt{2}z^{2} is positive definite, so there is a non-trivial compact factor V03V_{0}\simeq\mathbb{R}^{3}. Note that V0V_{0} embeds densely in 𝕋6\mathbb{T}^{6}.
On the other hand, one can check that in that setting EE is conjugate under (2I32I3I3I3)\begin{pmatrix}\sqrt{2}I_{3}&-\sqrt{2}I_{3}\\ I_{3}&I_{3}\end{pmatrix} to the block diagonal subalgebra M3()×M3()M_{3}(\mathbb{R})\times M_{3}(\mathbb{R}). If a06{0}a_{0}\in\mathbb{Z}^{6}\setminus\{0\}, both projections of a0a_{0} to the 3\mathbb{R}^{3} factors are non-zero (note that the direct sum decomposition is not defined over \mathbb{Q}) and therefore one always has a0E=6a_{0}E=\mathbb{R}^{6}, whence W0={0}W_{0}=\{0\}.
Existence of a large Fourier coefficient (μnδx0)^(a0)\widehat{(\mu^{*n}*\delta_{x_{0}})}(a_{0}) implies that up to a rational translation with small denominator, the starting point x0x_{0} is close to the image in 𝕋6\mathbb{T}^{6} of a ball of controlled radius in V0V_{0}. Note that if the starting point x0x_{0} lies on the embedded leaf V0V_{0}, then the random walk equidistributes with respect to (the image in 𝕋6\mathbb{T}^{6} of) the uniform probability measure on the sphere containing x0x_{0} for the quadratic form x2+y2+2z2x^{2}+y^{2}+\sqrt{2}z^{2} on V0V_{0}.

If the sequence (μnδx0)(\mu^{*n}*\delta_{x_{0}}) does not converge to the Haar measure m𝕋d\operatorname{m}_{\mathbb{T}^{d}} in the weak-* topology, then, by Weyl’s equidistribution criterion, there are a0d{0}a_{0}\in\mathbb{Z}^{d}\setminus\{0\} and t>0t>0 such that |(μnδx0)^(a0)|t\lvert\widehat{(\mu^{*n}*\delta_{x_{0}})}(a_{0})\rvert\geq t for an unbounded sequence of nn\in\mathbb{N}. Letting nn go to infinity along this sequence, we deduce the following qualitative statement from the above theorem.

Corollary 1.2 (Qualitative statement).

Let μ\mu be a probability measure on GLd()\operatorname{GL}_{d}(\mathbb{Z}) having a finite exponential moment. Denote by GGLd()G\subset\operatorname{GL}_{d}(\mathbb{R}) the algebraic group generated by μ\mu. Assume that GG is semisimple. Then for any point x0𝕋dx_{0}\in\mathbb{T}^{d}, either

μnδx0m𝕋d,\mu^{*n}*\delta_{x_{0}}\rightharpoonup^{*}\operatorname{m}_{\mathbb{T}^{d}},

or

x0d+V0+W0modd,x_{0}\in\mathbb{Q}^{d}+V_{0}+W_{0}\mod\mathbb{Z}^{d},

where V0V_{0} denotes the sum of all compact factors of GG in d\mathbb{R}^{d} and W0W_{0} is a proper rational subspace of d\mathbb{R}^{d} invariant under the action of the identity component GG^{\circ} of GG.

As a consequence, we recover the classification of orbit closures due to Guivarc’h and Starkov [22] and Muchnik [34].

Corollary 1.3 (Classification of orbit closures).

Let ΓGLd()\Gamma\subset\operatorname{GL}_{d}(\mathbb{Z}) be a subgroup whose Zariski closure GG is semisimple. Let x0𝕋dx_{0}\in\mathbb{T}^{d}. Then the orbit closure Γx¯\overline{\Gamma x} is either the whole 𝕋d\mathbb{T}^{d} or contained in a Γ\Gamma-invariant closed subset of the form

1qd+BV0(0,R)+γG/GγW0modd\frac{1}{q}\mathbb{Z}^{d}+\operatorname{B}_{V_{0}}(0,R)+\bigcup_{\gamma\in G/G^{\circ}}\gamma W_{0}\mod\mathbb{Z}^{d}

where qq is a nonzero integer, BV0(0,R)\operatorname{B}_{V_{0}}(0,R) is a ball in V0V_{0}, the sum of all compact factors of GG in d\mathbb{R}^{d} and W0W_{0} is a proper rational subspace invariant under the action of the identity component GG^{\circ} of GG.

The qualitative statement could also be reformulated more simply as follows.

Corollary 1.4 (Equidistribution).

Let μ\mu be a probability measure on GLd()\operatorname{GL}_{d}(\mathbb{Z}) having a finite exponential moment. Denote by ΓGLd()\Gamma\subset\operatorname{GL}_{d}(\mathbb{Z}) the subgroup generated by μ\mu. Assume that the Zariski closure of Γ\Gamma is semisimple. Then for any x0𝕋dx_{0}\in\mathbb{T}^{d}, either μnδx0m𝕋d\mu^{*n}*\delta_{x_{0}}\rightharpoonup^{*}\operatorname{m}_{\mathbb{T}^{d}} or x0x_{0} is contained in a proper Γ\Gamma-invariant closed subset.

A particularly simple case of the above results is when the group Γ\Gamma acts strongly irreducibly on d\mathbb{Q}^{d}, that is, when Γ\Gamma preserves no nontrivial finite union of proper subspaces of d\mathbb{Q}^{d}. Then, for any a0d{0}a_{0}\in\mathbb{Z}^{d}\setminus\{0\} and any γG\gamma\in G, one must have a0γE=(d)a_{0}\gamma E=(\mathbb{R}^{d})^{*}, so we obtain a simpler equidistribution statement.

Corollary 1.5 (Equidistribution of d\mathbb{Q}^{d}-irreducible random walks).

Assume that GG is semisimple and acts strongly irreducibly on d\mathbb{Q}^{d}. Then for every λ(0,1)\lambda\in(0,1), there exist C=C(μ,λ)0C=C(\mu,\lambda)\geq 0 such that the following holds.

Given x0𝕋dx_{0}\in\mathbb{T}^{d}, assume that for some t(0,12)t\in(0,\frac{1}{2}), a0da_{0}\in\mathbb{Z}^{d}, and nCloga0tn\geq C\log\frac{\lVert a_{0}\rVert}{t},

|(μnδx0)^(a0)|t.\lvert\widehat{(\mu^{*n}*\delta_{x_{0}})}(a_{0})\rvert\geq t.

Then there exists vV0v\in V_{0}, pdp\in\mathbb{Z}^{d} and q{0}q\in\mathbb{Z}\setminus\{0\} such that max(v,|q|)(a0t)C\max(\lVert v\rVert,\lvert q\rvert)\leq\left(\frac{\lVert a_{0}\rVert}{t}\right)^{C} and

d~(x0pqv,0)enλ.\tilde{d}\bigl{(}x_{0}-\frac{p}{q}-v,0\bigr{)}\leq e^{-n\lambda}.

In particular, if x0x_{0} does not lie on a rational translate of the V0V_{0} leaf in 𝕋d\mathbb{T}^{d}, then μnδx0\mu^{*n}*\delta_{x_{0}} converges to m𝕋d\operatorname{m}_{\mathbb{T}^{d}}.

It was observed by Benoist and Quint [5, Corollary 1.4] that if GG is semisimple without compact factors and acts irreducibly on d\mathbb{Q}^{d}, then m𝕋d\operatorname{m}_{\mathbb{T}^{d}} is the only atom-free μ\mu-stationary probability measure on 𝕋d\mathbb{T}^{d}. By the results of [6], this implies that the Cesàro averages 1nk=0n1μkδx0\frac{1}{n}\sum_{k=0}^{n-1}\mu^{*k}*\delta_{x_{0}} converge to m𝕋d\operatorname{m}_{\mathbb{T}^{d}}. The above corollary immediately shows that convergence also holds without the averaging process. When μ\mu is a symmetric probability measure on SLd()\operatorname{SL}_{d}(\mathbb{Z}), a general result of Bénard [2, Theorem 1] implies this qualitative statement, but without the symmetry assumption the result seems to be new.

Corollary 1.6.

Assume that GG is semisimple without compact factors and acts strongly irreducibly on d\mathbb{Q}^{d}. Then, for every x0x_{0} irrational in 𝕋d\mathbb{T}^{d}, the sequence of measures (μnδx0)n0(\mu^{*n}*\delta_{x_{0}})_{n\geq 0} converges in law to m𝕋d\operatorname{m}_{\mathbb{T}^{d}}.

One motivation to carry out the rather technical proof presented here is its application to the spectral gap property for subgroups of algebraic groups, modulo arbitrary integers. Indeed, following a strategy of Bourgain and Varjú [15], one can use Theorem 1.1 to answer a question of Salehi Golsefidy and Varjú [35, Question 2]. A particular case of the problem was studied in [26], and we hope to generalize those results in a forthcoming paper.

1.1. Outline of the proof

The paper is entirely devoted to the proof of Theorem 1.1, for which we use the strategy introduced in [11], and more precisely the variant used in [25] to avoid the proximality assumption. Section 2 deals with discretized algebraic combinatorics in semisimple algebras: we prove some Fourier decay estimate for multiplicative convolutions of measures satisfying natural non-concentration conditions, Theorem 2.1, generalizing results of Bourgain [10] for the real line. The main input for our proof is a sum-product theorem for representations of real Lie groups [24, Theorem 1.1], which easily implies the discretized sum-product theorem in semisimple algebras; then we use some L2L^{2}-flattening lemma similar to the one used by Bourgain and Gamburd in their work on the spectral gap property.

After that, in order to apply the combinatorial results of the previous section to the random walk, we need to check that the measure μn\mu^{*n} appropriately rescaled is not concentrated near proper affine subspaces of EE, nor near singular elements; this is done in Section 3. Just as in [25], the argument ultimately relies on the spectral gap property modulo primes obtained by Salehi Golsefidy and Varjú [35]. However, because the rescaling automorphism is no longer a homothety, the proof involves a detailed analysis of the behavior of the random walk with respect to a quasi-norm on the algebra EE. To help the reader understand the main ideas of the proof without having to go through all the technical details, we start with the simpler case where EE is simple; even in that case, the argument is different and simpler than the one presented in [25], where similar estimates are needed.

In Section 4, we prove Theorem 4.2, an important Fourier decay estimate for the law of the random walk. This simply follows from a combination of the two previous sections when the group GG is connected, but becomes more complicated without this assumption. We follow the argument used in [28, Appendix B], with minor modifications.

Section 5 makes the link between the random walk on GG and the random walk on 𝕋d\mathbb{T}^{d}. The Fourier decay obtained in the previous section shows that if μnδx0\mu^{*n}*\delta_{x_{0}} has one large Fourier coefficient, then reducing slightly the value of nn, the measure μnδx0\mu^{*n}*\delta_{x_{0}} has many large Fourier coefficients. Using a quantitative version of Wiener’s lemma, one infers a first “granulation statement”: μnδx0\mu^{*n}*\delta_{x_{0}} is concentrated near a finite set of well-separated points in 𝕋d\mathbb{T}^{d}.

To conclude the proof of Theorem 1.1, we run backwards the random walk, starting from the granulation estimate mentioned above. The argument uses in particular the diophantine properties of the random walk, and the exponential unstability of closed invariant subsets, obtained using a drift function, as in Eskin-Margulis [19] or Benoist-Quint [5]. This is the content of Section 6.

1.2. Concluding remarks

Affine random walks. After some first results of J.-B. Boyer [16], it was explained in [27] how to obtain quantitative equidistribution of affine random walks on the torus, under the assumption that the action on d\mathbb{R}^{d} is strongly irreducible. The arguments in that paper could be adapted to our setting.


More general homogeneous spaces. Benoist and Quint [3, 5, 6] have obtained equidistribution results that are valid in the much more general setting of homogeneous spaces of Lie groups. One drawback is that their convergence theorems are not quantitative, and only concern the Cesàro averages 1nk=0n1μnδx0\frac{1}{n}\sum_{k=0}^{n-1}\mu^{*n}*\delta_{x_{0}}.

On this subject, the first author has obtained, in collaboration with Lakrec and Lindenstrauss, some partial results for affine random walks on nilmanifolds [28]; these spaces may be seen as the simplest generalization of tori, but the analysis already becomes much more intricate. Very recently, in collaboration with Bénard [17], using a new approach avoiding Fourier analysis, the first author has also been able to obtain results for random walks on finite-volume spaces of the form G/ΛG/\Lambda, where GG is SO(2,1)\operatorname{SO}(2,1) or SO(3,1)\operatorname{SO}(3,1), and Λ\Lambda a lattice in GG.

In a slightly different direction, W. Kim [29] studied effective equidistribution of expanding translates in the space of affine lattices. Also in a different direction, Lindenstrauss and Mohammadi [32], Yang [40], and Lindenstrauss, Mohammadi and Wang [33] have studied effective density and equidistribution in some homogeneous spaces. Although these equidistribution results do not deal with random walks, some of the techniques used there are similar enough to ours to be mentioned here.

1.3. Notation

Here is a list of notation we use.

  • fgf\ll g, gfg\gg f, f=O(g)f=O(g), there exists a constant C>0C>0 such that fCgf\leq Cg.

  • fgf\asymp g if fgf\ll g and gfg\ll f.

  • B(x,r)\operatorname{B}(x,r), the ball of center xx and radius rr.

  • BV(,)\operatorname{B}_{V}(\boldsymbol{\,\cdot\,},\boldsymbol{\,\cdot\,}), ball in the ambient space VV.

  • HH^{\circ}, the identity component with respect to the Zariski topology of the algebraic group HH.

  • VV^{*}, the space of linear forms on a linear space VV.

  • λ1(μ,V)\lambda_{1}(\mu,V), the top Lyapunov exponent associated to the random walk on a Euclidean space VV defined by a probability measure μ\mu supported on a group acting linearly on VV.

  • μν\mu*\nu, multiplicative convolution.

  • μk=μμ\mu^{*k}=\mu*\dotsm*\mu, multiplicative convolution power.

  • μν\mu\boxplus\nu, additive convolution.

  • μk=μμ\mu^{\boxplus k}=\mu\boxplus\dotsb\boxplus\mu, additive convolution power.

  • μν\mu\boxminus\nu, the image measure of μν\mu\otimes\nu under the map (x,y)xy(x,y)\mapsto x-y.

  • 𝟙A(x)=1\operatorname{\mathbbm{1}}_{A}(x)=1 if xAx\in A, 𝟙A(x)=0\operatorname{\mathbbm{1}}_{A}(x)=0 otherwise.

  • #A\#A, cardinality of a finite set AA.

  • |A|\lvert A\rvert, Lebesgue measure for subsets AA of an Euclidean space or a torus.

  • ||\scaleobj0.7\lvert\boldsymbol{\,\cdot\,}\rvert^{\scaleobj{0.7}{\sim}}, a quasi-norm

  • d~(,)\tilde{d}(\boldsymbol{\,\cdot\,},\boldsymbol{\,\cdot\,}), a quasi-distance, usually associated to a quasi-norm.

  • B~(,)\tilde{\operatorname{B}}(\boldsymbol{\,\cdot\,},\boldsymbol{\,\cdot\,}), ball with respect to d~\tilde{d}.

  • []\mathbb{P}[\boldsymbol{\,\cdot\,}] and []\mathbb{P}[\boldsymbol{\,\cdot\,}\mid\boldsymbol{\,\cdot\,}], probability and conditional probability.

  • fμf_{*}\mu, image measure of μ\mu under the map ff.

  • d()\operatorname{\mathcal{M}}_{d}(\mathbb{R}), the space of d×dd\times d real matrices.

  • 𝒫(X)\mathcal{P}(X), the space of Borel probability measure on a topological space.

  • ,\langle\boldsymbol{\,\cdot\,},\boldsymbol{\,\cdot\,}\rangle, according to the context, the natural pairing V×VV^{*}\times V\to\mathbb{R} or the natural pairing d×𝕋d𝕋\mathbb{Z}^{d}\times\mathbb{T}^{d}\to\mathbb{T}.

2. Sum-product, L2L^{2}-flattening and Fourier decay

In this section, we study multiplicative convolutions of measures on a semisimple associative algebra EE. Our goal is to derive Theorem 2.1 below, which shows that under some natural non-concentration assumptions, such multiplicative convolutions admit a polynomial Fourier decay. This generalizes results of Bourgain [10] for E=E=\mathbb{R}, of Li [31] for E=E=\mathbb{R}\oplus\dots\oplus\mathbb{R}, and of [25] for a simple algebra EE.


Let EE be a normed real algebra of finite dimension. The determinant detE(a)\det_{E}(a) of an element aEa\in E is simply defined as the determinant of the multiplication map EEE\to E, xaxx\mapsto ax. Given ρ>0\rho>0, we let

SE(ρ)={xE||detE(x)|ρ}.S_{E}(\rho)=\{\,x\in E\,|\,\lvert\det\nolimits_{E}(x)\rvert\leq\rho\,\}.

If WEW\subset E is any subset, we let W(ρ)W^{(\rho)} denote the ρ\rho-neighborhood of WW, defined by

W(ρ)={xE|wW:xw<ρ}.W^{(\rho)}=\{\,x\in E\,|\,\exists w\in W:\,\lVert x-w\rVert<\rho\,\}.

The following definition summarizes the non-concentration conditions we shall need in order to prove some Fourier decay for multiplicative convolutions.

Definition (Non-concentration conditions).

Let ε>0\varepsilon>0, κ>0\kappa>0, τ>0\tau>0 be parameters. We say a measure η\eta on EE satisfies NC0(ε,κ,τ)\operatorname{NC}_{0}(\varepsilon,\kappa,\tau) at scale δ>0\delta>0 if

  1. (1)

    suppηB(0,δε)\operatorname{supp}\eta\subset\operatorname{B}(0,\delta^{-\varepsilon});

  2. (2)

    for every xEx\in E, η(x+SE(δε))δτ\eta(x+S_{E}(\delta^{\varepsilon}))\leq\delta^{\tau};

  3. (3)

    for every ρ[δ,1]\rho\in[\delta,1] and every proper affine subspace WEW\subset E, η(W(ρ))δερκ\eta(W^{(\rho)})\leq\delta^{-\varepsilon}\rho^{\kappa}.

We say that a measure η\eta on EE satisfies NC(ε,κ,τ)\operatorname{NC}(\varepsilon,\kappa,\tau) at scale δ>0\delta>0 if it can be written as a sum of measures

η=η0+η1with{η0satisfyingNC0(ε,κ,τ)η1(E)δτ.\eta=\eta_{0}+\eta_{1}\quad\mbox{with}\quad\left\{\begin{array}[]{l}\eta_{0}\ \mbox{satisfying}\ \operatorname{NC}_{0}(\varepsilon,\kappa,\tau)\\ \eta_{1}(E)\leq\delta^{\tau}.\end{array}\right.

Given a finite measure μ\mu on EE, its Fourier transform μ^\hat{\mu} is the function on the dual space EE^{*} given by the expression

ξE,μ^(ξ)=Ee2iπξ,xdμ(x).\forall\xi\in E^{*},\quad\hat{\mu}(\xi)=\int_{E}e^{2i\pi\langle\xi,x\rangle}\,\mathrm{d}\mu(x).

If ν\nu is another finite measure on EE, the multiplicative convolution μν\mu*\nu is defined as the image measure of μν\mu\otimes\nu on E×EE\times E under the map (x,y)xy(x,y)\mapsto xy. It should not be confused with the additive convolution μν\mu\boxplus\nu, image of μν\mu\otimes\nu under the map (x,y)x+y(x,y)\mapsto x+y.

Theorem 2.1 (Fourier decay of multiplicative convolutions).

Let EE be a normed finite-dimensional semisimple algebra over \mathbb{R}. Given κ>0\kappa>0, there exists s=s(E,κ)s=s(E,\kappa)\in\mathbb{N} and ε=ε(E,κ)>0\varepsilon=\varepsilon(E,\kappa)>0 such that for any parameter τ(0,εκ)\tau\in{(0,\varepsilon\kappa)} the following holds for any scale δ>0\delta>0 sufficiently small.

If η1,,ηs\eta_{1},\dotsc,\eta_{s} are probability measures on EE satisfying NC(ε,κ,τ)\operatorname{NC}(\varepsilon,\kappa,\tau) at scale δ\delta, then for all ξE\xi\in E^{*} with δ1+εξδ1ε\delta^{-1+\varepsilon}\leq\lVert\xi\rVert\leq\delta^{-1-\varepsilon},

|(η1ηs)(ξ)|δετ.\lvert(\eta_{1}*\dotsm*\eta_{s})^{\wedge}(\xi)\rvert\leq\delta^{\varepsilon\tau}.

For E=E=\mathbb{R}, this is due to Bourgain [10, Lemma 8.43]. For algebras of the form E=E=\mathbb{R}\oplus\dotsb\oplus\mathbb{R}, this is due to Li [31, Theorem 1.1]. We shall first prove this theorem when all ηi\eta_{i} are equal, i.e. η1==ηs=η\eta_{1}=\dots=\eta_{s}=\eta and then deduce the general statement from this particular case following the argument in [27, Proof of Theorem B.3]. Alternatively, one could adapt the first part of the proof to handle directly the general case, but this would make notation cumbersome.

The proof we give for Theorem 2.1 follows a strategy originating in the work of Bourgain, Glibichuk and Konyagin [14] on exponential sums in finite fields: one deduces the bound on the exponential sum from a combinatorial “sum-product” statement, using an L2L^{2}-flattening statement. In our case, the combinatorial input is a discretized sum-product theorem in semisimple algebras, which follows from a general sum-product statement for representations of real Lie groups obtained in [24, Theorem 2.3].

2.1. Sum-product in semisimple algebras

Sum-product estimates go back to the work of Erdős and Szemerédi [18] who showed that there exists some positive constant ε\varepsilon such that for any subset AA of integers,

|A+A|+|AA||A|1+ε\lvert A+A\rvert+\lvert AA\rvert\geq\lvert A\rvert^{1+\varepsilon}

where A+AA+A and AAAA denote respectively the sum-set and the product-set of AA, defined by A+A={a+b;a,bA}A+A=\{a+b\ ;\ a,b\in A\} and AA={ab;a,bA}AA=\{ab\ ;\ a,b\in A\}. In the following, we consider a normed semisimple algebra EE of finite dimension over \mathbb{R}, and our goal is to prove a similar statement for subsets AEA\subset E, with the cardinality replaced by the covering number 𝒩(A,δ)\mathcal{N}(A,\delta) of AA at small scale δ>0\delta>0. Recall that by definition, 𝒩(A,δ)\mathcal{N}(A,\delta) is the minimal cardinality of a cover of AA by balls of radius δ\delta in EE. In order to ensure that the covering number of AA at scale δ\delta grows under addition or multiplication, one of course has to assume that AA is not essentially equal to a ball in some subalgebra of EE. We make a stronger assumption and require that AA is not concentrated near any proper affine subspace of EE.

Definition (Affine non-concentration).

Let VV be a Euclidean space, and ε,κ>0\varepsilon,\kappa>0 two parameters. We say a subset AVA\subset V satisfies ANC(ε,κ)\operatorname{ANC}(\varepsilon,\kappa) at scale δ\delta if

  1. (1)

    AB(0,δε)A\subset\operatorname{B}(0,\delta^{-\varepsilon}) and

  2. (2)

    for every ρδ\rho\geq\delta and every proper affine subspace WVW\subset V, 𝒩(AW(ρ),δ)δερκ𝒩(A,δ)\mathcal{N}(A\cap W^{(\rho)},\delta)\leq\delta^{-\varepsilon}\rho^{\kappa}\mathcal{N}(A,\delta).

Essentially, we want to show that if EE is a semisimple algebra, then for every κ>0\kappa>0, there exists ε>0\varepsilon>0 such that for any set ABE(0,1)A\subset\operatorname{B}_{E}(0,1) satisfying ANC(ε,κ)\operatorname{ANC}(\varepsilon,\kappa) and δκ𝒩(A,δ)δdimE+κ\delta^{-\kappa}\leq\mathcal{N}(A,\delta)\leq\delta^{-\dim E+\kappa}, one has 𝒩(A+A,δ)+𝒩(AAA,δ)δε𝒩(A,δ)\mathcal{N}(A+A,\delta)+\mathcal{N}(AAA,\delta)\geq\delta^{-\varepsilon}\mathcal{N}(A,\delta). We shall prove a slightly more technical growth statement, involving the tensor algebra EEopE\otimes E^{\mathrm{op}}, where EopE^{\mathrm{op}} denotes the algebra with the same linear structure as EE but with multiplication (a,b)ba(a,b)\mapsto ba. Note that the algebra EEopE\otimes E^{\mathrm{op}} acts naturally on EE by

a,xE,bEop,(ab)x=axb.\forall a,x\in E,\,\forall b\in E^{\mathrm{op}},\quad(a\otimes b)x=axb.
Theorem 2.2 (Sum-product in semisimple algebras).

Let EE be a finite-dimensional real semisimple algebra. Given κ>0\kappa>0, there exists ε=ε(E,κ)\varepsilon=\varepsilon(E,\kappa) such that the following holds for all δ>0\delta>0 sufficiently small.

  1. (1)

    Let AA be a subset of EE satisfying ANC(ε,κ)\operatorname{ANC}(\varepsilon,\kappa) at scale δ\delta and

  2. (2)

    δκ𝒩(A,δ)δdimE+κ\delta^{-\kappa}\leq\mathcal{N}(A,\delta)\leq\delta^{-\dim E+\kappa}.

  3. (3)

    Let BEEopB\subset E\otimes E^{\mathrm{op}} be a subset satisfying ANC(ε,κ)\operatorname{ANC}(\varepsilon,\kappa) at scale δ\delta.

Then there exists fBf\in B such that

𝒩(A+A,δ)+𝒩(A+fA,δ)δε𝒩(A,δ).\mathcal{N}(A+A,\delta)+\mathcal{N}(A+fA,\delta)\geq\delta^{-\varepsilon}\mathcal{N}(A,\delta).

The theorem above is almost equivalent to the fact that one can obtain from AA a small ball in EE using a bounded number of sums and products. This is the content of the proposition below, which we obtain as a simple application of [24, Theorem 2.3]. For a subset AA in an algebra EE and ss\in\mathbb{N}^{*}, we let As\langle A\rangle_{s} denote the set of elements in EE that can be obtained as sums of at most ss products of at most ss elements of AA or A-A.

Proposition 2.3 (Bounded generation in semisimple algebras).

Let EE be a finite-dimensional real semisimple algebra. Given κ>0\kappa>0 and ε0>0\varepsilon_{0}>0, there exists ε=ε(E,κ,ε0)>0\varepsilon=\varepsilon(E,\kappa,\varepsilon_{0})>0 and s=s(E,κ,ε0)1s=s(E,\kappa,\varepsilon_{0})\geq 1 such that the following holds for all δ>0\delta>0 sufficiently small. If AB(0,δε)A\subset\operatorname{B}(0,\delta^{-\varepsilon}) satisfies ANC(ε,κ)\operatorname{ANC}(\varepsilon,\kappa) at scale δ\delta in EE, then

B(0,δε0)As+B(0,δ).\operatorname{B}(0,\delta^{\varepsilon_{0}})\subset\langle A\rangle_{s}+\operatorname{B}(0,\delta).
Proof.

We consider the group G=E×G=E^{\times} of invertible elements in EE and its action by multiplication on V=EV=E. By semisimplicity, we may decompose EE into a sum of non-trivial irreducible representations E=iViE=\oplus_{i}V_{i}. Let πi:GGL(Vi)\pi_{i}\colon G\to\operatorname{GL}(V_{i}) denote the representation of GG on ViV_{i}. By [24, Theorem 2.3], there is a neighbourhood UU of the identity in GG and constants ε=ε(E,κ,ε0)>0\varepsilon=\varepsilon(E,\kappa,\varepsilon_{0})>0 and s=s(E,κ,ε)1s=s(E,\kappa,\varepsilon)\geq 1 such that the following holds for all δ>0\delta>0 sufficiently small. Let A0A_{0} be a subset of UU and A1A_{1} a subset of BV(0,1)\operatorname{B}_{V}(0,1). Assume

  1. (1)

    for all i=1,,ki=1,\dotsc,k, for all ρδ\rho\geq\delta, 𝒩(πi(A0),ρ)δερκ\mathcal{N}(\pi_{i}(A_{0}),\rho)\geq\delta^{\varepsilon}\rho^{-\kappa},

  2. (2)

    for any linear subspace WVW\subset V which is not GG-invariant, there is aA0a\in A_{0} such that d(a,StabG(W))δεd(a,\operatorname{Stab}_{G}(W)^{\circ})\geq\delta^{\varepsilon},

  3. (3)

    for any proper GG-invariant linear subspace WVW\subset V, there is aA1a\in A_{1} such that d(a,W)δεd(a,W)\geq\delta^{\varepsilon}.

Then

BV(0,δε0)A0,A1s+B(0,δ).\operatorname{B}_{V}(0,\delta^{\varepsilon_{0}})\subset\langle A_{0},A_{1}\rangle_{s}+\operatorname{B}(0,\delta).

Here, A0,A1s\langle A_{0},A_{1}\rangle_{s} denotes the set of elements in VV that can be obtained as sums of at most ss products of at most ss elements of A0A_{0} and elements of A1(A1)A_{1}\cup(-A_{1}). In the argument below, we apply this result with ε\varepsilon replaced by O(ε/κ)O(\varepsilon/\kappa).

Our set AA is not necessarily contained in the neighborhood UU, but we may cover AA by translates of UU in EE, and then, by the pigeonhole principle, there is aAa\in A such that A0=(Aa)UA_{0}=(A-a)\cap U satisfies

𝒩(A0,δ)UδO(ε)𝒩(A,δ).\mathcal{N}(A_{0},\delta)\gg_{U}\delta^{O(\varepsilon)}\mathcal{N}(A,\delta).

This set A0A_{0} satisfies ANC(O(ε),κ)\operatorname{ANC}(O(\varepsilon),\kappa) at scale δ\delta. This non-concentration condition applied to affine suspaces parallel to jiVj\oplus_{j\neq i}V_{j} shows that the first condition above is verified. Moreover, if WEW\subset E is not GG-invariant, then the algebra generated by StabG(W)\operatorname{Stab}_{G}(W) is a proper subalgebra of EE. In particular, it is included in a proper affine subspace of EE, and by ANC(O(ε),κ)\operatorname{ANC}(O(\varepsilon),\kappa), there must exist aa in A0A_{0} such that d(a,StabG(W))δO(ε/κ)d(a,\operatorname{Stab}_{G}(W))\geq\delta^{O(\varepsilon/\kappa)}; so the second condition is also satisfied. To conclude, take A1=A0A_{1}=A_{0}, which satisfies the third condition with ε\varepsilon replaced by O(ε/κ)O(\varepsilon/\kappa). ∎

In short, Theorem 2.2 will follow from Proposition 2.3 applied to the set BB in the tensor algebra EEopE\otimes E^{\mathrm{op}}, and from the Plünnecke-Ruzsa inequality.

Proof of Theorem 2.2.

For K1K\geq 1, define

Rδ(A,K)={fEEop|𝒩(A+fA,δ)K𝒩(A,δ)}.R_{\delta}(A,K)=\bigl{\{}\,f\in E\otimes E^{\mathrm{op}}\,\big{|}\,\mathcal{N}(A+fA,\delta)\leq K\mathcal{N}(A,\delta)\,\bigr{\}}.

Let us show that Rδ(A,K)R_{\delta}(A,K) is almost stable under addition and multiplication. By Ruzsa’s covering lemma, if fRδ(A,K)f\in R_{\delta}(A,K), there exists a set XfX_{f} such that 𝒩(Xf,δ)=O(K)\mathcal{N}(X_{f},\delta)=O(K) and

fAAA+Xf.fA\subset A-A+X_{f}.

Therefore, for f1,f2f_{1},f_{2} in δ(A,K)\mathbb{R}_{\delta}(A,K), one has

A+(f1+f2)AA+f1A+f2A3A2A+Xf1+Xf2.A+(f_{1}+f_{2})A\subset A+f_{1}A+f_{2}A\subset 3A-2A+X_{f_{1}}+X_{f_{2}}.

With the Plünnecke-Ruzsa inequality, this yields 𝒩(A+(f1+f2)A,δ)KO(1)𝒩(A,δ)\mathcal{N}(A+(f_{1}+f_{2})A,\delta)\leq K^{O(1)}\mathcal{N}(A,\delta), i.e. f1+f2f_{1}+f_{2} is in Rδ(A,KO(1))R_{\delta}(A,K^{O(1)}). Similarly, f1f2Rδ(A,KO(1))f_{1}f_{2}\in R_{\delta}(A,K^{O(1)}). By induction, this implies that for ss\in\mathbb{N},

Rδ(A,K)s+BEEop(0,δ)Rδ(A,KOs(1)).\langle R_{\delta}(A,K)\rangle_{s}+\operatorname{B}_{E\otimes E^{\mathrm{op}}}(0,\delta)\subset R_{\delta}(A,K^{O_{s}(1)}).

Now assume for a contradiction that B{1}Rδ(A,δε)B\cup\{1\}\subset R_{\delta}(A,\delta^{-\varepsilon}). Since EE is a semisimple algebra, EEopE\otimes E^{\mathrm{op}} is also one. Thus, by Proposition 2.3 applied to the set BEEopB\subset E\otimes E^{\mathrm{op}}, for any ε0>0\varepsilon_{0}>0, there is s=s(E,κ,ε0)1s=s(E,\kappa,\varepsilon_{0})\geq 1 such that

BEEop(0,δε0)Bs+B(0,δ)\operatorname{B}_{E\otimes E^{\mathrm{op}}}(0,\delta^{\varepsilon_{0}})\subset\langle B\rangle_{s}+\operatorname{B}(0,\delta)

and therefore,

BEEop(0,δε0)Bs+B(0,δ)Rδ(A,δOs(ε)).\operatorname{B}_{E\otimes E^{\mathrm{op}}}(0,\delta^{\varepsilon_{0}})\subset\langle B\rangle_{s}+\operatorname{B}(0,\delta)\subset R_{\delta}(A,\delta^{-O_{s}(\varepsilon)}).

In particular, δε0Rδ(A,δOs(ε))\delta^{\varepsilon_{0}}\in R_{\delta}(A,\delta^{-O_{s}(\varepsilon)}). This certainly implies δε0Rδ(A,δOs(ε0+ε))\delta^{-\varepsilon_{0}}\in R_{\delta}(A,\delta^{-O_{s}(\varepsilon_{0}+\varepsilon)}) and then, using once more stability of Rδ(A,K)R_{\delta}(A,K) under product,

BEEop(0,1)Rδ(A,δOs(ε0+ε)).\operatorname{B}_{E\otimes E^{\mathrm{op}}}(0,1)\subset R_{\delta}(A,\delta^{-O_{s}(\varepsilon_{0}+\varepsilon)}).

If ε0\varepsilon_{0} and ε\varepsilon are chosen small enough, this contradicts Lemma 2.4 below. ∎

We are left to show the next lemma, stating that if AA has ANC(ε,κ)\operatorname{ANC}(\varepsilon,\kappa) at scale δ\delta, then BEEop(0,1)\operatorname{B}_{E\otimes E^{\mathrm{op}}}(0,1) is not contained in Rδ(A,δε)R_{\delta}(A,\delta^{-\varepsilon}).

Lemma 2.4.

Let E=E1ErE=E_{1}\oplus\dots\oplus E_{r} be a finite-dimensional real semisimple algebra decomposed as a direct sum of minimal two-sided ideals. Write πj:EEj\pi_{j}\colon E\to E_{j} for the corresponding projections.

Given κ>0\kappa>0, there exists ε=ε(E,κ)>0\varepsilon=\varepsilon(E,\kappa)>0 such that the following holds for all δ>0\delta>0 sufficiently small. Let AB(0,δε)A\subset\operatorname{B}(0,\delta^{-\varepsilon}) be a subset of E. Assume

  1. (1)

    𝒩(A,δ)δdimE+κ\mathcal{N}(A,\delta)\leq\delta^{-\dim E+\kappa}

  2. (2)

    for each j=1,,rj=1,\dotsc,r, maxxEj𝒩(Aπj1(BEj(x,ρ)),δ)ρκ𝒩(A,δ)\max_{x\in E_{j}}\mathcal{N}(A\cap\pi_{j}^{-1}(\operatorname{B}_{E_{j}}(x,\rho)),\delta)\leq\rho^{\kappa}\mathcal{N}(A,\delta), where ρ=δκκ+dimE\rho=\delta^{\frac{\kappa}{\kappa+\dim E}}.

Then there exists fBEEop(0,1)f\in\operatorname{B}_{E\otimes E^{\mathrm{op}}}(0,1) such that

𝒩(A+fA,δ)>δε𝒩(A,δ).\mathcal{N}(A+fA,\delta)>\delta^{-\varepsilon}\mathcal{N}(A,\delta).
Proof.

The image of EEopE\otimes E^{\mathrm{op}} in End(E)\operatorname{End}(E) is equal to the image of j=1rEjEjop\bigoplus_{j=1}^{r}E_{j}\otimes E_{j}^{\mathrm{op}}. Let fjf_{j}, j=1,,rj=1,\dots,r be a family of jointly independent random elements of BEjEjop(0,1)\operatorname{B}_{E_{j}\otimes E_{j}^{\mathrm{op}}}(0,1) distributed according to the Lebesgue measure on EjEjopE_{j}\otimes E_{j}^{\mathrm{op}}, and set

f=f1++frf=f_{1}+\dots+f_{r}

regarded as a random element of End(E)\operatorname{End}(E). In the following argument, probabilities and expectations are taken with respect to these random variables. For each jj, since the algebra EjE_{j} is simple, the action of EjEjopE_{j}\otimes E_{j}^{\mathrm{op}} on EjE_{j} is irreducible. Hence EjEjop(y)=EjE_{j}\otimes E_{j}^{\mathrm{op}}(y)=E_{j} for any non-zero yEjy\in E_{j} and consequently,

(2.1) δ>0,x,yEj,[fj(y)xδ]δdimEjydimEj.\forall\delta>0,\,\forall x,y\in E_{j},\quad\mathbb{P}\bigl{[}\,\lVert f_{j}(y)-x\rVert\leq\delta\,\bigr{]}\ll\delta^{\dim E_{j}}\lVert y\rVert^{-\dim E_{j}}.

Consider the map

φ:A×AE(x,y)x+fy\begin{array}[]{llcl}\varphi\colon&A\times A&\to&E\\ &(x,y)&\mapsto&x+fy\end{array}

The energy of the map φ\varphi at scale δ>0\delta>0 is defined as

δ(φ,A×A)=𝒩({(a,a,b,b)A×A×A×A|φ(a,b)φ(a,b)δ},δ).\mathcal{E}_{\delta}(\varphi,A\times A)=\mathcal{N}\bigl{(}\{\,(a,a^{\prime},b,b^{\prime})\in A\times A\times A\times A\ \,|\,\lVert\varphi(a,b)-\varphi(a^{\prime},b^{\prime})\rVert\leq\delta\,\},\delta\bigr{)}.

By the Cauchy-Schwarz inequality — see also [23, Lemma 12(i)],

𝒩(φ(A×A),δ)=𝒩(A+fA,δ)𝒩(A,δ)4δ(φ,A×A).\mathcal{N}(\varphi(A\times A),\delta)=\mathcal{N}(A+fA,\delta)\geq\frac{\mathcal{N}(A,\delta)^{4}}{\mathcal{E}_{\delta}(\varphi,A\times A)}.

Taking expectations and applying Jensen’s inequality, we find

(2.2) 𝔼[𝒩(A+fA,δ)]𝒩(A,δ)4𝔼[δ(φ,A×A)]\mathbb{E}\bigl{[}\mathcal{N}(A+fA,\delta)\bigr{]}\geq\frac{\mathcal{N}(A,\delta)^{4}}{\mathbb{E}\bigl{[}\mathcal{E}_{\delta}(\varphi,A\times A)\bigr{]}}

so it suffices to bound 𝔼[δ(φ,A×A)]\mathbb{E}\bigl{[}\mathcal{E}_{\delta}(\varphi,A\times A)\bigr{]} from above.

For that, let A~\tilde{A} be a maximal δ\delta-separated subset of AA. By [23, Lemma 12(ii)],

𝔼[δ(φ,A×A)]x,y,x,yA~[f(yy)B(xx,5δ)].\mathbb{E}\bigl{[}\mathcal{E}_{\delta}(\varphi,A\times A)\bigr{]}\leq\sum_{x,y,x^{\prime},y^{\prime}\in\tilde{A}}\mathbb{P}\bigl{[}\,f(y^{\prime}-y)\in\operatorname{B}(x-x^{\prime},5\delta)\,\bigr{]}.

Let ρ=δκdimE+κ\rho=\delta^{\frac{\kappa}{\dim E+\kappa}}. We split the sum into two parts according to whether

j=1,,r,πj(yy)ρ.\forall j=1,\dotsc,r,\quad\lVert\pi_{j}(y^{\prime}-y)\rVert\geq\rho.

If this is the case, then (2.1) implies

[f(yy)B(xx,5δ)]δdimEρdimE.\mathbb{P}\bigl{[}\,f(y^{\prime}-y)\in\operatorname{B}(x-x^{\prime},5\delta)\,\bigr{]}\ll\delta^{\dim E}\rho^{-\dim E}.

Otherwise, there is j{1,,r}j\in\{1,\dotsc,r\} such that πj(y)B(πj(y),ρ)\pi_{j}(y^{\prime})\in\operatorname{B}(\pi_{j}(y),\rho). For fixed yy the number of such yy^{\prime} in A~\tilde{A} is

#(A~πj1(B(πj(y),ρ)))𝒩(Aπj1(B(πj(y),ρ)),δ)ρκ𝒩(A,δ).\#\bigl{(}\tilde{A}\cap\pi_{j}^{-1}(\operatorname{B}(\pi_{j}(y),\rho))\bigr{)}\ll\mathcal{N}(A\cap\pi_{j}^{-1}(\operatorname{B}(\pi_{j}(y),\rho)),\delta)\leq\rho^{\kappa}\mathcal{N}(A,\delta).

Moreover for fixed y,yy,y^{\prime} and xx, we have

xA~[f(yy)B(xx,5δ)]1\sum_{x^{\prime}\in\tilde{A}}\mathbb{P}\bigl{[}\,f(y^{\prime}-y)\in\operatorname{B}(x-x^{\prime},5\delta)\,\bigr{]}\ll 1

because the balls B(x,5δ)\operatorname{B}(x^{\prime},5\delta) have overlap multiplicity at most O(1)O(1). Putting these considerations together, we obtain

𝔼[δ(φ,A×A)]\displaystyle\mathbb{E}\bigl{[}\mathcal{E}_{\delta}(\varphi,A\times A)\bigr{]} δdimEρdimE𝒩(A,δ)4+ρκ𝒩(A,δ)3\displaystyle\ll\delta^{\dim E}\rho^{-\dim E}\mathcal{N}(A,\delta)^{4}+\rho^{\kappa}\mathcal{N}(A,\delta)^{3}
(δκρdimE+ρκ)𝒩(A,δ)3\displaystyle\leq\bigl{(}\delta^{\kappa}\rho^{-\dim E}+\rho^{\kappa}\bigr{)}\mathcal{N}(A,\delta)^{3}
δκ2dimE+κ𝒩(A,δ)3\displaystyle\ll\delta^{\frac{\kappa^{2}}{\dim E+\kappa}}\mathcal{N}(A,\delta)^{3}

Combined with (2.2), this finishes the proof of the lemma. ∎

2.2. L2L^{2}-flattening

Our goal is now to translate the sum-product theorem obtained above in terms of measures on the semisimple algebra EE. The result we obtain is an L2L^{2}-flattening lemma for additive and multiplicative convolutions of measures on EE. Statements of this form already appear implicitly in the work of Bourgain [9, 10] on the Erdős-Volkmann ring conjecture, and were later much popularized by their application to the spectral gap problem by Bourgain and Gamburd [13, 12]. They are usually derived from the analogous combinatorial growth statement, via a decomposition of the measures into dyadic level sets, combined with an application of the Balog-Szemerédi-Gowers lemma.

Before we can state our result, we give a non-concentration condition for measures on EE, analogous to the one given for subsets in the previous paragraph.

Definition (Affine non-concentration for measures).

Let VV be a Euclidean space, and ε,κ>0\varepsilon,\kappa>0 two parameters. We say that a measure η\eta on VV satisfies ANC(ε,κ)\operatorname{ANC}(\varepsilon,\kappa) at scale δ\delta if

  1. (1)

    suppηB(0,δε)\operatorname{supp}\eta\subset\operatorname{B}(0,\delta^{-\varepsilon});

  2. (2)

    for every ρδ\rho\geq\delta and every proper affine subspace WVW\subset V, η(W(ρ))δερκ\eta(W^{(\rho)})\leq\delta^{-\varepsilon}\rho^{\kappa}.

In this paper, measures are often studied at some fixed small positive scale δ\delta. For that reason, it is convenient to define the regularized measure ηδ\eta_{\delta} of a measure η\eta on EE at scale δ\delta by

ηδ=ηPδ\eta_{\delta}=\eta\boxplus P_{\delta}

where Pδ=𝟙B(0,δ)|B(0,δ)|P_{\delta}=\frac{\operatorname{\mathbbm{1}}_{\operatorname{B}(0,\delta)}}{\lvert\operatorname{B}(0,\delta)\rvert} is the normalized indicator function of the ball of radius δ\delta centered at 0. The measure ηδ\eta_{\delta} will be identified with its density with respect to the Lebesgue measure on EE, and we write

η2,δ=ηδ2.\lVert\eta\rVert_{2,\delta}=\lVert\eta_{\delta}\rVert_{2}.
Proposition 2.5 (L2L^{2}-flattening).

Let EE be finite-dimensional semisimple algebra over \mathbb{R}. Given κ>0\kappa>0, there exists ε=ε(E,κ)\varepsilon=\varepsilon(E,\kappa) such that the following holds for all δ>0\delta>0 sufficiently small. Let η\eta be a probability measure on EE satisfying

  1. (1)

    η\eta is supported on ESE(δε)E\setminus S_{E}(\delta^{\varepsilon});

  2. (2)

    η\eta satisfies ANC(ε,κ)\operatorname{ANC}(\varepsilon,\kappa) at scale δ\delta on EE;

  3. (3)

    δκ+εη2,δ2δdimE+κε\delta^{-\kappa+\varepsilon}\leq\lVert\eta\rVert_{2,\delta}^{2}\leq\delta^{-\dim E+\kappa-\varepsilon}.

Then,

ηηηηηη2,δδεη2,δ.\lVert\eta*\eta*\eta\boxminus\eta*\eta*\eta\rVert_{2,\delta}\leq\delta^{\varepsilon}\lVert\eta\rVert_{2,\delta}.

We wish to deduce this proposition from Theorem 2.2. A first useful observation is that the non-concentration condition for measures is closely related to non-concentration for subsets.

Lemma 2.6.

Given an Euclidean space VV, and parameters ϵ>0\epsilon>0 and κ>0\kappa>0, the following holds for all δ>0\delta>0 sufficiently small.

  1. (1)

    If AVA\subset V has ANC(ε,κ)\operatorname{ANC}(\varepsilon,\kappa) at scale δ\delta, then there is a measure supported on AA which has ANC(2ε,κ)\operatorname{ANC}(2\varepsilon,\kappa) at scale δ\delta.

  2. (2)

    Let η\eta be a probability measure on VV satisfying ANC(ε,κ)\operatorname{ANC}(\varepsilon,\kappa) at scale δ\delta. If AVA\subset V is a subset such that η(A)δε\eta(A)\geq\delta^{\varepsilon} then there is a subset AAA^{\prime}\subset A which satisfies ANC(6ε,κ)\operatorname{ANC}(6\varepsilon,\kappa) at scale δ\delta.

Proof.

For the first item, let A~\tilde{A} be a maximal δ\delta-separated subset of AA. The normalized counting measure on A~\tilde{A} satisfies the desired property. The second item is slightly more subtle. Since the normalized restriction of η\eta to AA satisfies ANC(2ε,κ)\operatorname{ANC}(2\varepsilon,\kappa), we may assume without loss of generality that A=suppηA=\operatorname{supp}\eta. Let imini_{\min} be the largest integer such that 2iminδ2εdimV2^{i_{\min}}\leq\delta^{2\varepsilon\dim V}. For every integer iimini\geq i_{\min}, set

Ai,0={aA| 2i1<η(B(a,2δ))|B(0,δ)|2i}.A_{i,0}=\Bigl{\{}\,a\in A\,\Big{|}\,2^{i-1}<\frac{\eta(\operatorname{B}(a,2\delta))}{\lvert\operatorname{B}(0,\delta)\rvert}\leq 2^{i}\,\Bigr{\}}.

and then

A,0=AiiminAi,0.A_{-,0}=A\setminus\bigcup_{i\geq i_{\min}}A_{i,0}.

Next, for every iimini\geq i_{\min}, set Ai=Ai,0(δ)A_{i}=A_{i,0}^{(\delta)} and also A=A,0(δ)A_{-}=A_{-,0}^{(\delta)}. By this construction,

(2.3) ηδδ2εdimV𝟙A+iimin2i𝟙Ai\eta_{\delta}\ll\delta^{2\varepsilon\dim V}\operatorname{\mathbbm{1}}_{A_{-}}+\sum_{i\geq i_{\min}}2^{i}\operatorname{\mathbbm{1}}_{A_{i}}

and

(2.4) iimin,2i𝟙Aiη3δ.\forall i\geq i_{\min},\quad 2^{i}\operatorname{\mathbbm{1}}_{A_{i}}\ll\eta_{3\delta}.

Note that AiA_{i} is empty whenever ilog|B(0,δ)|log2+1i\geq-\frac{\log\lvert\operatorname{B}(0,\delta)\rvert}{\log 2}+1. Thus, integrating (2.3) and recalling suppηB(0,δε)\operatorname{supp}\eta\subset B(0,\delta^{-\varepsilon}) from ANC(ε,κ)\operatorname{ANC}(\varepsilon,\kappa) for η\eta, we obtain some iimini\geq i_{\min} such that 2i|Ai|δε.2^{i}\lvert A_{i}\rvert\geq\delta^{\varepsilon}. Fix this ii and set A=Ai,0A^{\prime}=A_{i,0}. If WW is a proper affine subspace in VV and ρδ\rho\geq\delta, we can bound using (2.4) and ANC(ε,κ)\operatorname{ANC}(\varepsilon,\kappa) for η\eta,

𝒩(AW(ρ),δ)\displaystyle\mathcal{N}(A^{\prime}\cap W^{(\rho)},\delta) δdimVV𝟙AiW(ρ)(x)dx\displaystyle\ll\delta^{-\dim V}\int_{V}~\operatorname{\mathbbm{1}}_{A_{i}\cap W^{(\rho)}}(x)\,\mathrm{d}x
δdimVV2i𝟙W(ρ)(x)dη3δ(x)\displaystyle\ll\delta^{-\dim V}\int_{V}~2^{-i}\operatorname{\mathbbm{1}}_{W^{(\rho)}}(x)\,\mathrm{d}\eta_{3\delta}(x)
=δdimV2iη3δ(W(ρ))\displaystyle=\delta^{-\dim V}2^{-i}\eta_{3\delta}(W^{(\rho)})
δdimV2iδερκ\displaystyle\leq\delta^{-\dim V}2^{-i}\delta^{-\varepsilon}\rho^{\kappa}

and using the above lower bound on 2i|Ai|2^{i}\lvert A_{i}\rvert, we get

𝒩(AW(ρ),δ)\displaystyle\mathcal{N}(A^{\prime}\cap W^{(\rho)},\delta) δdimV|Ai|δ2ερκ\displaystyle\ll\delta^{-\dim V}\lvert A_{i}\rvert\delta^{-2\varepsilon}\rho^{\kappa}
δ2ερκ𝒩(Ai,δ)\displaystyle\ll\delta^{-2\varepsilon}\rho^{\kappa}\mathcal{N}(A_{i},\delta)

This shows that AA^{\prime} satisfies ANC(3ε,κ)\operatorname{ANC}(3\varepsilon,\kappa) at scale δ\delta. ∎

The next lemma is similar in spirit to the previous one. Roughly speaking, given measures η\eta on VV and μ\mu on GL(V)\operatorname{GL}(V) such that the convolution μημη\mu*\eta\boxminus\mu*\eta has large L2L^{2}-norm at scale δ\delta, we construct related subsets AVA\subset V and BGL(V)B\subset\operatorname{GL}(V) such that ABAA-BA is not much larger than AA. This is the central part of the proof of Proposition 2.5; it relies on the Balog-Szemerédi-Gowers lemma.

Lemma 2.7.

Let VV be a Euclidean space and μ\mu a probability measure on GL(V)\operatorname{GL}(V) such that

gsuppμ,g+g1δε.\forall g\in\operatorname{supp}\mu,\quad\lVert g\rVert+\lVert g^{-1}\rVert\leq\delta^{-\varepsilon}.

Let η\eta be a probability measure on BV(0,δε)\operatorname{B}_{V}(0,\delta^{-\varepsilon}) such that

μημη2,δ>δεη2,δ.\lVert\mu*\eta\boxminus\mu*\eta\rVert_{2,\delta}>\delta^{\varepsilon}\lVert\eta\rVert_{2,\delta}.

Then there exist a subset ABV(0,δO(ε))A\subset\operatorname{B}_{V}(0,\delta^{-O(\varepsilon)}) and an element g1suppμg_{1}\in\operatorname{supp}\mu such that

δdimV+O(ε)η2,δ2𝒩(A,δ)δdimVO(ε)η2,δ2\delta^{-\dim V+O(\varepsilon)}\lVert\eta\rVert_{2,\delta}^{-2}\leq\mathcal{N}(A,\delta)\leq\delta^{-\dim V-O(\varepsilon)}\lVert\eta\rVert_{2,\delta}^{-2}

and

μ({gGL(V)|𝒩(Agg11A,δ)δO(ε)𝒩(A,δ)})δO(ε).\mu\bigl{(}\bigl{\{}\,g\in\operatorname{GL}(V)\,\big{|}\,\mathcal{N}(A-gg_{1}^{-1}A,\delta)\leq\delta^{-O(\varepsilon)}\mathcal{N}(A,\delta)\,\bigr{\}}\bigr{)}\geq\delta^{O(\varepsilon)}.

If moreover η\eta satisfies ANC(ε,κ)\operatorname{ANC}(\varepsilon,\kappa) in VV at scale δ\delta for some κ>0\kappa>0 then AA satisfies ANC(O(ε),κ)\operatorname{ANC}(O(\varepsilon),\kappa).

Proof.

We use the following rough comparison notation : for positive quantities ff and gg, we write fgf\lesssim g for fδO(ε)gf\leq\delta^{-O(\varepsilon)}g and fgf\sim g for fgf\lesssim g and gfg\lesssim f.

We have

μηδμηδ2μημη2,δηδ2.\lVert\mu*\eta_{\delta}\boxminus\mu*\eta_{\delta}\rVert_{2}\gtrsim\lVert\mu*\eta\boxminus\mu*\eta\rVert_{2,\delta}\gtrsim\lVert\eta_{\delta}\rVert_{2}.

As in the proof of Lemma 2.6, we can approximate ηδ\eta_{\delta} using dyadic level sets : there are δ\delta-discretized sets111A δ\delta-discretized set is a union of balls of radius δ\delta. (Ai)i0(A_{i})_{i\geq 0} in BV(0,δε)\operatorname{B}_{V}(0,\delta^{-\varepsilon}) such that AiA_{i} is empty for ilog1δi\gg\log\frac{1}{\delta} and

(2.5) ηδi02i𝟙Aiη3δ+𝟙A0.\eta_{\delta}\ll\sum_{i\geq 0}2^{i}\operatorname{\mathbbm{1}}_{A_{i}}\lesssim\eta_{3\delta}+\operatorname{\mathbbm{1}}_{A_{0}}.

By the pigeonhole principle, there are i,j0i,j\geq 0 such that

ηδ2\displaystyle\lVert\eta_{\delta}\rVert_{2} μηδμηδ2\displaystyle\lesssim\lVert\mu*\eta_{\delta}\boxminus\mu*\eta_{\delta}\rVert_{2}
2i+jμ𝟙Aiμ𝟙Aj2\displaystyle\lesssim 2^{i+j}\lVert\mu*\operatorname{\mathbbm{1}}_{A_{i}}\boxminus\mu*\operatorname{\mathbbm{1}}_{A_{j}}\rVert_{2}
2i+jGL(V)×GL(V)𝟙gAi𝟙gAj2d(μμ)(g,g).\displaystyle\lesssim 2^{i+j}\int_{\operatorname{GL}(V)\times\operatorname{GL}(V)}\lVert\operatorname{\mathbbm{1}}_{gA_{i}}\boxminus\operatorname{\mathbbm{1}}_{g^{\prime}A_{j}}\rVert_{2}\,\mathrm{d}(\mu\otimes\mu)(g,g^{\prime}).

In the last inequality, we used g𝟙Ai=|detg|1𝟙gAig*\operatorname{\mathbbm{1}}_{A_{i}}=\lvert\det{g}\rvert^{-1}\operatorname{\mathbbm{1}}_{gA_{i}} and |detg|1\lvert\det g\rvert\sim 1 for all gsuppμg\in\operatorname{supp}\mu. By the right-hand inequality in (2.5), we have

2i|Ai|1and2j|Aj|12ηδ22^{i}\lvert A_{i}\rvert\lesssim 1\quad\text{and}\quad 2^{j}\lvert A_{j}\rvert^{\frac{1}{2}}\lesssim\lVert\eta_{\delta}\rVert_{2}

and similarly

2j|Aj|1and2i|Ai|12ηδ2.2^{j}\lvert A_{j}\rvert\lesssim 1\quad\text{and}\quad 2^{i}\lvert A_{i}\rvert^{\frac{1}{2}}\lesssim\lVert\eta_{\delta}\rVert_{2}.

By Young’s inequality and the estimate on detg\det g, we have for all g,gsuppμg,g^{\prime}\in\operatorname{supp}\mu,

2i+j𝟙gAi𝟙gAj2ηδ2.2^{i+j}\lVert\operatorname{\mathbbm{1}}_{gA_{i}}\boxminus\operatorname{\mathbbm{1}}_{g^{\prime}A_{j}}\rVert_{2}\lesssim\lVert\eta_{\delta}\rVert_{2}.

Thus, by the pigeonhole principle again, there exists g0suppμg_{0}\in\operatorname{supp}\mu and a set B0suppμB_{0}\subset\operatorname{supp}\mu such that μ(B0)1\mu(B_{0})\gtrsim 1 and for all gB0g\in B_{0},

ηδ22i+j𝟙g0Ai𝟙gAj2ηδ2.\lVert\eta_{\delta}\rVert_{2}\gtrsim 2^{i+j}\lVert\operatorname{\mathbbm{1}}_{g_{0}A_{i}}\boxminus\operatorname{\mathbbm{1}}_{gA_{j}}\rVert_{2}\gtrsim\lVert\eta_{\delta}\rVert_{2}.

By the above estimates, this implies

𝟙g0Ai𝟙gAj22\displaystyle\lVert\operatorname{\mathbbm{1}}_{g_{0}A_{i}}\boxminus\operatorname{\mathbbm{1}}_{gA_{j}}\rVert_{2}^{2} 22i2jηδ22\displaystyle\gtrsim 2^{-2i-2j}\lVert\eta_{\delta}\rVert_{2}^{2}
2ij|Ai|12|Aj|12\displaystyle\gtrsim 2^{-i-j}\lvert A_{i}\rvert^{\frac{1}{2}}\lvert A_{j}\rvert^{\frac{1}{2}}
|Ai|32|Aj|32\displaystyle\gtrsim\lvert A_{i}\rvert^{\frac{3}{2}}\lvert A_{j}\rvert^{\frac{3}{2}}
|g0Ai|32|gAj|32.\displaystyle\sim\lvert g_{0}A_{i}\rvert^{\frac{3}{2}}\lvert gA_{j}\rvert^{\frac{3}{2}}.

By the Balog-Szemerédi-Gowers lemma [37, Theorem 6.10], for each gB0g\in B_{0} there are δ\delta-discretized subsets AgAiA_{g}\subset A_{i} and AgAjA^{\prime}_{g}\subset A_{j} such that

|Ag||Ai|,|Ag||Aj|, and𝒩(g0AggAg,δ)𝒩(g0Ag,δ)12𝒩(gAg,δ)12.\lvert A_{g}\rvert\sim\lvert A_{i}\rvert,\,\lvert A^{\prime}_{g}\rvert\sim\lvert A_{j}\rvert\text{, and}\quad\mathcal{N}(g_{0}A_{g}-gA^{\prime}_{g},\delta)\lesssim\mathcal{N}(g_{0}A_{g},\delta)^{\frac{1}{2}}\mathcal{N}(gA^{\prime}_{g},\delta)^{\frac{1}{2}}.

Set X=Ai×AjX=A_{i}\times A_{j} and Xg=Ag×AgXX_{g}=A_{g}\times A^{\prime}_{g}\subset X and write

|Xg1Xg2|dμ(g1)dμ(g2)\displaystyle\iint\lvert X_{g_{1}}\cap X_{g_{2}}\rvert\,\mathrm{d}\mu(g_{1})\,\mathrm{d}\mu(g_{2}) =𝟙Xg1(x)𝟙Xg2(x)dxdμ(g1)dμ(g2)\displaystyle=\iint\int\operatorname{\mathbbm{1}}_{X_{g_{1}}}(x)\operatorname{\mathbbm{1}}_{X_{g_{2}}}(x)\,\mathrm{d}x\,\mathrm{d}\mu(g_{1})\,\mathrm{d}\mu(g_{2})
=(𝟙Xg(x)dμ(g))2dx\displaystyle=\int\left(\int\operatorname{\mathbbm{1}}_{X_{g}(x)}\,\mathrm{d}\mu(g)\right)^{2}\,\mathrm{d}x
1|X|(𝟙Xg(x)dxdμ(g))2\displaystyle\geq\frac{1}{\lvert X\rvert}\left(\iint\operatorname{\mathbbm{1}}_{X_{g}}(x)\,\mathrm{d}x\,\mathrm{d}\mu(g)\right)^{2}
|X|.\displaystyle\gtrsim\lvert X\rvert.

This shows that there exists g1g_{1} and B1B0B_{1}\subset B_{0} such that μ(B1)1\mu(B_{1})\gtrsim 1 for all gg in B1B_{1}, |Xg1Xg||X|\lvert X_{g_{1}}\cap X_{g}\rvert\gtrsim\lvert X\rvert. Equivalently,

gB1,|Ag1Ag||Ag1||Ag| and |Ag1Ag||Ag1||Ag|.\forall g\in B_{1},\quad\lvert A_{g_{1}}\cap A_{g}\rvert\sim\lvert A_{g_{1}}\rvert\sim\lvert A_{g}\rvert\text{ and }\lvert A^{\prime}_{g_{1}}\cap A^{\prime}_{g}\rvert\sim\lvert A^{\prime}_{g_{1}}\rvert\sim\lvert A^{\prime}_{g}\rvert.

For subsets A,AVA,A^{\prime}\in V, write AAA\approx A^{\prime} if 𝒩(AA,δ)𝒩(A,δ)12𝒩(A,δ)12\mathcal{N}(A-A^{\prime},\delta)\lesssim\mathcal{N}(A,\delta)^{\frac{1}{2}}\mathcal{N}(A^{\prime},\delta)^{\frac{1}{2}}. Ruzsa’s triangle inequality [38, Lemma 2.6] is still valid for covering numbers at scale δ\delta, so if AAA\approx A^{\prime} and AA′′A^{\prime}\approx A^{\prime\prime}, then AA′′A\approx A^{\prime\prime}, and moreover, if AAA\approx A^{\prime} for some sets AA and AA^{\prime}, then AAA\approx A and AAA^{\prime}\approx A^{\prime}.

The above shows that for every gB0g\in B_{0}, g0AggAgg_{0}A_{g}\approx gA_{g}^{\prime}. This implies g0Agg0Agg_{0}A_{g}\approx g_{0}A_{g}, and since gg is δε\delta^{-\varepsilon}-Lipschitz, AgAgA_{g}\approx A_{g}. Therefore, for g1B0g_{1}\in B_{0} and gB1g\in B_{1} as above, we find Ag1Ag1AgAgA_{g_{1}}\approx A_{g_{1}}\cap A_{g}\approx A_{g}. Similarly, Ag1AgA^{\prime}_{g_{1}}\approx A^{\prime}_{g},. Finally

g0Ag1g0AggAggAg1gg11g0Ag1,g_{0}A_{g_{1}}\approx g_{0}A_{g}\approx gA^{\prime}_{g}\approx gA^{\prime}_{g_{1}}\approx gg_{1}^{-1}g_{0}A_{g_{1}},

showing that A=g0Ag1A=g_{0}A_{g_{1}} has all the desired properties.

For last assertion, note that η(Ai)1\eta(A_{i})\gtrsim 1. By the proof of Lemma 2.62, AiA_{i} satisfies ANC(O(ε),κ)\operatorname{ANC}(O(\varepsilon),\kappa) and hence so do Ag1A_{g_{1}} and AA. Note also that

𝒩(A,δ)𝒩(Ag1,δ)𝒩(Ai,δ)δdimV|Ai|δdimVη2,δ2.\mathcal{N}(A,\delta)\sim\mathcal{N}(A_{g_{1}},\delta)\sim\mathcal{N}(A_{i},\delta)\sim\delta^{-\dim V}\lvert A_{i}\rvert\sim\delta^{-\dim V}\lVert\eta\rVert_{2,\delta}^{2}.

To prove Proposition 2.5 we use the above lemma for the action of EEopE\otimes E^{\mathrm{op}} on EE, and then apply the sum-product theorem in EE.

Proof of Proposition 2.5.

Let μ\mu be the image measure of ηη\eta\otimes\eta in GL(E)\operatorname{GL}(E), so that μημη=ηηηηηη\mu*\eta\boxminus\mu*\eta=\eta*\eta*\eta\boxminus\eta*\eta*\eta. We argue by contradiction: Assuming

μημη2,δ>δεη2,δ,\lVert\mu*\eta\boxminus\mu*\eta\rVert_{2,\delta}>\delta^{\varepsilon}\lVert\eta\rVert_{2,\delta},

we shall construct sets A,BA,B satisfying all assumptions of Theorem 2.2, with κ\kappa and ε\varepsilon replaced by κ/2\kappa/2 and O(ε)O(\varepsilon) but violating its conclusion. By Lemma 2.7 there is a subset AEA\subset E satisfying ANC(O(ε),κ)\operatorname{ANC}(O(\varepsilon),\kappa) at scale δ\delta and an element g1suppμg_{1}\in\operatorname{supp}\mu such that

δκ+O(ε)𝒩(A,δ)δdimE+κO(ε).\delta^{-\kappa+O(\varepsilon)}\leq\mathcal{N}(A,\delta)\leq\delta^{-\dim E+\kappa-O(\varepsilon)}.

and

μ({gGL(V)|𝒩(Agg11A,δ)δO(ε)𝒩(A,δ)})δO(ε).\mu\bigl{(}\bigl{\{}\,g\in\operatorname{GL}(V)\,\big{|}\,\mathcal{N}(A-gg_{1}^{-1}A,\delta)\leq\delta^{-O(\varepsilon)}\mathcal{N}(A,\delta)\,\bigr{\}}\bigr{)}\geq\delta^{O(\varepsilon)}.

By definition of μ\mu, we may write g1=a1b1g_{1}=a_{1}\otimes b_{1}, and the above inequality becomes

(2.6) (ηa11)(b11η)({(a,b)E×E|𝒩(AaAb,δ)δO(ε)𝒩(A,δ)})δO(ε).(\eta*a_{1}^{-1})\otimes(b_{1}^{-1}*\eta)\bigl{(}\bigl{\{}\,(a,b)\in E\times E\,\big{|}\,\mathcal{N}(A-aAb,\delta)\leq\delta^{-O(\varepsilon)}\mathcal{N}(A,\delta)\,\bigr{\}}\bigr{)}\geq\delta^{O(\varepsilon)}.

Since a1,b1SE(δε)a_{1},b_{1}\not\in S_{E}(\delta^{\varepsilon}), the measures ηa11\eta*a_{1}^{-1} and b11ηb_{1}^{-1}*\eta satisfy ANC(O(ε),κ)\operatorname{ANC}(O(\varepsilon),\kappa) at scale δ\delta. Moreover, Lemma 2.8 below shows that (ηa11)˙(b11η)(\eta*a_{1}^{-1})\dot{\otimes}(b_{1}^{-1}*\eta) satisfies ANC(O(ε),κ2)\operatorname{ANC}(O(\varepsilon),\frac{\kappa}{2}). By Lemma 2.62, there exists a subset Bsupp((ηa11)˙(b11η))B\subset\operatorname{supp}((\eta*a_{1}^{-1})\dot{\otimes}(b_{1}^{-1}*\eta)) satisfying ANC(O(ε),κ2)\operatorname{ANC}(O(\varepsilon),\frac{\kappa}{2}). Equation (2.6) shows that 𝒩(A+fA,δ)δO(ε)𝒩(A,δ)\mathcal{N}(A+fA,\delta)\leq\delta^{-O(\varepsilon)}\mathcal{N}(A,\delta) for all ff in BB. Since η\eta is supported on ESE(δε)E\setminus S_{E}(\delta^{\varepsilon}), one also has 𝒩(fA,δ)δO(ε)𝒩(A,δ)\mathcal{N}(fA,\delta)\geq\delta^{O(\varepsilon)}\mathcal{N}(A,\delta), and so by Plünnecke’s inequality one also has 𝒩(A+A,δ)δO(ε)𝒩(A,δ)\mathcal{N}(A+A,\delta)\leq\delta^{-O(\varepsilon)}\mathcal{N}(A,\delta) and in turn.

𝒩(A+A,δ)+𝒩(A+fA,δ)δO(ε)𝒩(A,δ).\mathcal{N}(A+A,\delta)+\mathcal{N}(A+fA,\delta)\leq\delta^{-O(\varepsilon)}\mathcal{N}(A,\delta).

Thus, AA and BB violate the conclusion of Theorem 2.2 with parameters κ/2\kappa/2 and O(ε)O(\varepsilon). This yields the desired contradiction, provided ε\varepsilon is chosen small enough. ∎

Lemma 2.8.

Let V1V_{1} and V2V_{2} be finite-dimensional linear spaces. For each i=1,2i=1,2, let ηi\eta_{i} be a measure on ViV_{i} and denote by η1˙η2\eta_{1}\dot{\otimes}\eta_{2} the image measure of η1η2\eta_{1}\otimes\eta_{2} by the natural bilinear map V1×V2V1V2V_{1}\times V_{2}\to V_{1}\otimes V_{2}.

Given two parameters ε,κ>0\varepsilon,\kappa>0, the following holds for δ>0\delta>0 sufficiently small. If η1\eta_{1} and η2\eta_{2} both satisfy ANC(ε,κ)\operatorname{ANC}(\varepsilon,\kappa) at scale δ\delta, then η1˙η2\eta_{1}\dot{\otimes}\eta_{2} satisfies ANC(2ε,κ2)\operatorname{ANC}(2\varepsilon,\frac{\kappa}{2}) in V1V2V_{1}\otimes V_{2} at scale δ\delta.

Proof.

Let v1v_{1} and v2v_{2} be independent random variables taking values respectively in V1V_{1} and V2V_{2} and distributed according to η1\eta_{1} and η2\eta_{2}. To establish ANC(2ε,κ2)\operatorname{ANC}(2\varepsilon,\frac{\kappa}{2}) for η1˙η2\eta_{1}\dot{\otimes}\eta_{2}, it is enough to show that for any linear form φ(V1V2)\varphi\in(V_{1}\otimes V_{2})^{*} with φ=1\lVert\varphi\rVert=1, any tt\in\mathbb{R} and any ρδ\rho\geq\delta, we have

(2.7) [|φ(v1v2)t|<ρ]δερκ/2.\mathbb{P}\bigl{[}\,\lvert\varphi(v_{1}\otimes v_{2})-t\rvert<\rho\,\bigr{]}\ll\delta^{-\varepsilon}\rho^{\kappa/2}.

Note that (V1V2)=V1V2(V_{1}\otimes V_{2})^{*}=V_{1}^{*}\otimes V_{2}^{*}. Hence, letting (ψ1,,ψd)(\psi_{1},\dotsc,\psi_{d}) be a orthonormal basis of V1V_{1}^{*}, we can write φ(V1V2)\varphi\in(V_{1}\otimes V_{2})^{*} as

φ=ψ1φ1++ψdφd\varphi=\psi_{1}\otimes\varphi_{1}+\dots+\psi_{d}\otimes\varphi_{d}

where φ1,,φdV2\varphi_{1},\dotsc,\varphi_{d}\in V_{2}^{*} are uniquely determined. Moreover,

(2.8) 1=φ2=φ12++φd2.1=\lVert\varphi\rVert^{2}=\lVert\varphi_{1}\rVert^{2}+\dots+\lVert\varphi_{d}\rVert^{2}.

On the one hand, when v2v_{2} is fixed, the map

v1φ(v1v2)=i=1dψi(v1)φi(v2)v_{1}\mapsto\varphi(v_{1}\otimes v_{2})=\sum_{i=1}^{d}\psi_{i}(v_{1})\varphi_{i}(v_{2})

is the linear form i=1dφi(v2)ψiV1\sum_{i=1}^{d}\varphi_{i}(v_{2})\psi_{i}\in V_{1}^{*}, which has norm i=1d|φi(v2)|2\sum_{i=1}^{d}\lvert\varphi_{i}(v_{2})\rvert^{2}. Thus, by independence of v1v_{1} and v2v_{2} and property ANC(ε,κ)\operatorname{ANC}(\varepsilon,\kappa) for η1\eta_{1}, we can estimate the conditional probability

(2.9) [|φ(v1v2)t|<ρ|i=1d|φi(v2)|2ρ1/2]δερκ/2.\mathbb{P}\Bigl{[}\,\lvert\varphi(v_{1}\otimes v_{2})-t\rvert<\rho\,\Big{|}\,\sum_{i=1}^{d}\lvert\varphi_{i}(v_{2})\rvert^{2}\geq\rho^{1/2}\,\Bigr{]}\leq\delta^{-\varepsilon}\rho^{\kappa/2}.

On the other hand, by property ANC(ε,κ)\operatorname{ANC}(\varepsilon,\kappa) for η2\eta_{2}, for each i=1,,di=1,\dotsc,d,

[|φi(v2)|ρ1/2φi]δερκ/2.\mathbb{P}\bigl{[}\,\lvert\varphi_{i}(v_{2})\rvert\leq\rho^{1/2}\lVert\varphi_{i}\rVert\,\bigr{]}\leq\delta^{-\varepsilon}\rho^{\kappa/2}.

Hence, recalling (2.8),

(2.10) [i=1d|φi(v2)|2<ρ]dδερκ/2.\mathbb{P}\Bigl{[}\,\sum_{i=1}^{d}\lvert\varphi_{i}(v_{2})\rvert^{2}<\rho\,\Bigr{]}\leq d\delta^{-\varepsilon}\rho^{\kappa/2}.

Inequalities (2.9) and (2.10) together imply (2.7) and finish the proof of the lemma. ∎

2.3. Fourier decay

To prove Theorem 2.1 we apply the L2L^{2}-flattening Proposition 2.5 repeatedly. The measures we obtain are images of tensor powers ηk\eta^{k} under polynomial maps EkEE^{k}\to E, and we need to compare their Fourier decay to that of simple multiplicative convolutions of η\eta. This is the content of the next lemma. This technique will also be useful to weaken slightly the assumptions of Theorem 2.1, see Corollary 2.11 below.

Lemma 2.9.

Let EE be any real associative algebra, and let η\eta be a measure on EE with η(E)1\eta(E)\leq 1. Let μ=ηηηηηη\mu=\eta*\eta*\eta\boxminus\eta*\eta*\eta then for any integer m1m\geq 1,

ξE,|η3m^(ξ)|2mμm^(ξ).\forall\xi\in E^{*},\quad\lvert\widehat{\eta^{*3m}}(\xi)\rvert^{2^{m}}\leq\widehat{\mu^{*m}}(\xi).
Proof.

By [28, Lemma B.6], if η\eta, η\eta^{\prime}, η′′\eta^{\prime\prime} are probability measures on EE, then the Fourier transform of η(ηη)η′′\eta*(\eta^{\prime}\boxminus\eta^{\prime})*\eta^{\prime\prime} takes non-negative real values and moreover,

ξE,|(ηηη′′)(ξ)|2(η(ηη)η′′)(ξ).\forall\xi\in E^{*},\quad\lvert(\eta*\eta^{\prime}*\eta^{\prime\prime})^{\wedge}(\xi)\rvert^{2}\leq\bigl{(}\eta*(\eta^{\prime}\boxminus\eta^{\prime})*\eta^{\prime\prime}\bigr{)}^{\wedge}(\xi).

By a simple scaling argument we see that the same holds when η\eta, η\eta^{\prime}, η′′\eta^{\prime\prime} are finite measures with total mass η(E),η(E),η′′(E)1\eta(E),\eta^{\prime}(E),\eta^{\prime\prime}(E)\leq 1. Using this inequality mm times with measure η=η3\eta^{\prime}=\eta^{*3}, so that μ=ηη\mu=\eta^{\prime}\boxminus\eta^{\prime}, we get

μm^(ξ)\displaystyle\widehat{\mu^{*m}}(\xi) |μ(m1)η3^(ξ)|2|η3m^(ξ)|2m.\displaystyle\geq\lvert\widehat{\mu^{*(m-1)}*\eta^{*3}}(\xi)\rvert^{2}\geq\dots\geq\lvert\widehat{\eta^{*3m}}(\xi)\rvert^{2^{m}}.

We shall also need a lemma on Fourier decay for multiplicative convolutions of measures with small L2L^{2}-norm. In the case where E=E=\mathbb{R}, such bounds originate in the work of Falconer [20] on projection theorems, and appear explicitly in Bourgain [10, Theorem 7]. The result below is taken from [25, Lemma 2.9].

Lemma 2.10.

Let EE be a finite-dimensional real associative algebra with unit. The following holds for any parameters κ>0\kappa>0 and ε>0\varepsilon>0 and any scale δ>0\delta>0 small enough. Let η\eta and ν\nu be probability measures on EE. Assume

  1. (1)

    η2,δ2δκ\lVert\eta\rVert_{2,\delta}^{2}\leq\delta^{-\kappa},

  2. (2)

    suppηB(0,δε)\operatorname{supp}\eta\subset\operatorname{B}(0,\delta^{-\varepsilon}) and suppνB(0,δε)\operatorname{supp}\nu\subset\operatorname{B}(0,\delta^{-\varepsilon}),

  3. (3)

    for every proper affine subspace WEW\subset E, ν(W(δ))δ2κ\nu(W^{(\delta)})\leq\delta^{2\kappa}.

Then for ξE\xi\in E^{*} with δ1+εξδ1ε\delta^{-1+\varepsilon}\leq\lVert\xi\rVert\leq\delta^{-1-\varepsilon},

|ην^(ξ)|δκdimE+3O(ε).\lvert\widehat{\eta*\nu}(\xi)\rvert\leq\delta^{\frac{\kappa}{\dim E+3}-O(\varepsilon)}.

We can finally derive Theorem 2.1.

Proof of Theorem 2.1.

First case: η1==ηs=η\eta_{1}=\dots=\eta_{s}=\eta.
For a measure η\eta satisfying NC(ε,κ,τ)\operatorname{NC}(\varepsilon,\kappa,\tau) at scale δ\delta, we write ess(η)\operatorname{ess}(\eta) to denote the essential part of η\eta, defined as a measure on EE satisfying

  1. (1)

    ess(η)η\operatorname{ess}(\eta)\leq\eta and ess(η)(E)η(E)3δτ\operatorname{ess}(\eta)(E)\geq\eta(E)-3\delta^{\tau},

  2. (2)

    ess(η)\operatorname{ess}(\eta) is supported on B(0,δε)SE(δε)\operatorname{B}(0,\delta^{-\varepsilon})\setminus S_{E}(\delta^{\varepsilon}),

  3. (3)

    ess(η)\operatorname{ess}(\eta) satisfies ANC(ε,κ)\operatorname{ANC}(\varepsilon,\kappa) at scale δ\delta.

The second and third conditions from NC0\operatorname{NC}_{0} are invariant under translation, so that if μ\mu is a measure satisfying NC0(ε,κ,τ)\operatorname{NC}_{0}(\varepsilon,\kappa,\tau) and ν\nu any measure supported on BE(0,δε)\operatorname{B}_{E}(0,\delta^{-\varepsilon}), then μν\mu\boxplus\nu always satisfies NC0(2ε,κ,τ)\operatorname{NC}_{0}(2\varepsilon,\kappa,\tau). Therefore, if η\eta and η\eta^{\prime} satisfy NC(ε,κ,τ)\operatorname{NC}(\varepsilon,\kappa,\tau) at scale δ\delta, then ηη\eta\boxplus\eta^{\prime} satisfy NC(O(ε),κ,τ2)\operatorname{NC}(O(\varepsilon),\kappa,\frac{\tau}{2}) at scale δ\delta, with essential part ess(ηη)=ess(η)ess(η)\operatorname{ess}(\eta\boxplus\eta^{\prime})=\operatorname{ess}(\eta)\boxplus\operatorname{ess}(\eta^{\prime}). Similarly, ηη\eta\boxminus\eta^{\prime} and ηη\eta*\eta^{\prime} satisfy NC(O(ε),κ,τ2)\operatorname{NC}(O(\varepsilon),\kappa,\frac{\tau}{2}). We may therefore define inductively η0=ess(η)\eta_{0}=\operatorname{ess}(\eta), and for k0k\geq 0,

ηk+1=ess(ηk3ηk3),\eta_{k+1}=\operatorname{ess}\bigl{(}\eta_{k}^{*3}\boxminus\eta_{k}^{*3}\bigr{)},

to get, for each k0k\geq 0,

  1. (1)

    ηk(E)1δτOk(1)\eta_{k}(E)\geq 1-\delta^{\frac{\tau}{O_{k}(1)}},

  2. (2)

    ηk\eta_{k} is supported on B(0,δOk(ε))SE(δOk(ε))\operatorname{B}(0,\delta^{-O_{k}(\varepsilon)})\setminus S_{E}(\delta^{O_{k}(\varepsilon)}),

  3. (3)

    ηk\eta_{k} satisfies ANC(Ok(ε),κ)\operatorname{ANC}(O_{k}(\varepsilon),\kappa) at scale δ\delta.

Note that ANC(Ok(ε),κ)\operatorname{ANC}(O_{k}(\varepsilon),\kappa) implies

ηk2,δ2δdimE+κOk(ε).\lVert\eta_{k}\rVert_{2,\delta}^{2}\leq\delta^{-\dim E+\kappa-O_{k}(\varepsilon)}.

Set κ=κ3\kappa^{\prime}=\frac{\kappa}{3}. By Proposition 2.5 applied with κ\kappa^{\prime} instead of κ\kappa, there exists ε1=ε1(E,κ)\varepsilon_{1}=\varepsilon_{1}(E,\kappa^{\prime}) such that, provided ε>0\varepsilon>0 is small enough, we have for each 0kdimEε10\leq k\leq\left\lceil\frac{\dim E}{\varepsilon_{1}}\right\rceil, either ηk2,δ2δκ\lVert\eta_{k}\rVert^{2}_{2,\delta}\leq\delta^{-\kappa^{\prime}} or

ηk+12,δ2δε1ηk2,δ2.\lVert\eta_{k+1}\rVert_{2,\delta}^{2}\leq\delta^{\varepsilon_{1}}\lVert\eta_{k}\rVert_{2,\delta}^{2}.

Hence there exists sdimEε1s\leq\left\lceil\frac{\dim E}{\varepsilon_{1}}\right\rceil such that

ηs2,δ2δκ.\lVert\eta_{s}\rVert^{2}_{2,\delta}\leq\delta^{-\kappa^{\prime}}.

By Lemma 2.10 applied with κ\kappa^{\prime} instead of κ\kappa, for ξE\xi\in E^{*} with δ1+εξδ1ε\delta^{-1+\varepsilon}\leq\lVert\xi\rVert\leq\delta^{-1-\varepsilon},

|ηs2^(ξ)|δκO(1)O(ε)δκO(1)δτ,\bigl{\lvert}\widehat{\eta_{s}^{*2}}(\xi)\bigr{\rvert}\leq\delta^{\frac{\kappa^{\prime}}{O(1)}-O(\varepsilon)}\leq\delta^{\frac{\kappa^{\prime}}{O(1)}}\leq\delta^{\tau},

provided ε\varepsilon is chosen sufficiently small. Now, a first application of Lemma 2.9 to μ=ηs13ηs13\mu=\eta_{s-1}^{*3}\boxminus\eta_{s-1}^{*3} with m=2m=2 yields

ηs2^(ξ)+δτOs(1)μ2^(ξ)|ηs123^(ξ)|22.\widehat{\eta_{s}^{*2}}(\xi)+\delta^{\frac{\tau}{O_{s}(1)}}\geq\widehat{\mu^{*2}}(\xi)\geq\lvert\widehat{\eta_{s-1}^{*2\cdot 3}}(\xi)\rvert^{2^{2}}.

A second application of the same lemma to μ1=ηs23ηs23\mu_{1}=\eta_{s-2}^{*3}\boxminus\eta_{s-2}^{*3} with m=23m=2\cdot 3 gives

ηs123^(ξ)+δτOs(1)μ123^(ξ)|ηs2232^(ξ)|223\widehat{\eta_{s-1}^{*2\cdot 3}}(\xi)+\delta^{\frac{\tau}{O_{s}(1)}}\geq\widehat{\mu_{1}^{*2\cdot 3}}(\xi)\geq\lvert\widehat{\eta_{s-2}^{*2\cdot 3^{2}}}(\xi)\rvert^{2^{2\cdot 3}}

and repeating this process ss times, we finally obtain

|η023s^(ξ)|Os(1)|ηs2^(ξ)|+δτOs(1)δτ+δτOs(1)δτOs(1).\bigl{\lvert}\widehat{\eta_{0}^{*2\cdot 3^{s}}}(\xi)\bigr{\rvert}^{O_{s}(1)}\leq\bigl{\lvert}\widehat{\eta_{s}^{*2}}(\xi)\bigr{\rvert}+\delta^{\frac{\tau}{O_{s}(1)}}\leq\delta^{\tau}+\delta^{\frac{\tau}{O_{s}(1)}}\leq\delta^{\frac{\tau}{O_{s}(1)}}.

This allows to conclude:

|η23s^(ξ)||η023s^(ξ)|+Os(δτ)δετ,\bigl{\lvert}\widehat{\eta^{*2\cdot 3^{s}}}(\xi)\bigr{\rvert}\leq\bigl{\lvert}\widehat{\eta_{0}^{*2\cdot 3^{s}}}(\xi)\bigr{\rvert}+O_{s}(\delta^{\tau})\leq\delta^{\varepsilon\tau},

provided ε\varepsilon is sufficiently small. This proves the theorem, with parameter s(E,κ)=23ss(E,\kappa)=2\cdot 3^{s}.

General case
To deduce the general case from the previous one, we follow [28, Proof of Theorem B.3]. In short, one applies the previous case to the measures

ηλ=λ1(η1η1)++λs(ηsηs),\eta_{\lambda}=\lambda_{1}(\eta_{1}\boxminus\eta_{1})+\dots+\lambda_{s}(\eta_{s}\boxminus\eta_{s}),

where λ=(λ1,,λs)\lambda=(\lambda_{1},\dots,\lambda_{s}) in +s\mathbb{R}_{+}^{s} is such that λ1++λs1\lambda_{1}+\dots+\lambda_{s}\leq 1. The Fourier decay for η1ηs\eta_{1}*\dots*\eta_{s} can be deduced from that of ηληλ\eta_{\lambda}*\dots*\eta_{\lambda} for every λ\lambda using the fact that Fourier coefficients of ηληλ\eta_{\lambda}*\dots*\eta_{\lambda} can be written as polynomials in λ\lambda whose coefficients are essentially Fourier coefficients of η1ηs\eta_{1}*\dots*\eta_{s}. The reader is referred to [28] for details. ∎

We conclude this section by showing that the conclusion of Theorem 2.1 still holds if the non-concentration assumption is only satisfied for some additive convolution of the measures ηi\eta_{i}, i=1,,si=1,\dots,s. This will be useful when we study Fourier decay of random walks on linear groups.

Corollary 2.11.

Let EE be a normed finite-dimensional semisimple algebra over \mathbb{R}. Given DD\in\mathbb{N}^{*} and κ>0\kappa>0, there exists s=s(E,κ)s=s(E,\kappa)\in\mathbb{N} and ε=ε(E,κ,D)>0\varepsilon=\varepsilon(E,\kappa,D)>0 such that for any parameter τ(0,εκ)\tau\in{(0,\varepsilon\kappa)} the following holds for any scale δ>0\delta>0 sufficiently small.

If ηi\eta_{i}, i=1,,si=1,\dots,s are probability measures on EE such that each ηiD\eta_{i}^{\boxplus D} satisfies NC(ε,κ,τ)\operatorname{NC}(\varepsilon,\kappa,\tau) at scale δ\delta, then for all ξE\xi\in E^{*} with δ1+εξδ1ε\delta^{-1+\varepsilon}\leq\lVert\xi\rVert\leq\delta^{-1-\varepsilon},

|(η1ηs)(ξ)|δετ.\lvert(\eta_{1}*\dotsm*\eta_{s})^{\wedge}(\xi)\rvert\leq\delta^{\varepsilon\tau}.
Proof.

Let ξE\xi\in E^{*} with δ1+εξδ1ε\delta^{-1+\varepsilon}\leq\lVert\xi\rVert\leq\delta^{-1-\varepsilon}. Since all the measures ηiDηiD\eta_{i}^{\boxplus D}\boxminus\eta_{i}^{\boxplus D}, i=1,,si=1,\dots,s satisfy NC(ε,κ,τ)\operatorname{NC}(\varepsilon,\kappa,\tau) at scale δ\delta, Theorem 2.1 shows that

|((η1Dη1D)(ηsDηsD))(ξ)|δετ.\left\lvert\big{(}(\eta_{1}^{\boxplus D}\boxminus\eta_{1}^{\boxplus D})*\dotsm*(\eta_{s}^{\boxplus D}\boxminus\eta_{s}^{\boxplus D})\big{)}^{\wedge}(\xi)\right\rvert\leq\delta^{\varepsilon\tau}.

Applying [28, Lemma B.6] repeatedly ss times, we see that

|((η1Dη1D)(ηsDηsD))(ξ)|\displaystyle\left\lvert\big{(}(\eta_{1}^{\boxplus D}\boxminus\eta_{1}^{\boxplus D})*\dotsm*(\eta_{s}^{\boxplus D}\boxminus\eta_{s}^{\boxplus D})\big{)}^{\wedge}(\xi)\right\rvert
|((η1Dη1D)(ηs1Dηs1D)ηs)(ξ)|2D\displaystyle\qquad\qquad\geq\left\lvert\big{(}(\eta_{1}^{\boxplus D}\boxminus\eta_{1}^{\boxplus D})*\dotsm*(\eta_{s-1}^{\boxplus D}\boxminus\eta_{s-1}^{\boxplus D})*\eta_{s}\big{)}^{\wedge}(\xi)\right\rvert^{2D}
\displaystyle\qquad\qquad\geq\dots
|(η1ηs)(ξ)|(2D)s\displaystyle\qquad\qquad\geq\lvert(\eta_{1}*\dotsm*\eta_{s})^{\wedge}(\xi)\rvert^{(2D)^{s}}

so that

|(η1ηs)(ξ)|δετ(2D)s.\lvert(\eta_{1}*\dotsm*\eta_{s})^{\wedge}(\xi)\rvert\leq\delta^{\frac{\varepsilon\tau}{(2D)^{s}}}.

3. Non-concentration for random walks on semisimple groups

In this section we consider a probability measure μ\mu on SLd()\operatorname{SL}_{d}(\mathbb{Z}), and we prove some non-concentration property for the law of the associated random walk, viewed as a measure on the algebra generated by μ\mu. Let Γ\Gamma be the group generated by the support of μ\mu and GG be the Zariski closure of Γ\Gamma in SLd()\operatorname{SL}_{d}(\mathbb{R}). We assume that GG is semisimple and Zariski connected.


Let EE denote the \mathbb{R}-linear span of GG in d()\operatorname{\mathcal{M}}_{d}(\mathbb{R}), which is also the subalgebra generated by GG in d()\operatorname{\mathcal{M}}_{d}(\mathbb{R}). Since GG is semisimple, one may decompose EE into a direct sum of irreducible GG-modules. This gives a decomposition of EE into minimal left ideals, so that by the fundamental theorem of semisimple rings [39, §117], EE is a semisimple algebra. Let

(3.1) E=E1ErE=E_{1}\oplus\dots\oplus E_{r}

be the decomposition of EE into simple factors, i.e. into minimal two-sided ideals. For j=1,,rj=1,\dotsc,r, let πj:EEj\pi_{j}\colon E\to E_{j} denote the corresponding projections. Consider the top Lyapunov exponent associated to μ\mu on each of the factors EjE_{j}, defined by

λ1(μ,Ej)=limn+1nlogπj(g)dμn(g).\lambda_{1}(\mu,E_{j})=\lim_{n\to+\infty}\frac{1}{n}\int\log\lVert\pi_{j}(g)\rVert\,\mathrm{d}\mu^{*n}(g).

In order to study the law at time nn of the random walk, we shall use the rescaling automorphism φn:EE\varphi_{n}\colon E\to E defined by

(3.2) φn(g)=j=1renλ1(μ,Ej)πj(g).\varphi_{n}(g)=\sum_{j=1}^{r}e^{-n\lambda_{1}(\mu,E_{j})}\pi_{j}(g).

Recall that by Furstenberg’s theorem [21] on the positivity of the Lyapunov exponent, one has λ1(μ,Ej)0\lambda_{1}(\mu,E_{j})\geq 0 with equality if and only if πj(G)\pi_{j}(G) is compact. After reordering the factors, we may assume that λ1(μ,Ej)>0\lambda_{1}(\mu,E_{j})>0 if and only if jsj\leq s for some integer srs\leq r. Let E=E1EsE^{\prime}=E_{1}\oplus\dots\oplus E_{s} and π:EE\pi^{\prime}\colon E\to E^{\prime} the corresponding projection. Finally, for n1n\geq 1 we define

μn=(πφn)(μn).\mu_{n}=(\pi^{\prime}\circ\varphi_{n})_{*}(\mu^{*n}).

The goal of this section is as follows.

Proposition 3.1 (Non-concentration).

Let μ\mu be a probability measure on SLd()\operatorname{SL}_{d}(\mathbb{Z}) having a finite exponential moment. Let GG denote the algebraic group generated by μ\mu. Assume that GG is semisimple and Zariski connected, and denote by Ed()E\subset\operatorname{\mathcal{M}}_{d}(\mathbb{R}) the algebra generated by GG. Writing D=dimED=\dim E, there exists κ=κ(μ)>0\kappa=\kappa(\mu)>0 such that for any ε>0\varepsilon>0 there exists τ>0\tau>0 such that μnD\mu_{n}^{\boxplus D} satisfies NC(ε,κ,τ)\operatorname{NC}(\varepsilon,\kappa,\tau) at scale ene^{-n} in EE^{\prime} for all nn sufficiently large.

The readers can easily convince themselves that μn\mu_{n} does not satisfy NC(ε,κ,τ)\operatorname{NC}(\varepsilon,\kappa,\tau), especially the non-concentration condition near singular matrices. Hence taking an additive convolution power is necessary.

3.1. Non-concentration near affine subspaces

In this subsection we show that if μ\mu is a probability measure on SLd()\operatorname{SL}_{d}(\mathbb{Z}) generating a connected semisimple algebraic group GG, the law at time nn of the random walk associated to μ\mu is not concentrated near proper affine subspaces of the algebra generated by μ\mu.

We introduce a quasi-norm adapted to the random walk on the algebra EE generated by μ\mu. Given an element gg in EE, we write g=i=1rgig=\sum_{i=1}^{r}g_{i} according to the direct sum decomposition (3.1) and set

|g|\scaleobj0.7=max1isgi1λ1(μ,Ei).\lvert g\rvert^{\scaleobj{0.7}{\sim}}=\max_{1\leq i\leq s}\lVert g_{i}\rVert^{\frac{1}{\lambda_{1}(\mu,E_{i})}}.

Note that |g|=0\lvert g\rvert=0 if and only if gg lies in the sum

E0:=Es+1ErE_{0}:=E_{s+1}\oplus\dots\oplus E_{r}

of all compact factors. We denote by d~\tilde{d} the quasi-distance on EE given by d~(x,y)=|xy|\scaleobj0.7\tilde{d}(x,y)=\lvert x-y\rvert^{\scaleobj{0.7}{\sim}}. For instance, if WW is any affine subspace of EE, we write

d~(g,W)=infwW|gw|\scaleobj0.7.\tilde{d}(g,W)=\inf_{w\in W}\lvert g-w\rvert^{\scaleobj{0.7}{\sim}}.

Our goal is the following proposition.

Proposition 3.2 (Affine non-concentration on EE).

Let μ\mu be a probability measure on SLd()\operatorname{SL}_{d}(\mathbb{Z}) having a finite exponential moment. Let GG denote the algebraic group generated by μ\mu. Assume that GG is semisimple and Zariski connected, and denote by Ed()E\subset\operatorname{\mathcal{M}}_{d}(\mathbb{R}) the algebra generated by GG.

There exists κ=κ(μ)>0\kappa=\kappa(\mu)>0 such that for every n0n\geq 0 and ρen\rho\geq e^{-n}, for every affine hyperplane WEW\subset E such that WWE0W-W\supset E_{0},

μn({gG|d~(g,W)<ρminjJW|πj(g)|\scaleobj0.7})ρκ\mu^{*n}\bigl{(}\bigl{\{}\,g\in G\,\big{|}\,\tilde{d}(g,W)<\rho\min_{j\in J_{W}}\lvert\pi_{j}(g)\rvert^{\scaleobj{0.7}{\sim}}\,\bigr{\}}\bigr{)}\ll\rho^{\kappa}

where JW={ 1jr|VjWW}J_{W}=\{\,1\leq j\leq r\,|\,V_{j}\not\subset W-W\,\}.

Remark.

In general, it is not possible to replace the minimum minjJW|πj(g)|\scaleobj0.7\min_{j\in J_{W}}\lvert\pi_{j}(g)\rvert^{\scaleobj{0.7}{\sim}} by |g|\scaleobj0.7\lvert g\rvert^{\scaleobj{0.7}{\sim}}. This can be seen for example by taking G=G1×G1G=G_{1}\times G_{1} and μ=μ1μ1\mu=\mu_{1}\otimes\mu_{1}; in other words, the random walk is the direct product of two independent copies of a random walk on G1G_{1}. By the central limit theorem for random matrix products [8, Theorem 5.1, page 121], the probability to obtain at time nn an element g=(g1,g2)g=(g_{1},g_{2}) such that g1eng2\lVert g_{1}\rVert\leq e^{-\sqrt{n}}\lVert g_{2}\rVert has a positive limit cc. Therefore, for large nn,

μn({g=(g1,g2)G|g1<eng})c2.\mu^{*n}\bigl{(}\bigl{\{}\,g=(g_{1},g_{2})\in G\,\big{|}\,\lVert g_{1}\rVert<e^{-\sqrt{n}}\lVert g\rVert\,\bigr{\}}\bigr{)}\geq\frac{c}{2}.

and taking W={0}×Span(G1)W=\{0\}\times\operatorname{Span}_{\mathbb{R}}(G_{1}), we find

μn({gG|d~(g,W)<en|g|\scaleobj0.7})c2.\mu^{*n}\bigl{(}\bigl{\{}\,g\in G\,\big{|}\,\tilde{d}(g,W)<e^{-\sqrt{n}}\lvert g\rvert^{\scaleobj{0.7}{\sim}}\,\bigr{\}})\geq\frac{c}{2}.

3.1.1. The case of a simple algebra

For clarity, we first explain the proof of Proposition 3.2 when the algebra EE generated by GG is simple. In that case, the quasi-norm is a norm on EE and minjJW|πj(g)|\scaleobj0.7=g1λ1\min_{j\in J_{W}}\lvert\pi_{j}(g)\rvert^{\scaleobj{0.7}{\sim}}=\lVert g\rVert^{\frac{1}{\lambda_{1}}}. The key result in the proof is the following proposition, which we shall later apply to the irreducible action of G×GG\times G on EE.

Proposition 3.3.

Let μ\mu be a probability measure on SLd()\operatorname{SL}_{d}(\mathbb{Z}) with a finite exponential moment. Assume that the algebraic group GG generated by μ\mu is Zariski connected and acts irreducibly on V=dV=\mathbb{R}^{d}. There exists κ=κ(μ)\kappa=\kappa(\mu) such that for every vVv\in V, and any affine hyperplane WVW\subset V,

μn({gG|d(gv,W)ρgv})ρκ.\mu^{*n}\bigl{(}\{\,g\in G\,|\,d(gv,W)\leq\rho\lVert gv\rVert\,\}\bigr{)}\ll\rho^{\kappa}.
Proof.

First step: escape from affine subvarieties.
We claim that there exists c>0c>0 such that for every affine map ff on EE that is not identically zero on GG,

μn({gG|f(g)=0})ecn.\mu^{*n}\bigl{(}\{\,g\in G\,|\,f(g)=0\,\}\bigr{)}\ll e^{-cn}.

Indeed, by [3, Lemme 8.5], the group GG is semisimple. So the desired inequality is a particular case of [25, Proposition 3.7], whose proof is a combination of the spectral gap property modulo prime integers [35] and the Lang-Weil estimates on the number of points on algebraic varieties in finite fields.

Second step: a small neighborhood via a Diophantine property.
Let us show that there exist C,c>0C,c>0 such that for every non-zero polynomial map ff of degree at most 11 on GG,

μn({gG||f(g)|eCnf})ecn.\mu^{*n}\bigl{(}\{\,g\in G\,|\,\lvert f(g)\rvert\leq e^{-Cn}\lVert f\rVert\,\}\bigr{)}\ll e^{-cn}.

In the above, a polynomial map ff on GG is simply the restriction to GG of a polynomial map on EE; it is said to have degree at most DD if it is the restriction of a polynomial map on EE of degree at most DD. We endow the finite-dimensional space of polynomial maps of degree at most 11 with a fixed norm \lVert\cdot\rVert, say f=supgGBE(1,1)|f(g)|\lVert f\rVert=\sup_{g\in G\cap B_{E}(1,1)}\lvert f(g)\rvert. By the large deviation principle (see Theorem 3.11 below), there exists c>0c>0 such that for all nn large enough, μn({gG|g>e2nλ1(μ,V)})ecn\mu^{*n}(\{g\in G\ |\ \lVert g\rVert>e^{2n\lambda_{1}(\mu,V)}\})\leq e^{-cn}. Therefore, to prove the desired inequality, it suffices to show that for C0C\geq 0 large enough, the subset And()A_{n}\subset\operatorname{\mathcal{M}}_{d}(\mathbb{Z}) defined as

An={gΓ||f(g)|eCnf and ge2nλ1(μ,V)}A_{n}=\bigl{\{}\,g\in\Gamma\,\big{|}\,\lvert f(g)\rvert\leq e^{-Cn}\lVert f\rVert\text{ and }\lVert g\rVert\leq e^{2n\lambda_{1}(\mu,V)}\,\bigr{\}}

is included in GkerψG\cap\ker\psi for some affine map ψ:E\psi:E\to\mathbb{R} not identically zero on GG.

Suppose for a contradiction that this is not the case. Letting k=dim1[G]k=\dim\mathbb{R}_{\leq 1}[G] be the dimension of the space of polynomial maps on GG of degree at most 11, we may choose g1,,gkg_{1},\dots,g_{k} in AnA_{n} such that the linear map

L:1[G]kψ(ψ(g1),,ψ(gk))\begin{array}[]{lccc}L\colon&\mathbb{R}_{\leq 1}[G]&\to&\mathbb{R}^{k}\\ &\psi&\mapsto&(\psi(g_{1}),\dots,\psi(g_{k}))\end{array}

is bijective. Since it has integer coefficients and norm at most eC0ne^{C_{0}n}, we get L1eC1n\lVert L^{-1}\rVert\leq e^{C_{1}n}. In particular, feC1nLf=eC1nmax1ik|f(gi)|eC1neCnf\lVert f\rVert\leq e^{C_{1}n}\lVert Lf\rVert=e^{C_{1}n}\max_{1\leq i\leq k}\lvert f(g_{i})\rvert\leq e^{C_{1}n}e^{-Cn}\lVert f\rVert, which is the desired contradiction if C>C1C>C_{1}.

Third step: distance to proper subspaces.
We claim that there exist C,c>0C,c>0 such that for every vVv\in V and every affine hyperplane WVW\subset V,

μn({gG|d(gv,W)eCnv})ecn.\mu^{*n}(\{g\in G\ |\ d(gv,W)\leq e^{-Cn}\lVert v\rVert\})\ll e^{-cn}.

Indeed, let φW:V\varphi_{W}:V\to\mathbb{R} be an affine map such that kerφW=W\ker\varphi_{W}=W, and consider the affine map on GG given by

fv,W(g)=φW(gv)φW.f_{v,W}(g)=\frac{\varphi_{W}(gv)}{\lVert\varphi_{W}\rVert}.

Note that |fv,W(g)|d(gv,W)\lvert f_{v,W}(g)\rvert\asymp d(gv,W). Let B=BG(1,1)B=\operatorname{B}_{G}(1,1) denote the unit ball centered at the identity in GG. Note that vfv,W=supgB|fv,W(g)|\lVert v\rVert\asymp\lVert f_{v,W}\rVert=\sup_{g\in B}\lvert f_{v,W}(g)\rvert within constants independent of vv and WW. Indeed, otherwise, we may find vnv_{n} and WnW_{n} such that supgBd(gvn,Wn)0\sup_{g\in B}d(gv_{n},W_{n})\to 0. Extracting subsequences if necessary, we may assume that vnvv_{n}\to v and WnWW_{n}\to W; then for every gBg\in B, d(gv,W)=limd(gvn,Wn)=0d(gv,W)=\lim d(gv_{n},W_{n})=0. This implies that GvWG\cdot v\subset W and contradicts the assumption that GG acts irreducibly on VV. The desired inequality therefore follows from the previous step.

Fourth step: scaling.
First observe that increasing CC slightly, we can assume that for every vVv\in V and every affine hyperplane WVW\subset V,

μn({gG|d(gv,W)eCngv})ecn.\mu^{*n}\bigl{(}\{\,g\in G\,|\,d(gv,W)\leq e^{-Cn}\lVert gv\rVert\,\}\bigr{)}\ll e^{-cn}.

Indeed, by the large deviation estimate,

μn({gG|gve2λ1nv})1ecn\mu^{*n}\bigl{(}\{\,g\in G\,|\,\lVert gv\rVert\leq e^{2\lambda_{1}n}\lVert v\rVert\,\}\bigr{)}\geq 1-e^{-cn}

where λ1=λ1(μ,V)\lambda_{1}=\lambda_{1}(\mu,V) is the top Lyapunov exponent of μ\mu. To conclude, let κ=cC\kappa=\frac{c}{C}, where C,c>0C,c>0 are the constants obtained above. Choose mm\in\mathbb{N}^{*} such that ρ=eCm\rho=e^{-Cm} and write

μn({gG|d(gv,W)ρgv})\displaystyle\mu^{*n}\bigl{(}\{\,g\in G\,|\,d(gv,W)\leq\rho\lVert gv\rVert\,\}\bigr{)}
=μm({gG|d(gg1v,W)eCmgg1v})dμ(nm)(g1)\displaystyle\qquad\qquad=\int\mu^{*m}\bigl{(}\{\,g\in G\,|\,d(gg_{1}v,W)\leq e^{-Cm}\lVert gg_{1}v\rVert\,\}\bigr{)}\,\mathrm{d}\mu^{*(n-m)}(g_{1})
ecm=ρκ.\displaystyle\qquad\qquad\leq e^{-cm}=\rho^{\kappa}.

Proof of Proposition 3.2, case where EE is simple.

For xEx\in E, let Lx:EEL_{x}:E\to E and Rx:EER_{x}:E\to E denote the left and right multiplication by xx, respectively. Given a probability measure μ\mu on GG, we define a probability measure μ¯\bar{\mu} on GL(E)\operatorname{GL}(E) by

μ¯=12Lμ+12Rμ.\bar{\mu}=\frac{1}{2}L_{*}\mu+\frac{1}{2}R_{*}\mu.

The group generated by μ¯\bar{\mu} is isomorphic to G×GG\times G and acts irreducibly on EE. Moreover, in an appropriate basis, the elements of suppμ¯\operatorname{supp}\bar{\mu} have integer coefficients, so we may apply Proposition 3.3 to μ¯\bar{\mu}, with vector v=1Ev=1_{E}, the unit of EE. Note that if g¯\bar{g} is a random element distributed according to μ¯n\bar{\mu}^{*n}, then g¯1E\bar{g}\cdot 1_{E} has law μn\mu^{*n}, and therefore we find, uniformly over all affine hyperplanes WEW\subset E,

μn({gG|d(g,W)ρg})ρκ,\mu^{*n}\bigl{(}\{\,g\in G\,|\,d(g,W)\leq\rho\lVert g\rVert\,\}\bigr{)}\ll\rho^{\kappa},

which is exactly the content of Proposition 3.2 in the case where EE is simple. ∎

3.1.2. General case

The proof of Proposition 3.2 in the general case follows the same strategy as in the simple case, but the argument becomes slightly more technical, because the norm on EE is replaced by a quasi-norm, and EE contains proper ideals.

To state the appropriate generalization of Proposition 3.3, we consider a probability measure μ\mu on SLd()\operatorname{SL}_{d}(\mathbb{Z}) with some finite exponential moment, and let GG be the algebraic group generated by μ\mu. We assume that GG is Zariski connected and that the space V=dV=\mathbb{R}^{d} can be decomposed into a sum of irreducible representations of GG:

V=V1Vr.V=V_{1}\oplus\dots\oplus V_{r}.

We denote by πj:VVj\pi_{j}\colon V\to V_{j}, j=1,,rj=1,\dots,r the corresponding projections. To define a quasi-norm on VV, we fix α=(α1,,αs)\alpha=(\alpha_{1},\dotsc,\alpha_{s}) an ss-tuple of positive real numbers, where ss is some fixed integer 1sr1\leq s\leq r, and set

|v|\scaleobj0.7=|v|α\scaleobj0.7=max1jsπj(v)αj.\lvert v\rvert^{\scaleobj{0.7}{\sim}}=\lvert v\rvert^{\scaleobj{0.7}{\sim}}_{\alpha}=\max_{1\leq j\leq s}\lVert\pi_{j}(v)\rVert^{\alpha_{j}}.

For example, |v|\scaleobj0.7=0\lvert v\rvert^{\scaleobj{0.7}{\sim}}=0 if and only if vV0:=Vs+1Vrv\in V_{0}:=V_{s+1}\oplus\dots\oplus V_{r}. The quasi-distance associated to ||\scaleobj0.7\lvert\boldsymbol{\,\cdot\,}\rvert^{\scaleobj{0.7}{\sim}} on VV is given by

d~(v,w)=d~α(v,w)=|vw|α\scaleobj0.7.\tilde{d}(v,w)=\tilde{d}_{\alpha}(v,w)=\lvert v-w\rvert^{\scaleobj{0.7}{\sim}}_{\alpha}.

It satisfies a weak form of the triangle inequality:

u,v,wV,d~(u,w)αd~(u,v)+d~(v,w).\forall u,v,w\in V,\quad\tilde{d}(u,w)\ll_{\alpha}\tilde{d}(u,v)+\tilde{d}(v,w).

Given a subset WVW\subset V, and vVv\in V, we define the distance from vv to WW by

d~(v,W)=infwWd~(v,w).\tilde{d}(v,W)=\inf_{w\in W}\tilde{d}(v,w).
Remark.

In all our applications, we shall take ss so that V0=Vs+1VrV_{0}=V_{s+1}\oplus\dots\oplus V_{r} is the sum of all compact factors and

αj=1λ1(μ,Vj)forj=1,,s\alpha_{j}=\frac{1}{\lambda_{1}(\mu,V_{j})}\quad\text{for}\ j=1,\dotsc,s

to obtain a quasi-norm adapted to the random walk associated to μ\mu, same as the one defined in the introduction. However, the proof works in the more general setting of any choice of ss and α\alpha.

In the remainder of this subsection, ss and α\alpha are fixed and the implied constants in all Landau and Vinogradov notations may depend on dd and α\alpha.

Proposition 3.4.

Assume that GG is Zariski connected and that the linear span of GG in End(V)\operatorname{End}(V) contains πj\pi_{j} for j=1,,sj=1,\dotsc,s. Then there exists κ=κ(μ,α)>0\kappa=\kappa(\mu,\alpha)>0 such that for any vVv\in V and any affine hyperplane WVW\subset V with V0WWV_{0}\subset W-W,

n0,ρen,μn({gG|d~(gv,W)<ρminjJW|πj(gv)|\scaleobj0.7})ρκ\forall n\geq 0,\;\forall\rho\geq e^{-n},\quad\mu^{*n}\bigl{(}\bigl{\{}\,g\in G\,\big{|}\,\tilde{d}(gv,W)<\rho\min_{j\in J_{W}}\lvert\pi_{j}(gv)\rvert^{\scaleobj{0.7}{\sim}}\,\bigr{\}}\bigr{)}\ll\rho^{\kappa}

where JW={ 1jr|VjWW}J_{W}=\{\,1\leq j\leq r\,|\,V_{j}\not\subset W-W\,\}.

Remark.

The requirement that the linear span of GG contain πj\pi_{j}, j=1,,sj=1,\dots,s is here to exclude examples such as V=V1V1V=V_{1}\oplus V_{1}, with GG acting irreducibly on V1V_{1}. Indeed, in that case the diagonal subspace W={(v1,v1);v1V1}W=\{(v_{1},v_{1})\ ;\ v_{1}\in V_{1}\} is stable under GG, so the proposition cannot hold.

In the proof, we will use Lemma 3.5 and Lemma 3.6, whose proofs will be given right after.

Proof.

First and second step: spectral gap and Diophantine property.
Arguing exactly as in the proof of Proposition 3.3 we obtain that there exist C,c>0C,c>0 such that for every polynomial map ff of degree at most 11 on GG,

μn({gG||f(g)|eCnf})ecn.\mu^{*n}\bigl{(}\{\,g\in G\,|\,\lvert f(g)\rvert\leq e^{-Cn}\lVert f\rVert\,\}\bigr{)}\ll e^{-cn}.

As before, the norm on polynomial maps of degree at most 11 is defined by f=supgBG(1,1)|f(g)|\lVert f\rVert=\sup_{g\in B_{G}(1,1)}\lvert f(g)\rvert.
Third step: distance to proper subspaces.
We claim that there exist C1,c>0C_{1},c>0 such that for every vVv\in V and every affine hyperplane WW such that WWV0W-W\supset V_{0},

μn({gG|d~(gv,W)eC1nminjJW|πj(v)|\scaleobj0.7})ecn.\mu^{*n}\bigl{(}\bigl{\{}\,g\in G\,\big{|}\,\tilde{d}(gv,W)\leq e^{-C_{1}n}\min_{j\in J_{W}}\lvert\pi_{j}(v)\rvert^{\scaleobj{0.7}{\sim}}\,\bigr{\}}\bigr{)}\ll e^{-cn}.

To prove this, Lemma 3.5 below shows that it is enough to show that if BB is some large ball in GG, then

μn({gG|d~(gv,W)eC1nsuphBd~(hv,W)})ecn.\mu^{*n}\bigl{(}\bigl{\{}\,g\in G\,\big{|}\,\tilde{d}(gv,W)\leq e^{-C_{1}n}\sup_{h\in B}\tilde{d}(hv,W)\,\bigr{\}}\bigr{)}\ll e^{-cn}.

Now let φW:V\varphi_{W}\colon V\to\mathbb{R} be an affine map such that kerφW=W\ker\varphi_{W}=W, and denote by W\ell_{W} the linear part of φW\varphi_{W}; by Lemma 3.6 below, the distance to WW for the quasi-norm is given by

vV,d~(v,W)mini:W(ui)0|φW(v)W(ui)|αj(i),\forall v\in V,\quad\tilde{d}(v,W)\asymp\min_{i:\ell_{W}(u_{i})\neq 0}\left\lvert\frac{\varphi_{W}(v)}{\ell_{W}(u_{i})}\right\rvert^{\alpha_{j(i)}},

where (ui)1id(u_{i})_{1\leq i\leq d} is an orthonormal basis compatible with the quasi-norm and uiVj(i)u_{i}\in V_{j(i)} for i=1,,di=1,\dots,d. Therefore, if gg satisfies d~(gv,W)eC1nsuphBd~(hv,W)\tilde{d}(gv,W)\leq e^{-C_{1}n}\sup_{h\in B}\tilde{d}(hv,W), there must exist ii such that

|φW(gv)W(ui)|αj(i)αeC1nsuphB|φW(hv)W(ui)|αj(i)\left\lvert\frac{\varphi_{W}(gv)}{\ell_{W}(u_{i})}\right\rvert^{\alpha_{j(i)}}\ll_{\alpha}e^{-C_{1}n}\sup_{h\in B}\left\lvert\frac{\varphi_{W}(hv)}{\ell_{W}(u_{i})}\right\rvert^{\alpha_{j(i)}}

whence

|φW(gv)|αeC1nαj(i)suphB|φW(hv)|.\lvert\varphi_{W}(gv)\rvert\ll_{\alpha}e^{-\frac{C_{1}n}{\alpha_{j(i)}}}\sup_{h\in B}\lvert\varphi_{W}(hv)\rvert.

If C1αj(i)>C\frac{C_{1}}{\alpha_{j(i)}}>C, the previous step applied to the affine map f:gφW(gv)f\colon g\mapsto\varphi_{W}(gv) shows that the μn\mu^{*n}-measure of such points is bounded above by ecne^{-cn}, so the desired statement is proved.

Fourth step: scaling
Note that there are C2=C2(μ,α)>1C_{2}=C_{2}(\mu,\alpha)>1 and c=c(μ)>0c=c(\mu)>0 such that for any vector vVv\in V and any affine hyperplane WVW\subset V with V0WWV_{0}\subset W-W,

(3.3) n0,μn({gG|d~(gv,W)eC2nminjJW|πj(gv)|\scaleobj0.7})ecn.\forall n\geq 0,\quad\mu^{*n}\bigl{(}\bigl{\{}\,g\in G\,\big{|}\,\tilde{d}(gv,W)\leq e^{-C_{2}n}\min_{j\in J_{W}}\lvert\pi_{j}(gv)\rvert^{\scaleobj{0.7}{\sim}}\,\bigr{\}}\bigr{)}\ll e^{-cn}.

This readily follows from the previous step, and from the fact that, by the exponential moment assumption, there are C3=C3(μ)>1C_{3}=C_{3}(\mu)>1 and c=c(μ)>0c=c(\mu)>0 such that

μn({gG|geC3n})ecn.\mu^{*n}\bigl{(}\{\,g\in G\,|\,\lVert g\rVert\geq e^{C_{3}n}\,\}\bigr{)}\ll e^{-cn}.

Noting that for any j=1,,sj=1,\dotsc,s, |πj(gv)|\scaleobj0.7gαj|πj(v)|\scaleobj0.7\lvert\pi_{j}(gv)\rvert^{\scaleobj{0.7}{\sim}}\leq\lVert g\rVert^{\alpha_{j}}\lvert\pi_{j}(v)\rvert^{\scaleobj{0.7}{\sim}}, we obtain (3.3) by taking C2=C1+(max1jsαj)C3C_{2}=C_{1}+(\max_{1\leq j\leq s}\alpha_{j})C_{3}.

Finally, given enρ1e^{-n}\leq\rho\leq 1, set m=logρC2m=\left\lfloor\frac{-\log\rho}{C_{2}}\right\rfloor. Writing μn=μmμ(nm)\mu^{*n}=\mu^{*m}*\mu^{*(n-m)} and using the fact that (3.3) holds uniformly in vv, we find

μn({gG|d~(gv,W)<ρminjJW|πj(gv)|\scaleobj0.7})\displaystyle\mu^{*n}\bigl{(}\bigl{\{}\,g\in G\,\big{|}\,\tilde{d}(gv,W)<\rho\min_{j\in J_{W}}\lvert\pi_{j}(gv)\rvert^{\scaleobj{0.7}{\sim}}\,\bigr{\}}\bigr{)}
Gμm({gG|d~(ghv,W)eC2mminjJW|πj(ghv)|\scaleobj0.7})dμ(nm)(h)\displaystyle\leq\int_{G}\mu^{*m}\bigl{(}\bigl{\{}\,g\in G\,\big{|}\,\tilde{d}(ghv,W)\leq e^{-C_{2}m}\min_{j\in J_{W}}\lvert\pi_{j}(ghv)\rvert^{\scaleobj{0.7}{\sim}}\,\bigr{\}}\bigr{)}\,\mathrm{d}\mu^{*(n-m)}(h)
ecm\displaystyle\ll e^{-cm}
ρc/C2.\displaystyle\ll\rho^{c/C_{2}}.

This finishes the proof of Proposition 3.4. ∎

We are left to show the two technical lemmas on quasi-norms and distances to hyperplanes that we used in the proof.

Lemma 3.5.

Assume that the linear span of GG in End(V)\operatorname{End}(V) contains πj\pi_{j}, for j=1,,sj=1,\dotsc,s. Then there exists a ball BGB\subset G such that for any affine hyperplane WVW\subset V with V0WWV_{0}\subset W-W, and any vVv\in V,

minjJW|πj(v)|\scaleobj0.7supgBd~(gv,W).\min_{j\in J_{W}}\lvert\pi_{j}(v)\rvert^{\scaleobj{0.7}{\sim}}\ll\sup_{g\in B}\tilde{d}(gv,W).
Proof.

In the particular case r=1r=1 (irreducible case), one may assume that the quasi-norm is equal to the Euclidean norm. So the desired inequality with B=BG(1,1)B=\operatorname{B}_{G}(1,1) has already been proved in the third step of the proof of Proposition 3.3.

For the general case, first observe that by working in the quotient space V/V0V/V_{0}, we may assume that V0={0}V_{0}=\{0\}, that is, V=V1VsV=V_{1}\oplus\dots\oplus V_{s}. Then assume for a contradiction that for arbitrarily large RR and arbitrarily small c>0c>0, there exist vVv\in V and WVW\subset V such that

(3.4) gBG(1,R),d~(gv,W)cminjJW|πj(v)|\scaleobj0.7.\forall g\in\operatorname{B}_{G}(1,R),\quad\tilde{d}(gv,W)\leq c\min_{j\in J_{W}}\lvert\pi_{j}(v)\rvert^{\scaleobj{0.7}{\sim}}.

Fix j{1,,s}j\in\{1,\dots,s\}. Let us construct a basis of VjV_{j} such that

vj,1vj,dimVjπj(v)dimVj\bigl{\lVert}v_{j,1}\wedge\dotsm\wedge v_{j,\dim V_{j}}\bigr{\rVert}\gg\lVert\pi_{j}(v)\rVert^{\dim V_{j}}

and

(3.5) i=1,,dimVj,vj,iBG(1,1)πj(v)πj(v).\forall i=1,\dots,\dim V_{j},\qquad v_{j,i}\in\operatorname{B}_{G}(1,1)\pi_{j}(v)-\pi_{j}(v).

For that, we proceed iteratively. Assuming vj,1,,vj,iv_{j,1},\dots,v_{j,i} have been constructed, we know from the irreducible case applied in VjV_{j} with vector v=πj(v)v=\pi_{j}(v) and subspace W=πj(v)+Span(vj,1,,vj,i)W=\pi_{j}(v)+\operatorname{Span}(v_{j,1},\dots,v_{j,i}), that there exists gg in BG(1,1)\operatorname{B}_{G}(1,1) such that d(gπj(v)πj(v),Span(vj,k;ki))πj(v)d(g\pi_{j}(v)-\pi_{j}(v),\operatorname{Span}(v_{j,k};k\leq i))\gg\lVert\pi_{j}(v)\rVert. So we set vj,i+1=gπj(v)πj(v)v_{j,i+1}=g\pi_{j}(v)-\pi_{j}(v). By construction, the vectors vj,iv_{j,i}, i1i\geq 1 are linearly independent, and in the end, we get a basis for VjV_{j} with the desired property. Concatenate these bases to get a basis (u1,,ud)(u_{1},\dotsc,u_{d}) of VV, which has the property that

(3.6) u1udj=1sπj(v)dimVj.\lVert u_{1}\wedge\dotsm\wedge u_{d}\rVert\gg\prod_{j=1}^{s}\lVert\pi_{j}(v)\rVert^{\dim V_{j}}.

Let W0=WWW_{0}=W-W denote the direction of WW. By assumption, for j=1,,sj=1,\dots,s, there exist constants βj,k\beta_{j,k}\in\mathbb{R} and elements gkg_{k} in GG such that

πj=kβj,kgk.\pi_{j}=\sum_{k}\beta_{j,k}g_{k}.

Set RR large enough so that for all kk, BG(1,1)gkBG(1,R)\operatorname{B}_{G}(1,1)g_{k}\subset\operatorname{B}_{G}(1,R). Fix i{1,,d}i\in\{1,\dots,d\}. From (3.5) we may write ui=gπj(v)πj(v)u_{i}=g\pi_{j}(v)-\pi_{j}(v) for some gBG(1,1)g\in\operatorname{B}_{G}(1,1), so

ui=kβj,k(ggkgk)vu_{i}=\sum_{k}\beta_{j,k}(gg_{k}-g_{k})v

and (3.4) allows us to bound

d~(ui,W0)2ck|βj,k|minjJW|πj(v)|\scaleobj0.7.\tilde{d}(u_{i},W_{0})\leq 2c\sum_{k}\lvert\beta_{j,k}\rvert\min_{j\in J_{W}}\lvert\pi_{j}(v)\rvert^{\scaleobj{0.7}{\sim}}.

For each i=1,,di=1,\dotsc,d, let wiW0w_{i}\in W_{0} be such that

d~(ui,wi)cminjJW|πj(v)|\scaleobj0.7,\tilde{d}(u_{i},w_{i})\ll c\min_{j\in J_{W}}\lvert\pi_{j}(v)\rvert^{\scaleobj{0.7}{\sim}},

where the involved constant depends on the numbers βj,k\beta_{j,k}. Using the assumption that jJWVjW0\bigoplus_{j\not\in J_{W}}V_{j}\subset W_{0}, after adjusting wiw_{i}, we can moreover ensure that

wiuijJWVj.w_{i}-u_{i}\in\bigoplus_{j\in J_{W}}V_{j}.

We can bound

(3.7) w1wdu1udcminj1αjj=1sπj(v)dimVj.\bigl{\lVert}w_{1}\wedge\dotsm\wedge w_{d}-u_{1}\wedge\dotsm\wedge u_{d}\bigr{\rVert}\ll c^{\min_{j}\frac{1}{\alpha_{j}}}\prod_{j=1}^{s}\lVert\pi_{j}(v)\rVert^{\dim V_{j}}.

Indeed, developing the first wedge product using wi=ui+(wiui)w_{i}=u_{i}+(w_{i}-u_{i}) and then decomposing each vector along V1VsV_{1}\oplus\dots\oplus V_{s} and further developing the sum, we can express w1wdu1udw_{1}\wedge\dotsm\wedge w_{d}-u_{1}\wedge\dotsm\wedge u_{d} as a sum of wedge products of dd vectors of the following types

  1. (1)

    (first type) πj(wiui)\pi_{j}(w_{i}-u_{i}) with jJWj\in J_{W}, or

  2. (2)

    (second type) πj(ui)\pi_{j}(u_{i}) with 1js1\leq j\leq s.

In each wedge product, the first type appears at least once and the product is zero unless πj\pi_{j} appears exactly dimVj\dim V_{j} times. We can bound vectors of the first type by

πj(wiui)d~(wi,ui)1αjc1αjπj(v)\lVert\pi_{j}(w_{i}-u_{i})\rVert\leq\tilde{d}(w_{i},u_{i})^{\frac{1}{\alpha_{j}}}\ll c^{\frac{1}{\alpha_{j}}}\lVert\pi_{j}(v)\rVert

and vectors of the second type by

πj(ui)πj(v).\lVert\pi_{j}(u_{i})\rVert\ll\lVert\pi_{j}(v)\rVert.

This proves (3.7).

To conclude, we choose cc be to small enough so that (3.7) combined with (3.6) implies w1wd0w_{1}\wedge\dotsm\wedge w_{d}\neq 0 contradicting the condition that W0W_{0} is a proper linear subspace of VV. ∎

The second lemma is an elementary computation using the definition of the quasi-norm. It is instructive to convince oneself with a picture that the lemma holds when the quasi-norm is simply the euclidean norm on d\mathbb{R}^{d}.

Lemma 3.6.

Let (ui)1id(u_{i})_{1\leq i\leq d} be a union of orthonormal bases of each of the VjV_{j}, j=1,,rj=1,\dots,r. For i=1,,di=1,\dotsc,d, denote by j(i)j(i) the unique integer such that uiVj(i)u_{i}\in V_{j(i)}.

Let vVv\in V and let WVW\subset V be an affine hyperplane with with V0WWV_{0}\subset W-W. Let φW:V\varphi_{W}\colon V\to\mathbb{R} be an affine map such that

W={vV|φW(v)=0}.W=\{\,v\in V\,|\,\varphi_{W}(v)=0\,\}.

Let W:V\ell_{W}\colon V\to\mathbb{R} denote the linear part of φW\varphi_{W}. We have for any vVv\in V,

d~(v,W)mini:W(ui)0|φW(v)W(ui)|αj(i).\tilde{d}(v,W)\asymp\min_{i:\ell_{W}(u_{i})\neq 0}\left\lvert\frac{\varphi_{W}(v)}{\ell_{W}(u_{i})}\right\rvert^{\alpha_{j(i)}}.
Proof.

Note that W(ui)0\ell_{W}(u_{i})\neq 0 implies that j(i)JWj(i)\in J_{W} and JW{1,,s}J_{W}\subset\{1,\dotsc,s\} because V0WWV_{0}\subset W-W. It follows that αj(i)\alpha_{j(i)} is defined and positive.

For any i{1,,d}i\in\{1,\dotsc,d\} with W(ui)0\ell_{W}(u_{i})\neq 0, we have vφW(v)W(ui)uiWv-\frac{\varphi_{W}(v)}{\ell_{W}(u_{i})}u_{i}\in W. Hence

d~(v,W)|φW(v)W(ui)|αj(i).\tilde{d}(v,W)\leq\left\lvert\frac{\varphi_{W}(v)}{\ell_{W}(u_{i})}\right\rvert^{\alpha_{j(i)}}.

Let uVu\in V be such that vuWv-u\in W. Write u=i=1dxiuiu=\sum_{i=1}^{d}x_{i}u_{i}. Then

φW(v)=φW(vu)+W(u)=i=1dxiW(ui).\varphi_{W}(v)=\varphi_{W}(v-u)+\ell_{W}(u)=\sum_{i=1}^{d}x_{i}\ell_{W}(u_{i}).

It follows that there exists ii with W(ui)0\ell_{W}(u_{i})\neq 0 and such that

|xi|1d|φW(v)W(ui)|.\lvert x_{i}\rvert\geq\frac{1}{d}\left\lvert\frac{\varphi_{W}(v)}{\ell_{W}(u_{i})}\right\rvert.

This allows to conclude since

d~(v,vu)=|u|\scaleobj0.7πj(i)(u)αj(i)|xi|αj(i).\tilde{d}(v,v-u)=\lvert u\rvert^{\scaleobj{0.7}{\sim}}\geq\lVert\pi_{j(i)}(u)\rVert^{\alpha_{j(i)}}\geq\lvert x_{i}\rvert^{\alpha_{j(i)}}.

To conclude, we explain how to obtain Proposition 3.2 from Proposition 3.4. The argument is essentially the same as the one used in the particular case where EE is simple.

Proof of Proposition 3.2, general case.

For xEx\in E, let Lx:EEL_{x}:E\to E and Rx:EER_{x}:E\to E denote the left and right multiplication by xx, respectively. Then, define

L:EEndExLxandR:EEndExRx\begin{array}[]{rccc}L\colon&E&\to&\operatorname{End}E\\ &x&\mapsto&L_{x}\end{array}\quad\mbox{and}\quad\begin{array}[]{rccc}R\colon&E&\to&\operatorname{End}E\\ &x&\mapsto&R_{x}\end{array}

Given a probability measure μ\mu on GG, we define a probability measure μ¯\bar{\mu} on GL(E)\operatorname{GL}(E) by

μ¯=12Lμ+12Rμ.\bar{\mu}=\frac{1}{2}L_{*}\mu+\frac{1}{2}R_{*}\mu.

The group G¯\bar{G} generated by μ¯\bar{\mu} is isomorphic to G×GG\times G and the decomposition of EE into irreducible G¯\bar{G}-submodules is simply the decomposition into simple ideals E=jEjE=\oplus_{j}E_{j}. By definition, the algebra generated by GG contains the unit 1Ej1_{E_{j}} of EjE_{j} for each jj. It follows that the linear span of G¯\bar{G} contains all projections πj:EEj\pi_{j}\colon E\to E_{j}. Moreover, in an appropriate basis, the elements of suppμ¯\operatorname{supp}\bar{\mu} have integer coefficients, so we may apply Proposition 3.4 to μ¯\bar{\mu}, with vector v=1Ev=1_{E}. Note that if g¯\bar{g} is a random element distributed according to μ¯n\bar{\mu}^{*n}, then g¯1E\bar{g}\cdot 1_{E} has law μn\mu^{*n}, and therefore we obtain κ>0\kappa>0 such that uniformly over all affine hyperplanes WEW\subset E with WWE0W-W\supset E_{0},

n0,ρen,μn({gG|d~(g,W)ρminjJW|πj(g)|\scaleobj0.7})ρκ.\forall n\geq 0,\;\forall\rho\geq e^{-n},\quad\mu^{*n}\bigl{(}\bigl{\{}\,g\in G\,\big{|}\,\tilde{d}(g,W)\leq\rho\min_{j\in J_{W}}\lvert\pi_{j}(g)\rvert^{\scaleobj{0.7}{\sim}}\,\bigr{\}})\ll\rho^{\kappa}.

3.2. Non-concentration at singular matrices

As in the previous paragraph, μ\mu denotes a probability measure on SLd()\operatorname{SL}_{d}(\mathbb{Z}). We assume that the algebraic group GG generated by μ\mu is semisimple and connected, and let EE be the algebra generated by GG in d()\operatorname{\mathcal{M}}_{d}(\mathbb{R}). Recall that for xEx\in E, we defined detE(x)\det_{E}(x) to be the determinant of the map EEE\to E, yxyy\mapsto xy. Note that detE\det_{E} is a homogeneous polynomial function on EE of degree equal to D=dimED=\dim E. Recall also that

μn=(πφn)(μn),\mu_{n}=(\pi^{\prime}\circ\varphi_{n})_{*}(\mu^{*n}),

where π:EE\pi^{\prime}\colon E\to E^{\prime} is the projection to the direct sum E=E1EsE^{\prime}=E_{1}\oplus\dots\oplus E_{s} of all simple ideals with non-zero Lyapunov exponent, and φn:EE\varphi_{n}\colon E\to E is the scaling map defined in (3.2). As before, we write πj:EEj\pi_{j}:E\to E_{j}, j=1,,sj=1,\dots,s for the projection to the simple factors.

Lemma 3.7.

Given ω>0\omega>0 there exists c=c(μ,ω)>0c=c(\mu,\omega)>0 such that the following holds.

n0,yE,μnD({xE||detE(xy)|eωn})ecn.\forall n\geq 0,\,\forall y\in E^{\prime},\quad\mu_{n}^{\boxplus D}\bigl{(}\bigl{\{}\,x\in E^{\prime}\,\big{|}\,\lvert\det\nolimits_{E^{\prime}}(x-y)\rvert\leq e^{-\omega n}\,\bigr{\}}\bigr{)}\ll e^{-cn}.

Note that for all xx in EE^{\prime}, detE(x)=j=1sdetEj(πj(x))\det\nolimits_{E^{\prime}}(x)=\prod_{j=1}^{s}\det\nolimits_{E_{j}}(\pi_{j}(x)) and hence for every n0n\geq 0, and every xEx\in E,

detE(πφn(x))=j=1se(dimEj)λ1(μ,Ej)ndetEj(πj(x)).\det\nolimits_{E^{\prime}}(\pi^{\prime}\circ\varphi_{n}(x))=\prod_{j=1}^{s}e^{-(\dim E_{j})\lambda_{1}(\mu,E_{j})n}\det\nolimits_{E_{j}}(\pi_{j}(x)).

This immediately reduces the proof of Lemma 3.7 to the following.

Lemma 3.8.

Given ω>0\omega>0 there exists c=c(μ,ω)>0c=c(\mu,\omega)>0 such that the following holds for every j=1,,sj=1,\dotsc,s, all n0n\geq 0 and all yEjy\in E_{j},

(μn)D({xE||detEj(πj(x)y)|e(dimEj)λ1(μ,Ej)nωn})ecn.(\mu^{*n})^{\boxplus D}\bigl{(}\bigl{\{}\,x\in E\,\big{|}\,\lvert\det\nolimits_{E_{j}}(\pi_{j}(x)-y)\rvert\leq e^{(\dim E_{j})\lambda_{1}(\mu,E_{j})n-\omega n}\,\bigr{\}}\bigr{)}\ll e^{-cn}.

The idea is to apply [25, Proposition 3.2], where the case where EE is simple was treated. However, upon projecting to a simple factor, the random walk might no longer be defined with integer coefficients: simple factors of EE are only defined over a number field. So we cannot apply [25, Proposition 3.2] as it is stated. Nevertheless, we can remark that, in the proof of [25, Proposition 3.2], [25, Lemma 3.13] holds more generally for the projected random walk from EE to each EjE_{j} and then the rest of the proof of [25, Proposition 3.2] for a projected random walk is identical.

Here is the detailed proof. We need two ingredients from [25]. For a probability measure μ\mu on a semisimple Lie group GG and a finite-dimensional linear representation (ρ,V)(\rho,V) of GG over \mathbb{R}, recall that

λ1(μ,V)=limn+1nGlogρ(g)dμn(g)\lambda_{1}(\mu,V)=\lim_{n\to+\infty}\frac{1}{n}\int_{G}\log\lVert\rho(g)\rVert\,\mathrm{d}\mu^{*n}(g)

denotes the top Lyapunov exponent associated to the random walk induced on VV. By semisimplicity VV is a sum of irreducible sub-representations. The sum of irreducible sub-representations of same top Lyapunov exponent is a sum of isotypical components. For λ\lambda\in\mathbb{R}, we will denote by pλ:VVp_{\lambda}\colon V\to V the GG-equivariant projection onto

VV, irreducibleλ1(μ,V)λV.\sum_{\begin{subarray}{c}V^{\prime}\subset V,\text{ irreducible}\\ \lambda_{1}(\mu,V^{\prime})\geq\lambda\end{subarray}}V^{\prime}.

We also write [G]D\mathbb{R}[G]_{\leq D} for the set of polynomial maps of degree at most DD on GG, i.e. restrictions to GG of polynomial maps of degree at most DD on EE. We fix a norm on [G]D\mathbb{R}[G]_{\leq D}, for instance f=supgBG(1,1)|f(g)|\lVert f\rVert=\sup_{g\in\operatorname{B}_{G}(1,1)}\lvert f(g)\rvert. The following is [25, Proposition 3.17].

Proposition 3.9.

Let μ\mu be a probability measure on SLd()\operatorname{SL}_{d}(\mathbb{Z}) having a finite exponential moment. Let GG denote the Zariski closure of the subgroup generated by supp(μ)\operatorname{supp}(\mu) in SLd()\operatorname{SL}_{d}(\mathbb{R}). Assume that GG is semisimple and Zariski connected. Given D1D\geq 1, λ0\lambda\geq 0, and ω>0\omega>0, there is c=c(μ,D,λ,ω)>0c=c(\mu,D,\lambda,\omega)>0 such that the following holds for every f[G]Df\in\mathbb{R}[G]_{\leq D}.

n0,μn({gG||f(g)|e(λω)npλ(f)})ecn.\forall n\geq 0,\quad\mu^{*n}\bigl{(}\{\,g\in G\,|\,\lvert f(g)\rvert\leq e^{(\lambda-\omega)n}\lVert p_{\lambda}(f)\rVert\,\}\bigr{)}\ll e^{-cn}.

Here pλ:[G]D[G]Dp_{\lambda}\colon\mathbb{R}[G]_{\leq D}\to\mathbb{R}[G]_{\leq D} is defined as above.

The following is [25, Lemma 3.18]. For k1k\geq 1 and a measure μ\mu on GG, μk=μμ\mu^{\otimes k}=\mu\otimes\dotsm\otimes\mu denotes the product measure on Gk=G××GG^{k}=G\times\dotsm\times G. Again, [Gk]D\mathbb{R}[G^{k}]_{\leq D} denotes the space of restrictions to GkG^{k} of polynomial functions of degree at most DD on EkE^{k}.

Lemma 3.10.

Let VV be a Euclidean space. Let μ\mu be a Borel probability measure on SL(V)\operatorname{SL}(V) having a finite exponential moment. Let GG denote the Zariski closure of the subgroup generated by supp(μ)\operatorname{supp}(\mu) in SLd()\operatorname{SL}_{d}(\mathbb{R}). Assume that GG is Zariski connected, is not compact and acts irreducibly on VV.

Let EE denote the \mathbb{R}-span of GG in End(V)\operatorname{End}(V) and kdimEk\geq\dim E an integer. Let D1D\geq 1 an integer and f[E]Df\in\mathbb{R}[E]_{\leq D} be such that its homogeneous part fDf_{D} of degree DD does not vanish on EE. Define F[Gk]DF\in\mathbb{R}[G^{k}]_{\leq D} to be the polynomial function

(x1,,xk)Gk,F(x1,,xk)=f(x1++xk).\forall(x_{1},\dotsc,x_{k})\in G^{k},\quad F(x_{1},\dotsc,x_{k})=f(x_{1}+\dots+x_{k}).

Then we have

pDλ1(μ,V)(F)0p_{D\lambda_{1}(\mu,V)}(F)\neq 0

where pDλ1(μ,V):[Gk]D[Gk]Dp_{D\lambda_{1}(\mu,V)}\colon\mathbb{R}[G^{k}]_{\leq D}\to\mathbb{R}[G^{k}]_{\leq D} denotes the projection to the sum of irreducible GkG^{k}-subrepresentations M[Gk]DM\subset\mathbb{R}[G^{k}]_{\leq D} with λ1(μk,M)Dλ1(μ,V)\lambda_{1}(\mu^{\otimes k},M)\geq D\lambda_{1}(\mu,V).

Remark.

The conclusion of the above lemma can be improved to

pDλ1(μ,V)(F)[Gk]Dμ,D,kfD[E]D.\lVert p_{D\lambda_{1}(\mu,V)}(F)\rVert_{\mathbb{R}[G^{k}]_{\leq D}}\gg_{\mu,D,k}\lVert f_{D}\rVert_{\mathbb{R}[E]_{\leq D}}.

Indeed, it is enough to check it when f=fDf=f_{D} is homogeneous, and then one may assume fD[E]D=1\lVert f_{D}\rVert_{\mathbb{R}[E]_{\leq D}}=1. The left-hand side is a positive continuous function of fDf_{D}, so it admits a uniform positive lower bound on the unit sphere fD[E]D=1\lVert f_{D}\rVert_{\mathbb{R}[E]_{\leq D}}=1. This shows the desired lower bound.

Proof of Lemma 3.8.

Fix j{1,,s}j\in\{1,\dotsc,s\}. Remember that EjE_{j} is a simple algebra over \mathbb{R}. Using Wedderburn’s structure theorem, we can find a real vector space VjV_{j} and an irreducible faithful linear representation EjEnd(Vj)E_{j}\to\operatorname{End}(V_{j}). It is easy to see that λ1(μ,Ej)=λ1(πjμ,Vj)\lambda_{1}(\mu,E_{j})=\lambda_{1}\bigl{(}{\pi_{j}}_{*}\mu,V_{j}\bigr{)}. The Zariski closure of the subgroup generated by supp(πjμ)\operatorname{supp}({\pi_{j}}_{*}\mu) is precisely πj(G)\pi_{j}(G). It spans EjE_{j}, is Zariski connected, acts irreducibly on VjV_{j} and is not compact. Thus, we may apply Lemma 3.10 to πjμ{\pi_{j}}_{*}\mu with D=DjD=D_{j} and k=Dk=D.

Let yEjy\in E_{j} and consider the polynomial function f[Ej]f\in\mathbb{R}[E_{j}], f(x)=detEj(xy)f(x)=\det_{E_{j}}(x-y). The degree of ff is Dj=dimEjD_{j}=\dim E_{j} and its degree DjD_{j} homogeneous part is detEj\det_{E_{j}}. Recall D=dimED=\dim E. Consider F[πj(G)D]DjF\in\mathbb{R}[\pi_{j}(G)^{D}]_{\leq D_{j}} defined as

x1,,xDπj(G),F(x1,,xD)=f(x1++xD).\forall x_{1},\dotsc,x_{D}\in\pi_{j}(G),\quad F(x_{1},\dotsc,x_{D})=f(x_{1}+\dots+x_{D}).

By Lemma 3.10 and the remark that follows it

pDjλ1(μ,Ej)(F)[πj(G)D]DjdetEj[Ej]DjE1.\bigl{\lVert}p_{D_{j}\lambda_{1}(\mu,E_{j})}(F)\bigr{\rVert}_{\mathbb{R}[\pi_{j}(G)^{D}]_{\leq D_{j}}}\gg\lVert\det\nolimits_{E_{j}}\rVert_{\mathbb{R}[E_{j}]_{\leq D_{j}}}\gg_{E}1.

The linear map Θj:[πj(G)D][GD]\Theta_{j}\colon\mathbb{R}[\pi_{j}(G)^{D}]\to\mathbb{R}[G^{D}] obtained by precomposing (πj,,πj)(\pi_{j},\dotsc,\pi_{j}) is injective and sends irreducible πj(G)D\pi_{j}(G)^{D}-subrepresentations to irreducible GDG^{D}-subrepresentations. Moreover, for any irreducible πj(G)D\pi_{j}(G)^{D}-subrepresentation M[πj(G)D]M\subset\mathbb{R}[\pi_{j}(G)^{D}], we have

λ1((πjμ)D,M)=λ1(μD,Θj(M)).\lambda_{1}\bigl{(}({\pi_{j}}_{*}\mu)^{\otimes D},M\bigr{)}=\lambda_{1}\bigl{(}\mu^{\otimes D},\Theta_{j}(M)\bigr{)}.

It follows that

pDjλ1(μ,Ej)(F(πj,,πj))[GD]DjE1.\bigl{\lVert}p_{D_{j}\lambda_{1}(\mu,E_{j})}(F\circ(\pi_{j},\dotsc,\pi_{j}))\bigr{\rVert}_{\mathbb{R}[G^{D}]_{\leq D_{j}}}\gg_{E}1.

Then we obtain Lemma 3.8 by applying Proposition 3.9 to the measure μD\mu^{\otimes D} and the polynomial function F(πj,,πj)[GD]DjF\circ(\pi_{j},\dotsc,\pi_{j})\in\mathbb{R}[G^{D}]_{\leq D_{j}}. ∎

3.3. Proof of Proposition 3.1

In order to obtain the required non-concentration properties for the measure μn\mu_{n}, we shall use the basic large deviation estimates for matrix products that have already been used in the proof of Proposition 3.4. The statement below is taken from Boyer [16, Theorem A.5], which generalizes previous results of Le Page [30] and Bougerol [8, Theorem V.6.2].

Theorem 3.11 (Large deviation estimates).

Let μ\mu be a Borel probability measure on GLd()\operatorname{GL}_{d}(\mathbb{R}) having a finite exponential moment. For any ω>0\omega>0, there is c=c(μ,ω)>0c=c(\mu,\omega)>0, such that the following holds.

  1. (1)

    For all n1n\geq 1,

    μn({gΓ||1nloggλ1(μ,d)|ω})ωecn.\mu^{*n}\Bigl{(}\Bigl{\{}\,g\in\Gamma\,\Big{|}\,\left\lvert\frac{1}{n}\log\lVert g\rVert-\lambda_{1}(\mu,\mathbb{R}^{d})\right\rvert\geq\omega\,\Bigr{\}}\Bigr{)}\ll_{\omega}e^{-cn}.
  2. (2)

    Assume further that the group generated by supp(μ)\operatorname{supp}(\mu) acts irreducibly on d\mathbb{R}^{d}. For all n1n\geq 1 and all vd{0}v\in\mathbb{R}^{d}\setminus\{0\},

    μn({gΓ||1nloggvvλ1(μ,d)|ω})ωecn.\mu^{*n}\Bigl{(}\Bigl{\{}\,g\in\Gamma\,\Big{|}\,\left\lvert\frac{1}{n}\log\frac{\lVert gv\rVert}{\lVert v\rVert}-\lambda_{1}(\mu,\mathbb{R}^{d})\right\rvert\geq\omega\,\Bigr{\}}\Bigr{)}\ll_{\omega}e^{-cn}.

To prove Proposition 3.1, we shall only need the first item; the second item will be used later in Section 6.

Proof of Proposition 3.1.

Note that condition NC(ε,κ,τ)\operatorname{NC}(\varepsilon,\kappa,\tau) was defined for algebras endowed with a norm, and not with a quasi-norm. However, for some constants α,β>0\alpha,\beta>0, we have, for every vEv\in E^{\prime},

{v|v|\scaleobj0.7α if v1v|v|\scaleobj0.7β if v<1.\left\{\begin{array}[]{ll}\lVert v\rVert\leq{\lvert v\rvert^{\scaleobj{0.7}{\sim}}}^{\,\alpha}&\text{ if }\lVert v\rVert\geq 1\\ \lVert v\rVert\leq{\lvert v\rvert^{\scaleobj{0.7}{\sim}}}^{\,\beta}&\text{ if }\lVert v\rVert<1.\end{array}\right.

So if some measure satisfies condition NC(ε,κ,τ)\operatorname{NC}(\varepsilon,\kappa,\tau) for the quasi-norm ||\scaleobj0.7\lvert\boldsymbol{\,\cdot\,}\rvert^{\scaleobj{0.7}{\sim}} on EE^{\prime}, then it satisfies NC(αε,κβ,τ)\operatorname{NC}(\alpha\varepsilon,\frac{\kappa}{\beta},\tau) for the usual norm \lVert\cdot\rVert on EE^{\prime}. It is therefore sufficient to check the non-concentration properties of μn\mu_{n} for the quasi-distance d~\tilde{d}.

For that, let ε>0\varepsilon>0 be some small parameter. By Theorem 3.111 applied to each πjμ{\pi_{j}}_{*}\mu, there exists τ=τ(μ,ε)>0\tau=\tau(\mu,\varepsilon)>0 such that

μn({gE|j=1,,s,|πj(g)|\scaleobj0.7eεn})1eτn.\mu_{n}\bigl{(}\bigl{\{}\,g\in E^{\prime}\,\big{|}\,\forall j=1,\dotsc,s,\ \lvert\pi_{j}(g)\rvert^{\scaleobj{0.7}{\sim}}\geq e^{-\varepsilon n}\,\bigr{\}}\bigr{)}\geq 1-e^{-\tau n}.

Let ν0\nu_{0} be the restriction of μn\mu_{n} to such gg and write

μn=ν0+ν1\mu_{n}=\nu_{0}+\nu_{1}

so that ν1(E)eτn\nu_{1}(E)\leq e^{-\tau n}. By Proposition 3.2, there exists κ=κ(μ)>0\kappa=\kappa(\mu)>0 such that for any affine hyperplane WEW\subset E with E0WWE_{0}\subset W-W,

ρen,μn({gG|d~(g,W)<ρminjJW|πj(g)|\scaleobj0.7})ρκ.\forall\rho\geq e^{-n},\quad\mu^{*n}\bigl{(}\bigl{\{}\,g\in G\,\big{|}\,\tilde{d}(g,W)<\rho\min_{j\in J_{W}}\lvert\pi_{j}(g)\rvert^{\scaleobj{0.7}{\sim}}\,\bigr{\}}\bigr{)}\ll\rho^{\kappa}.

By definition of φn\varphi_{n} and of the quasi-norm ||\lvert\cdot\rvert on EE, we have |φn(g)|=en|g|\lvert\varphi_{n}(g)\rvert=e^{-n}\lvert g\rvert for every gGg\in G and therefore, for every affine hyperplane WEW\subset E^{\prime},

ρen,μn({gE|d~(g,W)<ρminjJW|πj(g)|\scaleobj0.7})ρκ.\forall\rho\geq e^{-n},\quad\mu_{n}\bigl{(}\bigl{\{}\,g\in E^{\prime}\,\big{|}\,\tilde{d}(g,W)<\rho\min_{j\in J_{W}}\lvert\pi_{j}(g)\rvert^{\scaleobj{0.7}{\sim}}\,\bigr{\}}\bigr{)}\ll\rho^{\kappa}.

By definition of ν0\nu_{0}, this implies

(3.8) ρen,ν0({gE|d~(g,W)<ρeεn})ρκ\forall\rho\geq e^{-n},\quad\nu_{0}\bigl{(}\bigl{\{}\,g\in E^{\prime}\,\big{|}\,\tilde{d}(g,W)<\rho e^{-\varepsilon n}\,\bigr{\}}\bigr{)}\ll\rho^{\kappa}

and this inequality is still valid for any convolution ν0η\nu_{0}\boxplus\eta, where η\eta is a finite measure with η(E)1\eta(E)\leq 1. On the other hand, Lemma 3.7 shows that for some τ1>0\tau_{1}>0,

(3.9) yE,μnD({xE||detE(xy)|eεn})eτ1n.\forall y\in E^{\prime},\quad\mu_{n}^{\boxplus D}\bigl{(}\bigl{\{}\,x\in E^{\prime}\,\big{|}\,\lvert\det\nolimits_{E^{\prime}}(x-y)\rvert\leq e^{-\varepsilon n}\,\bigr{\}}\bigr{)}\ll e^{-\tau_{1}n}.

Let η0\eta_{0} be the restriction of ν0μn(D1)\nu_{0}\boxplus\mu_{n}^{\boxplus(D-1)} to BE(0,e2εn)\operatorname{B}_{E^{\prime}}(0,e^{2\varepsilon n}), and write

μnD=η0+η1.\mu_{n}^{\boxplus D}=\eta_{0}+\eta_{1}.

By Theorem 3.111, we have η1(E)eτn\eta_{1}(E^{\prime})\leq e^{\tau n} for some τ2=τ2(ε)>0\tau_{2}=\tau_{2}(\varepsilon)>0, and by equations (3.8) and (3.9), the measure η0\eta_{0} satisfies NC0(2ε,κ2,τ)\operatorname{NC}_{0}(2\varepsilon,\frac{\kappa}{2},\tau) with τ=min(τ1,τ2)\tau=\min(\tau_{1},\tau_{2}). ∎

4. Fourier spectrum of the random walk

Let μ\mu be a probability measure on GLd()\operatorname{GL}_{d}(\mathbb{Z}). Denote by ΓGLd()\Gamma\subset\operatorname{GL}_{d}(\mathbb{Z}) the subgroup generated by supp(μ)\operatorname{supp}(\mu) and GGLd()G\subset\operatorname{GL}_{d}(\mathbb{R}) the Zariski closure of Γ\Gamma in GLd()\operatorname{GL}_{d}(\mathbb{R}). Under the assumption that μ\mu has a finite exponential moment and that GG is semisimple, we want to show some Fourier decay property for the measure μn\mu^{*n} on the algebra EE generated by GG.

4.1. Connected case

The result we need about Fourier decay for random walks is particularly transparent and easy to prove when the algebraic group GG generated by μ\mu is Zariski connected. So we first explain this particular case. Recall that φn:EE\varphi_{n}\colon E\to E is the rescaling automorphism given by (3.2), that π:EE\pi^{\prime}\colon E\to E^{\prime} denotes the projection to the direct sum of all non-compact factors in EE, and that for any integer n1n\geq 1, we let

μn=(πφn)(μn)\mu_{n}=(\pi^{\prime}\circ\varphi_{n})_{*}(\mu^{*n})

be the image of μn\mu^{*n} after rescaling and projection to EE^{\prime}. The proof of the Fourier decay for μn\mu_{n} will be a consequence of the results of Section 2 for multiplicative convolutions on semisimple algebras, and of the multiplicative structure of μn\mu_{n} simply expressed as

m,n,μn+m=μmμn.\forall m,n,\quad\mu_{n+m}=\mu_{m}*\mu_{n}.

We will denote by E{E^{\prime}}^{*} the space of linear forms on EE^{\prime} over the real numbers.

Theorem 4.1 (Fourier decay for random walks in EE^{\prime}).

Assume that GG is semisimple and Zariski connected, and that μ\mu has a finite exponential moment. Then there exists α0=α0(μ)>0\alpha_{0}=\alpha_{0}(\mu)>0 such that for every α1(0,α0)\alpha_{1}\in(0,\alpha_{0}), there exists c0=c0(μ,α1)>0c_{0}=c_{0}(\mu,\alpha_{1})>0 such that for all nn sufficiently large, for all ξE\xi\in{E^{\prime}}^{*} with

eα1nξeα0ne^{\alpha_{1}n}\leq\lVert\xi\rVert\leq e^{\alpha_{0}n}

the following estimate on the Fourier transform of μn\mu^{*n} holds:

|μn^(ξ)|ec0n.\left\lvert\widehat{\mu_{n}}(\xi)\right\rvert\leq e^{-c_{0}n}.

We let EE^{\prime} act on E{E^{\prime}}^{*} on the right by

ξE,x,yE,(ξx)(y)=ξ(xy).\forall\xi\in{E^{\prime}}^{*},\;\forall x,y\in E^{\prime},\quad(\xi\cdot x)(y)=\xi(xy).

Moreover, we let EE act on E{E^{\prime}}^{*} via π\pi^{\prime}.

Proof.

Let D=dimED=\dim E^{\prime}, ε=ε(E,κ,D)\varepsilon=\varepsilon(E^{\prime},\kappa,D) and s=s(E,κ)s=s(E^{\prime},\kappa) be the quantities given by Corollary 2.11. By Proposition 3.1, given α1(0,1)\alpha_{1}\in(0,1), there exists κ>0\kappa>0 such that for any ε>0\varepsilon>0, there exists τ>0\tau>0 such that μnD\mu_{n}^{\boxplus D} satisfies NC(α1ε2,κ,τ)\operatorname{NC}(\frac{\alpha_{1}\varepsilon}{2},\kappa,\tau) at scale ene^{-n} in EE^{\prime} for all nn sufficiently large. This formally implies that μnD\mu_{n}^{\boxplus D} satisfies NC(ε,κ,τ)\operatorname{NC}(\varepsilon,\kappa,\tau) at all scales δ[en,eα1n2]\delta\in[e^{-n},e^{-\frac{\alpha_{1}n}{2}}].

Without loss of generality, we may of course assume that τ(0,εκ)\tau\in(0,\varepsilon\kappa). Let ξE\xi\in{E^{\prime}}^{*} be such that eα1n2ξene^{\frac{\alpha_{1}n}{2}}\leq\lVert\xi\rVert\leq e^{n}. Taking δ=ξ1\delta=\lVert\xi\rVert^{-1}, we have δ[en,eα1n2]\delta\in[e^{-n},e^{-\frac{\alpha_{1}n}{2}}] so μnD\mu_{n}^{\boxplus D} satisfies NC(ε,κ,τ)\operatorname{NC}(\varepsilon,\kappa,\tau) at scale δ\delta. Therefore, Corollary 2.11 shows that μsn=μnμn\mu_{sn}=\mu_{n}*\dots*\mu_{n} satisfies

|μsn^(ξ)|eετn.\lvert\widehat{\mu_{sn}}(\xi)\rvert\leq e^{-\varepsilon\tau n}.

This shows the desired property if nsn\in s\mathbb{Z}. In general, take α0=14s\alpha_{0}=\frac{1}{4s}. For nn large and ξE\xi\in{E^{\prime}}^{*} such that eα1nξeα0ne^{\alpha_{1}n}\leq\lVert\xi\rVert\leq e^{\alpha_{0}n}, write n=sm+rn=sm+r, with 0r<s0\leq r<s and

μn^(ξ)=Gμsm^(ξx)dμr(x).\widehat{\mu_{n}}(\xi)=\int_{G}\widehat{\mu_{sm}}(\xi\cdot x)\,\mathrm{d}\mu_{r}(x).

Then, observe from the exponential moment assumption that outside of a set of μr\mu_{r}-measure at most ecne^{-cn}, one has eα1n2ξξxen2sξe^{-\frac{\alpha_{1}n}{2}}\lVert\xi\rVert\leq\lVert\xi\cdot x\rVert\leq e^{\frac{n}{2s}}\lVert\xi\rVert and so

eα1m2eα1n2ξxen2sen4sem.e^{\frac{\alpha_{1}m}{2}}\leq e^{\frac{\alpha_{1}n}{2}}\leq\lVert\xi\cdot x\rVert\leq e^{\frac{n}{2s}}e^{\frac{n}{4s}}\leq e^{m}.

For such ξx\xi\cdot x, we may bound

|μsm^(ξx)|eετmeετn2s\left\lvert\widehat{\mu_{sm}}(\xi\cdot x)\right\rvert\leq e^{-\varepsilon\tau m}\leq e^{-\frac{\varepsilon\tau n}{2s}}

whence

|μn^(ξ)|eετn2s+ecnec0n\lvert\widehat{\mu_{n}}(\xi)\rvert\leq e^{-\frac{\varepsilon\tau n}{2s}}+e^{-cn}\leq e^{-c_{0}n}

with c0=min(c2,ετ4s)c_{0}=\min(\frac{c}{2},\frac{\varepsilon\tau}{4s}). ∎

4.2. Disconnected case

As before, μ\mu denotes a probability measure on GLd()\operatorname{GL}_{d}(\mathbb{Z}), and GG the algebraic group generated by μ\mu. We still assume that μ\mu has a finite exponential moment and that GG is semisimple but no longer that it is Zariski connected. The identity component GG^{\circ} is then a finite index subgroup in GG. We now write E¯\bar{E} for the subalgebra generated by GG in d()\operatorname{\mathcal{M}}_{d}(\mathbb{R}). As before, we decompose

E¯=E¯1E¯r\bar{E}=\bar{E}_{1}\oplus\dots\oplus\bar{E}_{r}

into simple ideals. The rescaling automorphism φn:E¯E¯\varphi_{n}\colon\bar{E}\to\bar{E} is now defined by

(4.1) φn(g)=j=1renλ1(μ,E¯j)πj(g)\varphi_{n}(g)=\sum_{j=1}^{r}e^{-n\lambda_{1}(\mu,\bar{E}_{j})}\pi_{j}(g)

where λ1(μ,E¯j)\lambda_{1}(\mu,\bar{E}_{j}) denotes the top Lyapunov exponent associated to μ\mu on each of the factors E¯j\bar{E}_{j}. Also, we assume that λ1(μ,E¯j)=0\lambda_{1}(\mu,\bar{E}_{j})=0 if and only if j>sj>s and denote by π:E¯E¯=E¯1E¯s\pi^{\prime}\colon\bar{E}\to\bar{E}^{\prime}=\bar{E}_{1}\oplus\dots\oplus\bar{E}_{s} the projection to the non-compact factors.

Example.

When GG is not Zariski connected, we shall write E¯\bar{E} for the algebra generated by GG, and let EE denote the algebra generated by the identity component GG^{\circ} of GG. Let a0=(1101)a_{0}=\begin{pmatrix}1&1\\ 0&1\end{pmatrix}, a1=(1011)a_{1}=\begin{pmatrix}1&0\\ 1&1\end{pmatrix} and w=(0110)w=\begin{pmatrix}0&1\\ -1&0\end{pmatrix}. Then define by blocks A0=(w00a0)A_{0}=\begin{pmatrix}w&0\\ 0&a_{0}\end{pmatrix} and A1=(w00a1)A_{1}=\begin{pmatrix}w&0\\ 0&a_{1}\end{pmatrix} in SL4()\operatorname{SL}_{4}(\mathbb{Z}) and set

μ=14(δA0+δA1+δA01+δA11).\mu=\frac{1}{4}(\delta_{A_{0}}+\delta_{A_{1}}+\delta_{A_{0}^{-1}}+\delta_{A_{1}^{-1}}).

One has G(/4)×SL2()G\simeq(\mathbb{Z}/4\mathbb{Z})\times\operatorname{SL}_{2}(\mathbb{R}) and E¯×M2()\bar{E}\simeq\mathbb{C}\times M_{2}(\mathbb{R}). On the other hand, the algebra generated by GG^{\circ} is E×M2()E\simeq\mathbb{R}\times M_{2}(\mathbb{R}) if one identifies {(abba);a,b}\mathbb{C}\simeq\{\begin{pmatrix}a&b\\ -b&a\end{pmatrix}\ ;\ a,b\in\mathbb{R}\} and 1\mathbb{R}\simeq\mathbb{R}1. The law μn\mu^{*n} of the random walk at time nn is supported by EE if nn is even, and by A0Ei×M2()A_{0}E\simeq i\mathbb{R}\times M_{2}(\mathbb{R}) if nn is odd. It is always supported on a proper subspace of E¯\bar{E}.

To overcome this issue, we shall use the algebra EE¯E\subset\bar{E} generated by the identity component GG^{\circ} in GG. The group GG^{\circ} has finite index in GG and we let

F=G/G.F=G/G^{\circ}.

With a slight abuse of notation, we identify FF with a set of representatives in GG and write GG as a disjoint union

G=γFγG.G=\bigsqcup_{\gamma\in F}\gamma G^{\circ}.

Any measure ν\nu on GG can then be decomposed uniquely in the form

ν=γFγνγ\nu=\sum_{\gamma\in F}\gamma_{*}\nu_{\gamma}

where each νγ\nu_{\gamma} is a measure on EE. Finally, we let E=π(E)E^{\prime}=\pi^{\prime}(E), and for n1n\geq 1 and γF\gamma\in F,

μn,γ=(πφn)[(μn)γ].\mu_{n,\gamma}=(\pi^{\prime}\circ\varphi_{n})_{*}[(\mu^{*n})_{\gamma}].

Fourier decay for (integer coefficient) random walks on non-connected semisimple groups can be stated as follows.

Theorem 4.2 (Fourier decay for random walks in EE^{\prime}).

Let μ\mu, GG, GG^{\circ} and FF be as above. Then there exists α0=α0(μ)>0\alpha_{0}=\alpha_{0}(\mu)>0 such that for every α1(0,α0)\alpha_{1}\in(0,\alpha_{0}), there exists c0=c0(μ,α1)>0c_{0}=c_{0}(\mu,\alpha_{1})>0 such that for all nn sufficiently large, all γF\gamma\in F and ξE\xi\in{E^{\prime}}^{*} with

eα1nξeα0ne^{\alpha_{1}n}\leq\lVert\xi\rVert\leq e^{\alpha_{0}n}

the following estimate on the Fourier transform of μn\mu^{*n} holds:

|μn,γ^(ξ)|ec0n.\left\lvert\widehat{\mu_{n,\gamma}}(\xi)\right\rvert\leq e^{-c_{0}n}.

One can show that the above theorem is still valid under the assumption that the measure μ\mu is supported on the group GLd(¯)\operatorname{GL}_{d}(\overline{\mathbb{Q}}) of matrices with algebraic coefficients. It seems a difficult problem to prove the same statement without any such assumption on the support of μ\mu.

4.3. Induced random walk on the identity component

In order to prove Theorem 4.2, we shall use the induced random walk on GG^{\circ}, whose definition is given below. Since by definition GG^{\circ} is connected, this will allow us to use the results of Section 3. The drawback is that we can no longer use the simple identity μsn=μnμn\mu_{sn}=\mu_{n}*\dots*\mu_{n}; so we shall have to write μsn\mu_{sn} as a weighted sum of convolutions related to the induced measure μ\mu^{\circ} on the identity component, which makes the argument more technical. The argument is identical to the one given in [28, Appendix B], but we include it for completeness.


Let (gn)n1(g_{n})_{n\geq 1} be a sequence of independent random variables distributed according to μ\mu. Consider the return times to GG^{\circ},

τ(1)=inf{n1gng1G}\tau(1)=\inf\{\,n\geq 1\mid g_{n}\dotsm g_{1}\in G^{\circ}\,\}

and recursively for m2m\geq 2,

τ(m)=inf{n>τ(m1)gng1G}.\tau(m)=\inf\{\,n>\tau(m-1)\mid g_{n}\dotsm g_{1}\in G^{\circ}\,\}.

Those are the return times of a Markov chain on the finite space G/GG/G^{\circ}, so that for every m1m\geq 1, τ(m)\tau(m) is almost surely finite. In fact, by Kac’s formula [7, Lemma 5.4]

𝔼[τ(1)]=[G:G].\mathbb{E}[\tau(1)]=[G:G^{\circ}].

The random variables (gτ(m)gτ(m1)+1)m0(g_{\tau(m)}\dots g_{\tau(m-1)+1})_{m\geq 0} are independent and identically distributed with law μ\mu^{\circ}, the law of gτ(1)g1g_{\tau(1)}\dotsm g_{1}. Note that μ\mu^{\circ} is a probability measure on GG^{\circ} and has the following properties [1, Lemmas 4.40 and 4.42].

Lemma 4.3.

Let μ\mu be a probability measure on a real algebraic group GG and μ\mu^{\circ} the induced measure on the identity component GG^{\circ}. Let T=[G:G]T=[G:G^{\circ}]. If μ\mu admits some finite exponential moment, then:

  1. (1)

    The measure μ\mu^{\circ} has some finite exponential moment;

  2. (2)

    For every ω>0\omega>0, there exists c=c(μ,ω)>0c=c(\mu,\omega)>0 such that for all mm sufficiently large, [|τ(m)Tm|ωm]ecm.\mathbb{P}\bigl{[}\lvert\tau(m)-Tm\rvert\geq\omega m\bigr{]}\leq e^{-cm}.

In order to prove Theorem 4.2, we shall need to relate the random walk defined by μ\mu and the one defined by μ\mu^{\circ}. For that, we introduce, for m1m\geq 1 and 1\ell\geq 1, the law ν\nu_{\ell} of the random variable

gτ(m)g1conditional to the eventτ(m)=.g_{\tau(m)}\dotsm g_{1}\quad\text{conditional to the event}\quad\tau(m)=\ell.

Naturally, ν\nu_{\ell} is also the law of the variable gg1g_{\ell}\dotsm g_{1} conditional to τ(m)=\tau(m)=\ell. On the one hand, we may relate the measures ν\nu_{\ell} to (μ)m(\mu^{\circ})^{*m} with the formula

(4.2) (μ)m=pν.(\mu^{\circ})^{*m}=\sum_{\ell\in\mathbb{N}}p_{\ell}\nu_{\ell}.

where p=[τ(m)=]p_{\ell}=\mathbb{P}[\tau(m)=\ell]. Here, we are hiding the dependency of ν\nu_{\ell} and pp_{\ell} on mm in order to make notation less cumbersome.

On the other hand, writing 1++s+k=n\ell_{1}+\dotsb+\ell_{s}+k=n for some natural integers n,sn,s and 1,,s\ell_{1},\dots,\ell_{s}, we have

(4.3) μn=1++s+k=np1psμkνsν1+(([τ(sm)>n]))\mu^{*n}=\sum_{\ell_{1}+\dotsb+\ell_{s}+k=n}p_{\ell_{1}}\dotsm p_{\ell_{s}}\mu^{*k}*\nu_{\ell_{s}}*\dotsm*\nu_{\ell_{1}}+((\mathbb{P}[\tau(sm)>n]))

where the notation ((t))((t)) for some positive quantity tt means some unspecified positive measure of total mass at most tt. These two formulae will allow us to use the non-concentration properties of (μ)m(\mu^{\circ})^{*m} to prove some Fourier decay estimate for μn\mu^{*n}.

Before we derive Theorem 4.2, we note that the scaling automorphism φm\varphi_{m}^{\circ} on EE associated to μ\mu^{\circ} is simply given by φn=φmT\varphi_{n}^{\circ}=\varphi_{mT}, where T=[G:G]T=[G:G^{\circ}]. This readily follows from the fact that if E¯i\bar{E}_{i} is any simple ideal in E¯\bar{E} and VV any GG^{\circ}-irreducible submodule of E¯i\bar{E}_{i}, then λ1(μ,V)=Tλ1(μ,E¯i)\lambda_{1}(\mu^{\circ},V)=T\lambda_{1}(\mu,\bar{E}_{i}).

Proof of Theorem 4.2.

Let α1>0\alpha_{1}>0 be a given small number. Since the algebraic group generated by μ\mu^{\circ} is connected, Proposition 3.1 applies to the induced random walk on GG^{\circ}. We let κ=κ(μ)>0\kappa=\kappa(\mu^{\circ})>0 be the constant given by that proposition. Let D=dimED=\dim E and s=s(E,κ)1s=s(E,\kappa)\geq 1 and ε=ε(E,κ,D)>0\varepsilon=\varepsilon(E,\kappa,D)>0 be the constants given by Corollary 2.11.

Given α1>0\alpha_{1}>0, Proposition 3.1 shows that for all mm large enough, the measure

(πφmT)(((μ)m)D((μ)m)D)(\pi^{\prime}\circ\varphi_{mT})_{*}\bigl{(}((\mu^{\circ})^{*m})^{\boxplus D}\boxminus((\mu^{\circ})^{*m})^{\boxplus D}\bigr{)}

satisfies NC(α1ε2,κ,τ)\operatorname{NC}(\frac{\alpha_{1}\varepsilon}{2},\kappa,\tau) in EE^{\prime} at scale eme^{-m} for some τ>0\tau>0. This implies that the same measure satisfies NC(ε2,κ,τ)\operatorname{NC}(\frac{\varepsilon}{2},\kappa,\tau) in EE^{\prime} at all scales δ[em,eα1m]\delta\in{[e^{-m},e^{-\alpha_{1}m}]}. Without loss of generality, we may assume that τ<κε/2\tau<\kappa\varepsilon/2 and τ<ε/2\tau<\varepsilon/2.

Let ω=ω(μ,α1)\omega=\omega(\mu,\alpha_{1}) be a constant whose value is to be determined later. Fix n1n\geq 1 large, and set m=(12ω)nTsm=\left\lfloor(1-2\omega)\frac{n}{Ts}\right\rfloor, where T=[G:G]T=[G:G^{\circ}]. Everything below is true for nn sufficiently large (larger than some n0n_{0} depending on μ\mu and α1\alpha_{1}). The letter cc denotes a small positive constant, whose value may vary from one line to the other, depending on μ\mu and α1\alpha_{1} but independent of nn.

By Lemma 4.3, we have

[τ(sm)>nωn]ecn\mathbb{P}[\tau(sm)>n-\omega n]\leq e^{-cn}

and

[τ(sm)<n3ωn]ecn.\mathbb{P}[\tau(sm)<n-3\omega n]\leq e^{-cn}.

Put

={peα1τ2Dm}.\mathcal{L}=\{\,\ell\in\mathbb{N}\mid p_{\ell}\geq e^{-\frac{\alpha_{1}\tau}{2D}m}\,\}.

We can bound

(1,,s)sp1pssneα1τ2Dmecn.\sum_{(\ell_{1},\dotsc,\ell_{s})\not\in\mathcal{L}^{s}}p_{\ell_{1}}\dotsm p_{\ell_{s}}\leq sne^{-\frac{\alpha_{1}\tau}{2D}m}\leq e^{-cn}.

Thus, (4.3) becomes

μn=1,,s,ωnk3ωn1++s+k=np1psμkνsν1+((ecn)).\mu^{*n}=\sum_{\begin{subarray}{c}\ell_{1},\dotsc,\ell_{s}\in\mathcal{L},\,\omega n\leq k\leq 3\omega n\\ \ell_{1}+\dotsb+\ell_{s}+k=n\end{subarray}}p_{\ell_{1}}\dotsm p_{\ell_{s}}\mu^{*k}*\nu_{\ell_{s}}*\dotsm*\nu_{\ell_{1}}+((e^{-cn})).

Let γF\gamma\in F. To finish the proof of the theorem, it suffices to establish an upper bound of the form ecne^{-cn} for the quantity

I1,,s,k(ξ):=\displaystyle I_{\ell_{1},\dotsc,\ell_{s},k}(\xi):= γGe(ξπφn(γ1g))d(μkνsν1)(g)\displaystyle\int_{\gamma G^{\circ}}e\bigl{(}\xi\circ\pi^{\prime}\circ\varphi_{n}(\gamma^{-1}g)\bigr{)}\,\mathrm{d}\bigl{(}\mu^{*k}*\nu_{\ell_{s}}*\dotsm*\nu_{\ell_{1}}\bigr{)}(g)
=\displaystyle= gγGe(ξπ(φnsmT(γ1g)φsmT(h)))dμk(g)d(νsν1)(h)\displaystyle\iint_{g\in\gamma G^{\circ}}e\bigl{(}\xi\circ\pi^{\prime}\bigl{(}\varphi_{n-smT}(\gamma^{-1}g)\varphi_{smT}(h)\bigr{)}\bigr{)}\,\mathrm{d}\mu^{*k}(g)\,\mathrm{d}(\nu_{\ell_{s}}*\dotsm*\nu_{\ell_{1}})(h)
=\displaystyle= γG((πφsmT)(νsν1))(ξφnsmT(γ1g))dμk(g)\displaystyle\int_{\gamma G^{\circ}}\bigl{(}(\pi^{\prime}\circ\varphi_{smT})_{*}(\nu_{\ell_{s}}*\dotsm*\nu_{\ell_{1}})\bigr{)}^{\wedge}\bigl{(}\xi\cdot\varphi_{n-smT}(\gamma^{-1}g)\bigr{)}\,\mathrm{d}\mu^{*k}(g)

uniformly for all 1,,s\ell_{1},\dotsc,\ell_{s}\in\mathcal{L} and ωnk3ωn\omega n\leq k\leq 3\omega n with 1++s+k=n\ell_{1}+\dotsb+\ell_{s}+k=n.

First, we claim that uniformly for all \ell\in\mathcal{L}, the measure

(πφmT)(νDνD)(\pi^{\prime}\circ\varphi_{mT})_{*}\bigl{(}\nu_{\ell}^{\boxplus D}\boxminus\nu_{\ell}^{\boxplus D}\bigr{)}

satisfies NC(ε,κ,τ/2)\operatorname{NC}(\varepsilon,\kappa,\tau/2) in EE^{\prime} at all scales δ[em,e2α1m]\delta\in{[e^{-m},e^{-2\alpha_{1}m}]}, provided that m1m\geq 1 is large enough. Indeed, developing ((μ)m)D((μ)m)D((\mu^{\circ})^{*m})^{\boxplus D}\boxminus((\mu^{\circ})^{*m})^{\boxplus D} using (4.2), we see that for any 1\ell\geq 1,

((μ)m)D((μ)m)D=p2D(νDνD)+((1)).((\mu^{\circ})^{*m})^{\boxplus D}\boxminus((\mu^{\circ})^{*m})^{\boxplus D}=p_{\ell}^{2D}\bigl{(}\nu_{\ell}^{\boxplus D}\boxminus\nu_{\ell}^{\boxplus D}\bigr{)}+((1)).

Observe that given two measures η,η\eta,\eta^{\prime} such that η=δση+((1))\eta=\delta^{\sigma}\eta^{\prime}+((1)), if η\eta satisfies NC(ε,κ,τ)\operatorname{NC}(\varepsilon,\kappa,\tau), then η\eta^{\prime} satisfies NC(ε+σ,κ,τσ)\operatorname{NC}(\varepsilon+\sigma,\kappa,\tau-\sigma). Therefore, the inequality p2Deα1τmp_{\ell}^{2D}\geq e^{-\alpha_{1}\tau m} for \ell\in\mathcal{L} together with the fact that the left-hand side rescaled by πφmT\pi^{\prime}\circ\varphi_{mT} satisfies NC(ε2,κ,τ)\operatorname{NC}(\frac{\varepsilon}{2},\kappa,\tau) in EE^{\prime} at all scales δ[em,eα1m]\delta\in{[e^{-m},e^{-\alpha_{1}m}]} show our claim. By Corollary 2.11, this implies, for all ζ(E)\zeta\in(E^{\prime})^{*} such that e2α1mζeme^{2\alpha_{1}m}\leq\lVert\zeta\rVert\leq e^{m},

|((πφsmT)(νsν1))(ζ)|eα1ετ(2D)smecn.\bigl{\lvert}\bigl{(}(\pi^{\prime}\circ\varphi_{smT})_{*}(\nu_{\ell_{s}}*\dotsm*\nu_{\ell_{1}})\bigr{)}^{\wedge}(\zeta)\bigr{\rvert}\leq e^{-\frac{\alpha_{1}\varepsilon\tau}{(2D)^{s}}m}\leq e^{-cn}.

Note that for any gγGg\in\gamma G^{\circ},

(4.4) ξg11ξφnsmT(γ1g)ξφnsmTg.\lVert\xi\rVert\lVert g^{-1}\rVert^{-1}\ll\lVert\xi\cdot\varphi_{n-smT}(\gamma^{-1}g)\rVert\ll\lVert\xi\rVert\lVert\varphi_{n-smT}\rVert\lVert g\rVert.

On the one hand, we have 0nsmT3ωn0\leq n-smT\leq 3\omega n. Hence, there exists a constant C=C(μ)1C=C(\mu)\geq 1 such that

φnsmTeCωn.\lVert\varphi_{n-smT}\rVert\leq e^{C\omega n}.

On the other hand, using the assumption that μ\mu has a finite exponential moment and Markov’s inequality, we can find a constant C=C(μ)1C=C(\mu)\geq 1 such that for any k1k\geq 1, the μk\mu^{*k}-measure of the set of gΓg\in\Gamma such that

(4.5) geCkandg1eCk\lVert g\rVert\leq e^{Ck}\quad\text{and}\quad\lVert g^{-1}\rVert\leq e^{Ck}

is at least 1ek1-e^{-k}.

Set α0=14Ts\alpha_{0}=\frac{1}{4Ts} and let ξ(E)\xi\in(E^{\prime})^{*} be such that eα1nξeα0ne^{\alpha_{1}n}\leq\lVert\xi\rVert\leq e^{\alpha_{0}n}. Using k3ωnk\leq 3\omega n, we have, for any gsupp(μk)g\in\operatorname{supp}(\mu^{*k}) satisfying (4.5),

e(α14Cω)nξφnsmT(γ1g)e(α0+5Cω)n.e^{(\alpha_{1}-4C\omega)n}\leq\lVert\xi\cdot\varphi_{n-smT}(\gamma^{-1}g)\rVert\leq e^{(\alpha_{0}+5C\omega)n}.

With the choice ω=min{α18C,120CTs}\omega=\min\{\frac{\alpha_{1}}{8C},\frac{1}{20CTs}\}, we can guarantee that this implies

eα1meα1n/2ξφnsmT(γ1g)em.e^{\alpha_{1}m}\leq e^{\alpha_{1}n/2}\leq\lVert\xi\cdot\varphi_{n-smT}(\gamma^{-1}g)\rVert\leq e^{m}.

Putting everything together, we obtain

|I1,,s,k(ξ)|ecn+ekecn+eωn.\lvert I_{\ell_{1},\dotsc,\ell_{s},k}(\xi)\rvert\leq e^{-cn}+e^{-k}\leq e^{-cn}+e^{-\omega n}.

for all 1,,s\ell_{1},\dotsc,\ell_{s}\in\mathcal{L}, ωnk3ωn\omega n\leq k\leq 3\omega n with 1++s+k=n\ell_{1}+\dotsb+\ell_{s}+k=n. This concludes the proof of the theorem. ∎

5. From Fourier decay to granular structure

As in the previous section, μ\mu denotes a probability measure on GLd()\operatorname{GL}_{d}(\mathbb{Z}) and we study the random walk associated to μ\mu on 𝕋d\mathbb{T}^{d}, with starting distribution ν𝒫(𝕋d)\nu\in\mathcal{P}(\mathbb{T}^{d}). The law of the walk at time nn is νn=μnν\nu_{n}=\mu^{*n}*\nu. The goal of this section is to show that if νn\nu_{n} has a large Fourier coefficient, then the starting distribution ν\nu must have some strong concentration property.

5.1. Concentration statement for the random walk

In order to state the main proposition of this section, we need to set up some notation. As before, GG denotes the algebraic subgroup generated by μ\mu, Ed()E\subset\operatorname{\mathcal{M}}_{d}(\mathbb{R}) denotes the algebra generated by the identity component GG^{\circ} of GG, FF denotes the finite group G/GG/G^{\circ} and T=#FT=\#F.

Changing notation slightly, we now consider a decomposition of EE

E=E0E1ErE=E_{0}\oplus E_{1}\oplus\dots\oplus E_{r}

into maximal sums of minimal ideals with same Lyapunov exponent for the action of μ\mu^{\circ}. We assume that the summands are ordered so that

λ1(μ,E1)>>λ1(μ,Er)>0=λ1(μ,E0).\lambda_{1}(\mu^{\circ},E_{1})>\dots>\lambda_{1}(\mu^{\circ},E_{r})>0=\lambda_{1}(\mu^{\circ},E_{0}).

Here, E0E_{0} is eventually trivial.

The group GG acts naturally on the space V=dV=\mathbb{R}^{d} and for 1jr1\leq j\leq r, we let ViV_{i} be the sum of all simple GG^{\circ}-submodules WVW\subset V such that λ1(μ,W)=λ1(μ,Ei)\lambda_{1}(\mu^{\circ},W)=\lambda_{1}(\mu^{\circ},E_{i}). Equivalently, ViV_{i} is also the sum of all simple GG-submodules WVW\subset V such that λ1(μ,W)=1Tλ1(μ,Ei)\lambda_{1}(\mu,W)=\frac{1}{T}\lambda_{1}(\mu^{\circ},E_{i}). We have

V=V0V1Vr.V=V_{0}\oplus V_{1}\oplus\dots\oplus V_{r}.

Let πi:VVi\pi_{i}\colon V\to V_{i} denote the corresponding projection. Define a quasi-norm on VV by

|v|\scaleobj0.7=max0isπi(v)1λ1(μ,Vi)\lvert v\rvert^{\scaleobj{0.7}{\sim}}=\max_{0\leq i\leq s}\lVert\pi_{i}(v)\rVert^{\frac{1}{\lambda_{1}(\mu,V_{i})}}

where by convention

π0(v)10=π0(v)+={0ifπ0(v)1+otherwise.\lVert\pi_{0}(v)\rVert^{\frac{1}{0}}=\lVert\pi_{0}(v)\rVert^{+\infty}=\left\{\begin{array}[]{cl}0&\mbox{if}\ \lVert\pi_{0}(v)\rVert\leq 1\\ +\infty&\mbox{otherwise}.\end{array}\right.

This induces a quasi distance on 𝕋d\mathbb{T}^{d}. For x,y𝕋dx,y\in\mathbb{T}^{d}, define

d~(x,y)={|vw|\scaleobj0.7if there are lifts vV of x and wV of y such that vw12,1otherwise.\tilde{d}(x,y)=\left\{\begin{array}[]{ll}\lvert v-w\rvert^{\scaleobj{0.7}{\sim}}&\quad\text{if there are lifts $v\in V$ of $x$ and $w\in V$ of $y$ such that $\lVert v-w\rVert\leq\frac{1}{2}$,}\\ 1&\quad\text{otherwise.}\end{array}\right.

Neighborhoods of subsets of 𝕋d\mathbb{T}^{d} with respect to this quasi-distance will be denoted by N~bd(,)\operatorname{\tilde{N}bd}(\boldsymbol{\,\cdot\,},\boldsymbol{\,\cdot\,}). Finally, for a rational subspace WVW\subset V, we let WmoddW\;\mathrm{mod}\;\mathbb{Z}^{d} denote its projection in 𝕋d\mathbb{T}^{d}, which is a subtorus.

Proposition 5.1.

Let μ\mu be a probability measure on GLd()\operatorname{GL}_{d}(\mathbb{Z}) with some finite exponential moment and ν\nu be a Borel probability measure on 𝕋d\mathbb{T}^{d}. If the algebraic group GG generated by μ\mu is semisimple, then there exist C=C(μ)0C=C(\mu)\geq 0 and τ>0\tau>0 such that the following holds.

Assume that for some t(0,12)t\in(0,\frac{1}{2}),

|μnν^(a0)|tfor some a0d and nCloga0t.\lvert\widehat{\mu^{*n}*\nu}(a_{0})\rvert\geq t\quad\text{for some }a_{0}\in\mathbb{Z}^{d}\text{ and }n\geq C\log\frac{\lVert a_{0}\rVert}{t}.

Then, there exists γF\gamma\in F such that, denoting

W=(a0γE)W=(a_{0}\gamma E)^{\perp}

there exists a finite subset X𝕋dX\subset\mathbb{T}^{d} such that

(XX)N~bd(Wmodd,e(12τ)n)={0}(X-X)\cap\operatorname{\tilde{N}bd}(W\;\mathrm{mod}\;\mathbb{Z}^{d},e^{-(1-2\tau)n})=\{0\}

and

ν(X+N~bd(Wmodd,e(1τ)n))tO(1).\nu\bigl{(}X+\operatorname{\tilde{N}bd}(W\;\mathrm{mod}\;\mathbb{Z}^{d},e^{-(1-\tau)n})\bigr{)}\geq t^{O(1)}.

The proof of Proposition 5.1 is in two steps: First, using the results of Section 4, one shows that the inequality |μnν^(a0)|t\lvert\widehat{\mu^{*n}*\nu}(a_{0})\rvert\geq t implies that μnν\mu^{*n}*\nu has many large Fourier coefficients (reducing slightly the value of nn) and then, one applies a Fourier analysis lemma originating in the work of Bourgain, Furman, Lindenstrauss and Mozes [11, Proposition 7.5]. We start with the statement and proof of a general version of that lemma adapted to our needs.

5.2. A quantitative version of Wiener’s lemma

Wiener’s lemma in harmonic analysis states that a measure ν\nu on the torus 𝕋d\mathbb{T}^{d} is atom-free if and only if its Fourier series tends to zero in density, i.e. given t>0t>0, the proportion of vectors ada\in\mathbb{Z}^{d} in a large ball B(0,N)B(0,N) such that |ν^(a)|t\lvert\hat{\nu}(a)\rvert\geq t, tends to zero as NN goes to infinity. In their paper [11], Bourgain, Furman, Lindenstrauss and Mozes observed that this statement could be made quantitative: If B(0,N)B(0,N) contains a proportion at least s>0s>0 of integer vectors satisfying |ν^(a)|t\lvert\hat{\nu}(a)\rvert\geq t, then there exists a ball B=B(x,1N)B=B(x,\frac{1}{N}) of radius 1N\frac{1}{N} in 𝕋d\mathbb{T}^{d} such that ν(B)(st)3\nu(B)\gg(st)^{3}, where the involved constant depends only on dd. In order to later be able to use the quasi-norm adapted to a random walk, we need to generalize this statement. It turns out to be most convenient to formulate the lemma in terms of convex sets and polar pairs.


We will need to generalize slightly the notation 𝒩(,δ)\mathcal{N}(\boldsymbol{\,\cdot\,},\delta) of covering number. Instead of covering a set by balls, we will use translates of a convex body. Recall that a convex body is a compact convex set in d\mathbb{R}^{d}, symmetric about 0, i.e. such that B=BB=-B, and containing 0 in its interior. Given a convex body BdB\subset\mathbb{R}^{d}, and AVA\subset V a bounded non-empty subset, we define the covering number of AA by BB by

𝒩(A,B)=min{N1|x1,,xNV,Ai=1N(xi+B)}.\mathcal{N}(A,B)=\min\Bigl{\{}\,N\geq 1\,\Big{|}\,\exists x_{1},\dotsc,x_{N}\in V,\,A\subset\bigcup_{i=1}^{N}(x_{i}+B)\,\Bigr{\}}.

We shall also say that AA is BB-separated, if (AA)B={0}(A-A)\cap B=\{0\}. Let us briefly list some useful properties of covering numbers. These may be used without explicit mention in the rest of the paper; the elementary proofs are left to the reader. Notice that if \lVert\cdot\rVert is the norm on VV for which BB is the unit ball, then 𝒩(,B)\mathcal{N}(\cdot,B) is simply the covering number at scale 11 for the distance associated to the norm.

  • Let f:VWf\colon V\to W be a linear map to another Euclidean space WW. Then, for any set AVA\subset V and any convex body BWB\subset W,

    (5.1) 𝒩(f(A),f(B))𝒩(A,B)\mathcal{N}\bigl{(}f(A),f(B)\bigr{)}\leq\mathcal{N}(A,B)

    with equality if ff is a linear isomorphism.

  • Let BVB^{\prime}\subset V be another convex body, then

    𝒩(A,B)𝒩(A,B)𝒩(B,B).\mathcal{N}(A,B^{\prime})\leq\mathcal{N}(A,B)\mathcal{N}(B,B^{\prime}).
  • The previous point combined with John’s ellipsoid theorem [36, Theorem 2A, page 87] shows that for any convex body BB there is an ellipsoid EE such that for all non-empty subsets AVA\subset V,

    𝒩(A,B)𝒩(A,E)\mathcal{N}(A,B)\asymp\mathcal{N}(A,E)

    where the implied constant in the \asymp notation depends only on dimV\dim V. In particular,

    𝒩(A,2B)𝒩(A,B)\mathcal{N}(A,2B)\asymp\mathcal{N}(A,B)

    within constants depending only on dimV\dim V.

  • In AA, maximal BB-separated subsets have cardinality at least N(A,B)N(A,B).

  • If BB is symmetric and AA is 2B2B-separated then 𝒩(A,B)#A\mathcal{N}(A,B)\geq\#A.

  • For any set AVA\subset V and any convex body BWB\subset W,

    𝒩(A,2B)𝒩(A,B)\mathcal{N}(A,2B)\asymp\mathcal{N}(A,B)

    where the implied constant depends only on dimV\dim V.

  • Let f:VWf\colon V\to W be a surjective linear map between Euclidean spaces. Let B,CVB,C\subset V be convex bodies and let ACA\subset C be a subset of CC. We have

    (5.2) 𝒩(f(A),f(B))𝒩(f(C),f(B))𝒩(A,B)𝒩(C,B)\frac{\mathcal{N}\bigl{(}f(A),f(B)\bigr{)}}{\mathcal{N}\bigl{(}f(C),f(B)\bigr{)}}\gg\frac{\mathcal{N}(A,B)}{\mathcal{N}(C,B)}

    where the implied constant in the \gg notation depends only on dimV\dim V.

Let d\mathbb{R}^{d} be endowed with the usual scalar product. Given a convex body CdC\subset\mathbb{R}^{d}, its polar set CC^{*} is defined by

C={xd|yC,x,y1}.C^{*}=\{\,x\in\mathbb{R}^{d}\,|\,\forall y\in C,\ \langle x,y\rangle\leq 1\,\}.

If CB(0,2)C\supset B(0,2), then CB(0,12)C^{*}\subset B(0,\frac{1}{2}) and we naturally identify CC^{*} with its projection to 𝕋d\mathbb{T}^{d}. The quantitative version of Wiener’s lemma that we need is given by the proposition below.

Proposition 5.2.

Let ν\nu be a probability measure on 𝕋d\mathbb{T}^{d} and write for t>0t>0

At={ad||ν^(a)|t}.A_{t}=\{\,a\in\mathbb{Z}^{d}\,|\,\lvert\hat{\nu}(a)\rvert\geq t\,\}.

Assume that for some convex bodies BCdB\subset C\subset\mathbb{R}^{d} containing B(0,1)B(0,1), we have, for some c0dc_{0}\in\mathbb{Z}^{d} and some s>0s>0

(5.3) 𝒩(At(c0+C),B)s|C||B|.\mathcal{N}(A_{t}\cap(c_{0}+C),B)\geq s\cdot\frac{\lvert C\rvert}{\lvert B\rvert}.

Then there exists a BB^{*}-separated subset X𝕋dX\subset\mathbb{T}^{d} such that

ν(X+C)ds3/2t6.\nu(X+C^{*})\gg_{d}s^{3/2}t^{6}.
Proof.

The implied constants in the Vinogradov notation in this proof depend only on dd. We shall need two auxiliary functions; the first one corresponds to the pair of convex sets (C,C)(C,C^{*}), the second to (B,B)(B,B^{*}):

  1. (1)

    There exists a smooth function ψ:𝕋d0\psi\colon\mathbb{T}^{d}\to\mathbb{R}_{\geq 0} such that

    1. (a)

      𝕋dψ=1\int_{\mathbb{T}^{d}}\psi=1,

    2. (b)

      ψ1|C|𝟙C\psi\ll\frac{1}{\lvert C^{*}\rvert}\operatorname{\mathbbm{1}}_{C^{*}},

    3. (c)

      ψ^𝟙2Cd\widehat{\psi}\gg\operatorname{\mathbbm{1}}_{2C\cap\mathbb{Z}^{d}}.

  2. (2)

    There exists a smooth function φ:𝕋d0\varphi\colon\mathbb{T}^{d}\to\mathbb{R}_{\geq 0} such that

    1. (a)

      φ𝟙B\varphi\gg\operatorname{\mathbbm{1}}_{B^{*}},

    2. (b)

      φ^\widehat{\varphi} is real and positive and φ^1|B|2𝟙B𝟙B1|B|𝟙2B\widehat{\varphi}\ll\frac{1}{\lvert B\rvert^{2}}\operatorname{\mathbbm{1}}_{B}\boxminus\operatorname{\mathbbm{1}}_{B}\leq\frac{1}{\lvert B\rvert}\operatorname{\mathbbm{1}}_{2B}.

One obtains ψ\psi by taking any smooth symmetric bump function supported on 116C\frac{1}{16}C^{*} with integral ψ=1\int\psi=1. The third property follows from the fact that for every ξ2Cd\xi\in 2C\cap\mathbb{Z}^{d} and x116Cx\in\frac{1}{16}C^{*}, one has ξ,x18\langle\xi,x\rangle\leq\frac{1}{8} and hence (e(ξ,x))12\Re\bigl{(}e(\langle\xi,x\rangle)\bigr{)}\geq\frac{1}{2}. The function φ\varphi can be given explicitly by the formula φ(x)=|1|B|a18Bde(a,x)|2\varphi(x)=\Bigl{\lvert}\frac{1}{\lvert B\rvert}\sum_{a\in\frac{1}{8}B\cap\mathbb{Z}^{d}}e(\langle a,x\rangle)\Bigr{\rvert}^{2} for all xx in 𝕋d\mathbb{T}^{d}. The second item is immediate by definition of φ\varphi, and the first one follows from the fact that by Minkowski’s first theorem on convex bodies, one has #(18Bd)|B|\#(\frac{1}{8}B\cap\mathbb{Z}^{d})\gg\lvert B\rvert.

Pick a maximal 2B2B-separated subset AAt(c0+C)A^{\prime}\subset A_{t}\cap(c_{0}+C) such that all coefficients ν^(a)\hat{\nu}(a), aAa\in A^{\prime} fall in the same quadrant of \mathbb{C}. One still has #A𝒩(At(c0+C),B)\#A^{\prime}\gg\mathcal{N}(A_{t}\cap(c_{0}+C),B) and moreover

|aAν^(a)|t#A2.\Bigl{\lvert}\sum_{a\in A^{\prime}}\hat{\nu}(a)\Bigr{\rvert}\geq\frac{t\#A^{\prime}}{\sqrt{2}}.

By the Cauchy-Schwarz inequality,

a,bAν^(ab)=𝕋d|aAe(a,x)|2dν(x)|𝕋daAe(a,x)dν(x)|2t2(#A)2.\sum_{a,b\in A^{\prime}}\hat{\nu}(a-b)=\int_{\mathbb{T}^{d}}\bigl{\lvert}\sum_{a\in A^{\prime}}e(\langle a,x\rangle)\bigr{\rvert}^{2}\,\mathrm{d}\nu(x)\geq\Bigl{\lvert}\int_{\mathbb{T}^{d}}\sum_{a\in A^{\prime}}e(\langle a,x\rangle)\,\mathrm{d}\nu(x)\Bigr{\rvert}^{2}\gg t^{2}(\#A^{\prime})^{2}.

Hence, there exists a translate AA of AA^{\prime} such that AAA2CA\subset A^{\prime}-A^{\prime}\subset 2C and

|aAν^(a)|t2#A\Bigl{\lvert}\sum_{a\in A}\hat{\nu}(a)\Bigr{\rvert}\gg t^{2}\#A

and

(5.4) #A=#As|C||B|.\#A=\#A^{\prime}\gg s\cdot\frac{\lvert C\rvert}{\lvert B\rvert}.

Consider the function f:𝕋df\colon\mathbb{T}^{d}\to\mathbb{R} defined by

x𝕋d,f(x)=aAe(a,x).\forall x\in\mathbb{T}^{d},\quad f(x)=\sum_{a\in A}e(\langle a,x\rangle).

On the one hand, using the definition of ff, the properties of φ\varphi and the fact that AA is 2B2B-separated, one has, for any y𝕋dy\in\mathbb{T}^{d},

y+B|f|2\displaystyle\int_{y+B^{*}}\lvert f\rvert^{2} =𝕋d𝟙B(xy)|f(x)|2dx\displaystyle=\int_{\mathbb{T}^{d}}\operatorname{\mathbbm{1}}_{B^{*}}(x-y)\lvert f(x)\rvert^{2}\,\mathrm{d}x
a1,a2Aφ(xy)e(a1a2,x)dx\displaystyle\ll\sum_{a_{1},a_{2}\in A}\int\varphi(x-y)e(\langle a_{1}-a_{2},x\rangle)\,\mathrm{d}x
a1,a2Aφ^(a1a2)\displaystyle\leq\sum_{a_{1},a_{2}\in A}\hat{\varphi}(a_{1}-a_{2})
1|B|a1,a2A𝟙2B(a1a2)\displaystyle\ll\frac{1}{\lvert B\rvert}\sum_{a_{1},a_{2}\in A}\operatorname{\mathbbm{1}}_{2B}(a_{1}-a_{2})
#A|B|.\displaystyle\ll\frac{\#A}{\lvert B\rvert}.

On the other hand, from the properties of ψ\psi and of those of AA,

|𝕋dfd(νψ)|=|aAν^(a)ψ^(a)||aAν^(a)|t2#A.\Bigl{\lvert}\int_{\mathbb{T}^{d}}f\,\mathrm{d}(\nu\boxplus\psi)\Bigr{\rvert}=\Bigl{\lvert}\sum_{a\in A}\hat{\nu}(a)\hat{\psi}(a)\Bigr{\rvert}\gg\Bigl{\lvert}\sum_{a\in A}\hat{\nu}(a)\Bigr{\rvert}\gg t^{2}\#A.

Let (yi)iI(y_{i})_{i\in I} be a maximal family of (4B)(4B^{*})-separated points in 𝕋d\mathbb{T}^{d}. Then the translates (yi+B)iI(y_{i}+B^{*})_{i\in I} are disjoint and have a total volume 1\gg 1. By Fubini’s theorem,

dxiIx+yi+Bfd(νψ)=iI|yi+B|𝕋dfd(νψ).\int\,\mathrm{d}x\sum_{i\in I}\int_{x+y_{i}+B^{*}}f\,\mathrm{d}(\nu\boxplus\psi)=\sum_{i\in I}\lvert y_{i}+B^{*}\rvert\int_{\mathbb{T}^{d}}f\,\mathrm{d}(\nu\boxplus\psi).

Hence, translating all yiy_{i} by some x𝕋dx\in\mathbb{T}^{d} if necessary, we may assume that

(5.5) iI|yi+Bfd(νψ)|t2#A.\sum_{i\in I}\Bigl{\lvert}\int_{y_{i}+B^{*}}f\,\mathrm{d}(\nu\boxplus\psi)\Bigr{\rvert}\gg t^{2}\#A.

By the Cauchy-Schwarz inequality, for each iIi\in I,

|yi+Bfd(νψ)|\displaystyle\Bigl{\lvert}\int_{y_{i}+B^{*}}f\,\mathrm{d}(\nu\boxplus\psi)\Bigr{\rvert} yi+B|f|2yi+B(νψ)2\displaystyle\leq\sqrt{\int_{y_{i}+B^{*}}\lvert f\rvert^{2}}\sqrt{\int_{y_{i}+B^{*}}(\nu\boxplus\psi)^{2}}
#A|B|(νψ)(yi+B)maxyi+Bνψ\displaystyle\ll\sqrt{\frac{\#A}{\lvert B\rvert}}\sqrt{(\nu\boxplus\psi)(y_{i}+B^{*})\max_{y_{i}+B^{*}}\nu\boxplus\psi}
ν(yi+B+C)hi#A|C||B|\displaystyle\ll\nu(y_{i}+B^{*}+C^{*})\sqrt{\frac{h_{i}\#A\lvert C\rvert}{\lvert B\rvert}}

where hi=|C|maxyi+Bνψν(yi+B+C)h_{i}=\frac{\lvert C^{*}\rvert\max_{y_{i}+B^{*}}\nu\boxplus\psi}{\nu(y_{i}+B^{*}+C^{*})}. Recalling (5.4) and (5.5), we obtain some constant L=L(d)>1L=L(d)>1 depending only on dd such that

iIν(yi+B+C)hi1/2s1/2t2L.\sum_{i\in I}\nu(y_{i}+B^{*}+C^{*})h_{i}^{1/2}\geq\frac{s^{1/2}t^{2}}{L}.

On the other hand, since for every xx in 𝕋d\mathbb{T}^{d}, one has (νψ)(x)ν(x+C)|C|(\nu\boxplus\psi)(x)\ll\frac{\nu(x+C^{*})}{\lvert C^{*}\rvert}, so

hi=|C|maxyi+Bνψν(yi+B+C)maxxyi+Bν(x+C)ν(yi+B+C)1h_{i}=\frac{\lvert C^{*}\rvert\max_{y_{i}+B^{*}}\nu\boxplus\psi}{\nu(y_{i}+B^{*}+C^{*})}\ll\frac{\max_{x\in y_{i}+B^{*}}\nu(x+C^{*})}{\nu(y_{i}+B^{*}+C^{*})}\leq 1

and therefore, increasing the value of LL if necessary, we may assume that

i,hiL.\forall i,\quad h_{i}\leq L.

Finally, (yi)iI(y_{i})_{i\in I} is 4B4B^{*}-separated so

iIν(yi+B+C)iIν(yi+2B)1\sum_{i\in I}\nu(y_{i}+B^{*}+C^{*})\leq\sum_{i\in I}\nu(y_{i}+2\cdot B^{*})\leq 1

and we may set J={iI|hist44L2}J=\{\,i\in I\,|\,h_{i}\geq\frac{st^{4}}{4L^{2}}\,\} to find

iJν(yi+B+C)s1/2t22L3/2.\sum_{i\in J}\nu(y_{i}+B^{*}+C^{*})\geq\frac{s^{1/2}t^{2}}{2L^{3/2}}.

For each iJi\in J, fix xiyi+Bx_{i}\in y_{i}+B^{*} such that

(νψ)(xi)=maxyi+Bνψ(\nu\boxplus\psi)(x_{i})=\max_{y_{i}+B^{*}}\nu\boxplus\psi

and let

X={xi;iJ}.X=\{x_{i}\ ;\ i\in J\}.

Since the family (yi)iI(y_{i})_{i\in I} is 4B4B^{*}-separated, the set XX is BB^{*}-separated. For the second property, note that for each ii in JJ,

ν(xi+C)|C|(νψ)(xi)=hiν(yi+B+C)st4ν(yi+B+C).\nu(x_{i}+C^{*})\geq\lvert C^{*}\rvert(\nu\boxplus\psi)(x_{i})=h_{i}\nu(y_{i}+B^{*}+C^{*})\gg st^{4}\nu(y_{i}+B^{*}+C^{*}).

so that

ν(X+C)=iJν(xi+C)st4iJν(yi+B+C)s3/2t6.\nu(X+C^{*})=\sum_{i\in J}\nu(x_{i}+C^{*})\gg st^{4}\sum_{i\in J}\nu(y_{i}+B^{*}+C^{*})\gg s^{3/2}t^{6}.

5.3. Proof of Proposition 5.1

To prove Proposition 5.1, we follow the same pattern as in [25]. The only difference here is that we need to find the correct polar pairs (B,B)(B,B^{*}) and (C,C)(C,C^{*}) to apply Proposition 5.2. Before we start the proof, we record to elementary lemmas from [25].

Lemma 5.3 ([25, Lemma 4.3] Additive structure of large Fourier coefficients).

Let μ\mu be a Borel probability measure on SLd()\operatorname{SL}_{d}(\mathbb{Z}) and ν\nu a Borel probability measure on 𝕋d\mathbb{T}^{d}. If

|μν^(a0)|t0>0,\lvert\widehat{\mu*\nu}(a_{0})\rvert\geq t_{0}>0,

then for any integer k1k\geq 1, the set

A={gd()|ν^(a0g)|t02k/2}A=\bigl{\{}g\in\operatorname{\mathcal{M}}_{d}(\mathbb{Z})\mid\lvert\hat{\nu}(a_{0}g)\rvert\geq t_{0}^{2k}/2\bigr{\}}

satisfies

(μkμk)(A)t02k2.\bigl{(}\mu^{\boxplus k}\boxminus\mu^{\boxplus k}\bigr{)}(A)\geq\frac{t_{0}^{2k}}{2}.
Lemma 5.4 ([25, Lemma 4.4] Regularity from Fourier decay).

Given D1D\geq 1 and α>0\alpha>0 sufficiently small, there exist constants c=c(D,α)>0c=c(D,\alpha)>0 and C1=C1(D,α)>0C_{1}=C_{1}(D,\alpha)>0 such that the following holds for all 0<δ<ct0<\delta<ct. Let η\eta be a Borel measure on D\mathbb{R}^{D}, of total mass μ(D)1\mu(\mathbb{R}^{D})\leq 1. Let AA be a subset of D\mathbb{R}^{D}. Assume

  1. (1)

    supp(η)B(0,δα)\operatorname{supp}(\eta)\subset B(0,\delta^{-\alpha}),

  2. (2)

    for all ξD\xi\in\mathbb{R}^{D} with δαξδ1α\delta^{-\alpha}\leq\lVert\xi\rVert\leq\delta^{-1-\alpha}, |η^(ξ)|ξC1\lvert\hat{\eta}(\xi)\rvert\leq\lVert\xi\rVert^{-C_{1}},

  3. (3)

    η(A)t\eta(A)\geq t.

Then there exists aDa\in\mathbb{R}^{D} such that

𝒩(AB(a,δβ),δ)ctD+1(δβδ)D,\mathcal{N}(A\cap B(a,\delta^{\beta}),\delta)\geq ct^{D+1}\Bigl{(}\frac{\delta^{\beta}}{\delta}\Bigr{)}^{D},

where β=(2D+1)α\beta=(2D+1)\alpha.

We are ready to prove the main proposition of this section.

Proof of Proposition 5.1.

We shall use Lemma 5.4 with D=dimED=\dim E^{\prime} and

αmin1irλ1(μ,Vi)3(2D+1)maxiλ1(μ,Vi)soβmin1irλ1(μ,Vi)3max1irλ1(μ,Vi).\alpha\leq\frac{\min_{1\leq i\leq r}\lambda_{1}(\mu,V_{i})}{3(2D+1)\max_{i}\lambda_{1}(\mu,V_{i})}\quad\mbox{so}\quad\beta\leq\frac{\min_{1\leq i\leq r}\lambda_{1}(\mu,V_{i})}{3\max_{1\leq i\leq r}\lambda_{1}(\mu,V_{i})}.

Let C1=C1(D,α)C_{1}=C_{1}(D,\alpha) be as in the lemma. Take α0=α0(μ)\alpha_{0}=\alpha_{0}(\mu) as in Theorem 4.2 with the additional condition that

α0β<min1irλ1(μ,Vi)\alpha_{0}\beta<\min_{1\leq i\leq r}\lambda_{1}(\mu,V_{i})

and set α1=αα02\alpha_{1}=\frac{\alpha\alpha_{0}}{2} and c0=c0(μ,α1)c_{0}=c_{0}(\mu,\alpha_{1}) as in Theorem 4.2. Set also k0=α0C1c0k_{0}=\bigl{\lceil}\frac{\alpha_{0}C_{1}}{c_{0}}\bigr{\rceil} and k=Tk0k=Tk_{0}, where T=#FT=\#F.

Assume

|μnν^(a0)|t.\lvert\widehat{\mu^{*n}*\nu}(a_{0})\rvert\geq t.

By Lemma 5.3, there is a subset Ad()A\subset\operatorname{\mathcal{M}}_{d}(\mathbb{Z}) such that

gA,|ν^(a0g)|t2k2\forall g\in A,\quad\lvert\hat{\nu}(a_{0}g)\rvert\geq\frac{t^{2k}}{2}

and

((μn)k(μn)k)(A)t2k2.\bigl{(}(\mu^{*n})^{\boxplus k}\boxminus(\mu^{*n})^{\boxplus k}\bigr{)}(A)\geq\frac{t^{2k}}{2}.

Note that μn\mu^{*n} is supported on γFγE\bigcup_{\gamma\in F}\gamma E. We can cover μn\mu^{*n} by its restrictions (μn)|γE(\mu^{*n})_{|\gamma E} to each subspace γE\gamma E. Thanks to the choice of kk and the commutativity of additive convolutions, there exists γF\gamma\in F such that

((μn)|γE)k0(a probability measure)(A)t2k2T2k.\left((\mu^{*n})_{|\gamma E}\right)^{\boxplus k_{0}}\boxplus(\text{a probability measure})(A)\geq\frac{t^{2k}}{2T^{2k}}.

Hence for some x1d()x_{1}\in\operatorname{\mathcal{M}}_{d}(\mathbb{Z}), we have

((μn)|γE)k0(x1+A)t2k2T2k.\left((\mu^{*n})_{|\gamma E}\right)^{\boxplus k_{0}}(x_{1}+A)\geq\frac{t^{2k}}{2T^{2k}}.

Let η\eta^{\prime} be the pushforward of ((μn)|γE)k0\left((\mu^{*n})_{|\gamma E}\right)^{\boxplus k_{0}} under the map g(πφn)(γ1g)g\mapsto(\pi^{\prime}\circ\varphi_{n})(\gamma^{-1}g) and let

A=(πφn)(Eγ1(x1+A))EA^{\prime}=(\pi^{\prime}\circ\varphi_{n})\bigl{(}E\cap\gamma^{-1}(x_{1}+A)\bigr{)}\subset E^{\prime}

so that

(5.6) η(A)t2k.\eta^{\prime}(A^{\prime})\gg t^{2k}.

Lemma 5.4 will be used at scale δ=eα0n2\delta=e^{-\frac{\alpha_{0}n}{2}}. By the definition of μn,γ\mu_{n,\gamma} in Section 4, we have η=μn,γk0\eta^{\prime}=\mu_{n,\gamma}^{\boxplus k_{0}}. By Theorem 4.2, for all ξE\xi\in E^{\prime*} with δα=eα1nξeα0n=δ2\delta^{-\alpha}=e^{\alpha_{1}n}\leq\lVert\xi\rVert\leq e^{\alpha_{0}n}=\delta^{-2}, we have

|η^(ξ)|ek0c0nξC1.\lvert\widehat{\eta^{\prime}}(\xi)\rvert\leq e^{-k_{0}c_{0}n}\leq\lVert\xi\rVert^{-C_{1}}.

In view of the large deviation principle for μn\mu^{*n}, we may replace η\eta^{\prime} by its restriction to B(0,δα)B(0,\delta^{-\alpha}) while maintaining (5.6) and the conclusion of Theorem 4.2. Thus by Lemma 5.4 applied to η\eta^{\prime} and AA^{\prime}, there exists x2BE(0,δα)x_{2}\in\operatorname{B}_{E^{\prime}}(0,\delta^{-\alpha}) such that

(5.7) 𝒩(AB(x2,δβ),δ)tO(1)δD(1β).\mathcal{N}\bigl{(}A^{\prime}\cap B(x_{2},\delta^{\beta}),\delta\bigr{)}\gg t^{O(1)}\delta^{-D(1-\beta)}.

Now define convex bodies in EE by

C0=φn(BE(0,δβ)×BE0(0,R))andB0=φn(BE(0,δ)×BE0(0,R))C_{0}=\varphi_{-n}\big{(}\operatorname{B}_{E^{\prime}}(0,\delta^{\beta})\times\operatorname{B}_{E_{0}}(0,R)\big{)}\quad\mbox{and}\quad B_{0}=\varphi_{-n}\big{(}\operatorname{B}_{E^{\prime}}(0,\delta)\times\operatorname{B}_{E_{0}}(0,R)\big{)}

where R=Oμ(k)R=O_{\mu}(k) is a constant large enough so that γ1AE×BE0(0,R)\gamma^{-1}A\subset E^{\prime}\times\operatorname{B}_{E_{0}}(0,R). Note that C0B0C_{0}\supset B_{0} and, since we took α0β<min1irλ1(μ,Vi)\alpha_{0}\beta<\min_{1\leq i\leq r}\lambda_{1}(\mu,V_{i}),

B0BE(0,1).B_{0}\supset\operatorname{B}_{E}(0,1).

Then inequality (5.7) implies that for some x3x_{3} in EE,

𝒩(γ1A(C0+x3),B0)tO(1)δD(1β)tO(1)|C0||B0|.\mathcal{N}(\gamma^{-1}A\cap(C_{0}+x_{3}),B_{0})\gg t^{O(1)}\delta^{-D(1-\beta)}\asymp t^{O(1)}\cdot\frac{\lvert C_{0}\rvert}{\lvert B_{0}\rvert}.

Indeed, with

f1:EExπφn(γ1x1+x)\begin{array}[]{cccc}f_{1}\colon&E&\to&E^{\prime}\\ &x&\mapsto&\pi^{\prime}\circ\varphi_{n}(\gamma^{-1}x_{1}+x)\end{array}

one has f1(γ1A)Af_{1}(\gamma^{-1}A)\supset A^{\prime}, πφn(B0)=BE(0,δ)\pi^{\prime}\circ\varphi_{n}(B_{0})=\operatorname{B}_{E^{\prime}}(0,\delta), and taking x3Ex_{3}\in E^{\prime} such that πφn(γ1x1+x3)=x2\pi^{\prime}\circ\varphi_{n}(\gamma^{-1}x_{1}+x_{3})=x_{2}, f1(C0+x3)=BE(x2,δβ)f_{1}(C_{0}+x_{3})=\operatorname{B}_{E^{\prime}}(x_{2},\delta^{\beta}). The choice of RR guarantees that f1(γ1A(C0+x3))=f1(γ1A)f1(C0+x3)f_{1}\bigl{(}\gamma^{-1}A\cap(C_{0}+x_{3})\bigr{)}=f_{1}(\gamma^{-1}A)\cap f_{1}(C_{0}+x_{3}). One concludes using the inequality (5.1) on f1f_{1}.

Now let

C1=a0γC0andB1=a0γB0C_{1}=a_{0}\gamma C_{0}\quad\mbox{and}\quad B_{1}=a_{0}\gamma B_{0}

and apply (5.2) to f:xa0γxf\colon x\mapsto a_{0}\gamma x to obtain, with c0=a0γx3c_{0}=a_{0}\gamma x_{3},

𝒩(a0A(C1+c0),B1)𝒩(C1,B1)𝒩(γ1A(C0+x3),B0)𝒩(C0,B0)tO(1)\frac{\mathcal{N}\bigl{(}a_{0}A\cap(C_{1}+c_{0}),B_{1}\bigr{)}}{\mathcal{N}(C_{1},B_{1})}\gg\frac{\mathcal{N}\bigl{(}\gamma^{-1}A\cap(C_{0}+x_{3}),B_{0}\bigr{)}}{\mathcal{N}(C_{0},B_{0})}\gg t^{O(1)}

whence

𝒩(a0A(C1+c0),B1)tO(1)|C1|a0γE|B1|a0γE.\mathcal{N}\bigl{(}a_{0}A\cap(C_{1}+c_{0}),B_{1}\bigr{)}\geq t^{O(1)}\frac{\lvert C_{1}\rvert_{a_{0}\gamma E}}{\lvert B_{1}\rvert_{a_{0}\gamma E}}.

Since C1B1C_{1}\subset B_{1} and B1B_{1} contains a ball of radius 11 in a0γEa_{0}\gamma E, we may set

C=C1+Bd(0,2)andB=B1+Bd(0,2)C=C_{1}+\operatorname{B}_{\mathbb{R}^{d}}(0,2)\quad\mbox{and}\quad B=B_{1}+\operatorname{B}_{\mathbb{R}^{d}}(0,2)

to get convex bodies in d\mathbb{R}^{d} containing Bd(0,2)\operatorname{B}_{\mathbb{R}^{d}}(0,2) such that

𝒩(a0A(C+c0),B)tO(1)|C||B|.\mathcal{N}\bigl{(}a_{0}A\cap(C+c_{0}),B\bigr{)}\geq t^{O(1)}\frac{\lvert C\rvert}{\lvert B\rvert}.

Since ν^(a0g)tO(1)\hat{\nu}(a_{0}g)\geq t^{O(1)} for every gAg\in A, Proposition 5.2 shows that there exists a BB^{*}-separated subset X𝕋dX\subset\mathbb{T}^{d} such that

ν(X+C)tO(1).\nu(X+C^{*})\geq t^{O(1)}.

To conclude, it remains to describe the sets BB^{*} and CC^{*}. For that, first consider a decomposition of the space of linear forms on d\mathbb{R}^{d} into irreducible components under the right action of GG^{\circ}

(d)=V=V(1)V(r)(\mathbb{R}^{d})^{*}=V^{\prime}=V^{(1)}\oplus\dots\oplus V^{(r)}

and write pip_{i}, i=1,,ki=1,\dots,k for the corresponding projections. Since a0γa_{0}\gamma is an integer vector, and each V(i)V^{(i)} is defined over a number field, there exists a constant C>0C>0 such that for each ii such that pi(a0γ)0p_{i}(a_{0}\gamma)\neq 0, one has

a0Cpi(a0γ)a0.\lVert a_{0}\rVert^{-C}\ll\lVert p_{i}(a_{0}\gamma)\rVert\ll\lVert a_{0}\rVert.

Therefore, for any ε>0\varepsilon>0, we may choose Cε0C_{\varepsilon}\geq 0 such that nCεloga0tn\geq C_{\varepsilon}\log\frac{\lVert a_{0}\rVert}{t} implies, for i=1,,ki=1,\dots,k,

pi(a0γ)=0oreεnpi(a0γ)eεn.p_{i}(a_{0}\gamma)=0\quad\mbox{or}\quad e^{-\varepsilon n}\leq\lVert p_{i}(a_{0}\gamma)\rVert\leq e^{\varepsilon n}.

Thus, if pi(a0γ)0p_{i}(a_{0}\gamma)\neq 0 and λ1(μ,V(i))0\lambda_{1}(\mu,V^{(i)})\neq 0, then

pi(a0γ)B0BV(i)(0,e(λ1(μ,V(i))+ε)nδ)p_{i}(a_{0}\gamma)B_{0}\subset\operatorname{B}_{V^{(i)}}(0,e^{(\lambda_{1}(\mu,V^{(i)})+\varepsilon)n}\delta)

and

pi(a0γ)C0BV(i)(0,e(λ1(μ,V(i))ε)nδβ).p_{i}(a_{0}\gamma)C_{0}\supset\operatorname{B}_{V^{(i)}}(0,e^{(\lambda_{1}(\mu,V^{(i)})-\varepsilon)n}\delta^{\beta}).

Now consider the decomposition of V=dV=\mathbb{R}^{d} according to Lyapunov exponents

V=V0V1Vr,V=V_{0}\oplus V_{1}\oplus\dots\oplus V_{r},

where for i=0,,ri=0,\dots,r, ViV_{i} is the sum of all irreducible GG-submodules of VV with Lyapunov exponent λ1(μ,Ei)\lambda_{1}(\mu,E_{i}). Since W=(a0γE)W=(a_{0}\gamma E)^{\perp} is a submodule, it can be written

W=W0Wr,whereWi=WVi,i=0,,r.W=W_{0}\oplus\dots\oplus W_{r},\quad\mbox{where}\quad W_{i}=W\cap V_{i},\quad i=0,\dots,r.

An elementary computation based on the above observations shows that for some compact subset A0V0A_{0}\subset V_{0} containing BV0(0,12)\operatorname{B}_{V_{0}}(0,\frac{1}{2}), (we identify subsets of BV(0,12)\operatorname{B}_{V}(0,\frac{1}{2}) with their projections in 𝕋d\mathbb{T}^{d})

BA0×1ir{viVi|d(vi,Wi)e(λ1(μ,Vi)+ε)nδ1andvi12}B^{*}\supset A_{0}\times\prod_{1\leq i\leq r}\{\,v_{i}\in V_{i}\,|\,d(v_{i},W_{i})\leq e^{-(\lambda_{1}(\mu,V_{i})+\varepsilon)n}\delta^{-1}\ \mbox{and}\ \lVert v_{i}\rVert\leq\frac{1}{2}\,\}

and

CA0×1ir{viVi|d(vi,Wi)e(λ1(μ,Vi)ε)nδβandvi12}.C^{*}\subset A_{0}\times\prod_{1\leq i\leq r}\{\,v_{i}\in V_{i}\,|\,d(v_{i},W_{i})\leq e^{-(\lambda_{1}(\mu,V_{i})-\varepsilon)n}\delta^{-\beta}\ \mbox{and}\ \lVert v_{i}\rVert\leq\frac{1}{2}\,\}.

Recalling δ=eα0n2\delta=e^{-\frac{\alpha_{0}n}{2}} and setting τ=α05maxiλ1(μ,Vi)>0\tau=\frac{\alpha_{0}}{5\max_{i}\lambda_{1}(\mu,V_{i})}>0, we may choose ε>0\varepsilon>0 small enough so that

e(λ1(μ,Vi)+ε)nδ1=e(λ1(μ,Vi)+εα02)ne(12τ)λ1(μ,Vi)ne^{-(\lambda_{1}(\mu,V_{i})+\varepsilon)n}\delta^{-1}=e^{-(\lambda_{1}(\mu,V_{i})+\varepsilon-\frac{\alpha_{0}}{2})n}\geq e^{-(1-2\tau)\lambda_{1}(\mu,V_{i})n}

and

e(λ1(μ,Vi)ε)nδβ=e(λ1(μ,Vi)εβα02)ne(1τ)λ1(μ,Vi)n.e^{-(\lambda_{1}(\mu,V_{i})-\varepsilon)n}\delta^{-\beta}=e^{-(\lambda_{1}(\mu,V_{i})-\varepsilon-\frac{\beta\alpha_{0}}{2})n}\leq e^{-(1-\tau)\lambda_{1}(\mu,V_{i})n}.

Finally, since A0A_{0} can be covered by a bounded number of translates of BV0(0,12)\operatorname{B}_{V_{0}}(0,\frac{1}{2}), we may assume A0BV0(0,12)A_{0}\subset\operatorname{B}_{V_{0}}(0,\frac{1}{2}), and then BN~bd(Wmodd,e(12τ)n)B^{*}\supset\operatorname{\tilde{N}bd}(W\;\mathrm{mod}\;\mathbb{Z}^{d},e^{-(1-2\tau)n}) while CN~bd(Wmodd,e(1τ)n)C^{*}\subset\operatorname{\tilde{N}bd}(W\;\mathrm{mod}\;\mathbb{Z}^{d},e^{-(1-\tau)n}). ∎

For some technical reason, we shall have to work on a union of tori 𝕋d×F\mathbb{T}^{d}\times F, where Γ\Gamma acts diagonally. For a measure ν\nu on 𝕋d×F\mathbb{T}^{d}\times F and ada\in\mathbb{Z}^{d}, we write μnν^(a,1)\widehat{\mu^{*n}*\nu}(a,1) for the Fourier coefficient at frequency aa of the restriction of ν\nu to 𝕋d×{1}\mathbb{T}^{d}\times\{1\} viewed as a measure on 𝕋d\mathbb{T}^{d}.

A more careful look at the proof gives us the following slightly more precise version of Proposition 5.1.

Corollary 5.5.

Let μ\mu be a probability measure on GLd()\operatorname{GL}_{d}(\mathbb{Z}) with finite exponential moment. Assume the algebraic group GG generated by μ\mu is semisimple and write F=G/GF=G/G^{\circ}. Let ν\nu be a Borel probability measure on 𝕋d×F\mathbb{T}^{d}\times F. Then there exist C=C(μ)0C=C(\mu)\geq 0 and τ>0\tau>0 such that the following holds.

Assume that for some t(0,12)t\in(0,\frac{1}{2}),

|μnν^(a0,1)|tfor some a0d and nCloga0t.\bigl{\lvert}\widehat{\mu^{*n}*\nu}(a_{0},1)\bigr{\rvert}\geq t\quad\text{for some }a_{0}\in\mathbb{Z}^{d}\text{ and }n\geq C\log\frac{\lVert a_{0}\rVert}{t}.

Then, there exists γF\gamma\in F such that, denoting

W=(a0γE)W=(a_{0}\gamma E)^{\perp}

there exists a finite subset X𝕋d×{γ1}X\subset\mathbb{T}^{d}\times\{\gamma^{-1}\} such that

(XX)N~bd(Wmodd,e(12τ)n)={0}(X-X)\cap\operatorname{\tilde{N}bd}(W\;\mathrm{mod}\;\mathbb{Z}^{d},e^{-(1-2\tau)n})=\{0\}

and

ν(X+N~bd(Wmodd,e(1τ)n))tO(1).\nu\bigl{(}X+\operatorname{\tilde{N}bd}(W\;\mathrm{mod}\;\mathbb{Z}^{d},e^{-(1-\tau)n})\bigr{)}\geq t^{O(1)}.

Here, of course, the addition on 𝕋d×{γ1}\mathbb{T}^{d}\times\{\gamma^{-1}\} is defined for the torus coordinate.

Proof sketch.

Decomposing the measure (μnν)𝕋d×1(\mu^{*n}*\nu)_{\mathbb{T}^{d}\times 1} as

(μnν)|𝕋d×{1}=γFμ|γGnν|𝕋d×{γ1}(\mu^{*n}*\nu)_{|\mathbb{T}^{d}\times\{1\}}=\sum_{\gamma\in F}\mu^{*n}_{|\gamma G^{\circ}}*\nu_{|\mathbb{T}^{d}\times\{\gamma^{-1}\}}

we find that for some γ\gamma, one has, up to a constant depending only on [G:G][G:G^{\circ}],

|(μ|γGnν|𝕋d×{γ1})(a0)|t.\bigl{\lvert}(\mu^{*n}_{|\gamma G^{\circ}}*\nu_{|\mathbb{T}^{d}\times\{\gamma^{-1}\}})^{\wedge}(a_{0})\bigr{\rvert}\gg t.

Let k=dimEk=\dim E^{\prime}. Lemma 5.3 shows that there exists a set Ad()A\subset\operatorname{\mathcal{M}}_{d}(\mathbb{Z}) such that

gA,|ν|𝕋d×{γ1}^(a0g)|t2k.\forall g\in A,\qquad\lvert\widehat{\nu_{|\mathbb{T}^{d}\times\{\gamma^{-1}\}}}(a_{0}g)\rvert\gg t^{2k}.

and

((μ|γGn)k(μ|γGn)k)(A)t2k\left((\mu^{*n}_{|\gamma G^{\circ}})^{\boxplus k}\boxminus(\mu^{*n}_{|\gamma G^{\circ}})^{\boxplus k}\right)(A)\gg t^{2k}

Reasoning as in the proof of Proposition 5.1, we deduce that ν|𝕋d×{γ1}\nu_{|\mathbb{T}^{d}\times\{\gamma^{-1}\}} has many large Fourier coefficients in a0γEa_{0}\gamma E and therefore must be concentrated around a finite subset of well-separated translates of neighborhoods of W=(a0γE)W=(a_{0}\gamma E)^{\perp}. ∎

6. Concentration and unstability of the random walk

In this section, we finally prove the main result of the paper. We consider a probability measure μ\mu on GLd()\operatorname{GL}_{d}(\mathbb{Z}) and the associated random walk on the torus 𝕋d\mathbb{T}^{d}, starting from a point x0𝕋dx_{0}\in\mathbb{T}^{d}. Letting Γ\Gamma be the group generated by suppμ\operatorname{supp}\mu and GG the Zariski closure of Γ\Gamma, we assume that GG is semisimple as an algebraic group, and we show — in a quantitative way — that if the law μnδx0\mu^{*n}*\delta_{x_{0}} of the random walk is not exponentially close to the Haar measure on 𝕋d\mathbb{T}^{d}, then the starting point x0x_{0} is exponentially close to a proper closed invariant subset.


But let us first introduce a new space on which it is convenient to study the random walk, especially to overcome issues related to being Zariski disconnected. As before, GG^{\circ} denotes the identity component of GG, and F=G/GF=G/G^{\circ}. The subalgebra of d()\operatorname{\mathcal{M}}_{d}(\mathbb{R}) generated by GG^{\circ} is denoted by EE. In order to keep track of the coset modulo GG^{\circ}, we let

Y0=𝕋d×FY_{0}=\mathbb{T}^{d}\times F

and let Γ\Gamma act on Y0Y_{0} diagonally.

Let W0W_{0} be a rational GG^{\circ}-invariant subspace of V=dV=\mathbb{R}^{d}. We can define a factor of Y0YY_{0}\to Y by

Y=γFV/(γW0+d)×{γ}.Y=\bigsqcup_{\gamma\in F}V/(\gamma W_{0}+\mathbb{Z}^{d})\times\{\gamma\}.

The action of Γ\Gamma on YY is defined in the obvious way. This way, the natural projection Y0YY_{0}\to Y is Γ\Gamma-equivariant.

Let V0V_{0} denote the sum of all compact factors of GG in V=dV=\mathbb{R}^{d}, that is, the sum of irreducible subrepresentations WVW\subset V such that λ1(μ,W)=0\lambda_{1}(\mu,W)=0. Given a0da_{0}\in\mathbb{Z}^{d} for which the random walk μnδx0\mu^{*n}*\delta_{x_{0}} has large Fourier at a0a_{0}, we shall set

W0=V0+(a0E)W_{0}=V_{0}+(a_{0}E)^{\perp}

and study the random walk on the space YY associated to this W0W_{0}.

In the introduction, we defined the quasi-distance adapted to the random walk on V=dV=\mathbb{R}^{d} and on 𝕋d\mathbb{T}^{d}. Similarly, we can define a quasi-distance on each of the tori V/(d+γW0)V/(\mathbb{Z}^{d}+\gamma W_{0}) in YY. Together, theses quasi-distances define a quasi-distance on YY by the formula

d~((x1,γ1),(x2,γ2))={d~(x1,x2)if γ1=γ2,+otherwise.\tilde{d}\bigl{(}(x_{1},\gamma_{1}),(x_{2},\gamma_{2})\bigr{)}=\begin{cases}\tilde{d}(x_{1},x_{2})&\text{if }\gamma_{1}=\gamma_{2},\\ +\infty&\text{otherwise.}\end{cases}

Below we will state a slightly more precise version of Theorem 1.1. We fix a GG-invariant Euclidean norm on V0V_{0} and write BV0(0,R)\operatorname{B}_{V_{0}}(0,R) for the closed ball in V0V_{0} with radius R>0R>0 with respect to this norm. Given some parameter Q>0Q>0, we note that

BV0(0,Q)+qQ1qdV=d\operatorname{B}_{V_{0}}(0,Q)+\bigcup_{q\leq Q}\frac{1}{q}\mathbb{Z}^{d}\subset V=\mathbb{R}^{d}

is Γ\Gamma-invariant. As a consequence, the set

ZQ=γF(BV0(0,Q)+qQ1qdmod(γW0+d))×{γ}Z_{Q}=\bigsqcup_{\gamma\in F}\bigl{(}\operatorname{B}_{V_{0}}(0,Q)+\bigcup_{q\leq Q}\frac{1}{q}\mathbb{Z}^{d}\;\mathrm{mod}\;(\gamma W_{0}+\mathbb{Z}^{d})\bigr{)}\times\{\gamma\}

is a Γ\Gamma-invariant closed subset of YY.

Theorem 6.1.

Assume that μ\mu has a finite exponential moment and the algebraic group GG is semisimple. Then for every λ(0,1)\lambda\in(0,1), there exist C=C(μ,λ)0C=C(\mu,\lambda)\geq 0 such that the following holds.

Given a0da_{0}\in\mathbb{Z}^{d}, let W0=V0+(a0E)W_{0}=V_{0}+(a_{0}E)^{\perp} and YY be defined as above. For any x0𝕋dx_{0}\in\mathbb{T}^{d}, if

(6.1) |(μnδx0^)(a0)|t for some t(0,12) and nCloga0t,\bigl{\lvert}(\widehat{\mu^{*n}*\delta_{x_{0}}})(a_{0})\bigr{\rvert}\geq t\quad\text{ for some }t\in(0,\frac{1}{2})\text{ and }n\geq C\log\frac{\lVert a_{0}\rVert}{t},

then there is γ0F\gamma_{0}\in F such that writing y0=(x0modγ0W0,γ0)Yy_{0}=(x_{0}\mod\gamma_{0}W_{0},\gamma_{0})\in Y, we have

d~(y0,ZQ)eλn for some Q(a0t)C.\tilde{d}(y_{0},Z_{Q})\leq e^{-\lambda n}\quad\text{ for some }Q\leq\left(\frac{\lVert a_{0}\rVert}{t}\right)^{C}.

To obtain Theorem 1.1 from this theorem, it suffices to lift y0y_{0} and ZQZ_{Q} to Y0=𝕋d×FY_{0}=\mathbb{T}^{d}\times F and then project to 𝕋d\mathbb{T}^{d}.

We proceed to the proof of Theorem 6.1. We fix the meaning of a0da_{0}\in\mathbb{Z}^{d}, W0VW_{0}\subset V, x0𝕋dx_{0}\in\mathbb{T}^{d} as in the statement. By the pigeonhole principle, (6.1) implies that there is γ0F\gamma_{0}\in F such that

(6.2) |μnδ(x0,γ0)^(a0,1)|t#F,\bigl{\lvert}\widehat{\mu^{*n}*\delta_{(x_{0},\gamma_{0})}}(a_{0},1)\bigr{\rvert}\geq\frac{t}{\#F},

where the notation μnδ(x0,γ0)^(a0,1)\widehat{\mu^{*n}*\delta_{(x_{0},\gamma_{0})}}(a_{0},1) is defined in the paragraph preceding Corollary 5.5. This choice of γ0\gamma_{0} determines y0Yy_{0}\in Y. We fix this y0y_{0} for the rest of the proof.

6.1. Bootstrapping concentration

In order to prove Theorem 6.1, we start from the granulation estimate obtained in the previous section as Proposition 5.1. The first step is then to run backwards the random walk to increase the concentration.

Proposition 6.2 (High concentration).

Assume (6.1). Given η>0\eta>0, there exists n1ηloga0tn_{1}\asymp_{\eta}\log\frac{\lVert a_{0}\rVert}{t} and ρ>0\rho>0 with |logρ|n1\lvert\log\rho\rvert\asymp n_{1} such that for some yYy\in Y,

μ(nn1)δy0(B~(y,ρ))ρη.\mu^{*(n-n_{1})}*\delta_{y_{0}}(\tilde{\operatorname{B}}(y,\rho))\geq\rho^{\eta}.
Proof.

Using (6.2) and Corollary 5.5 and observing that (a0γE)=γ1(a0E)(a_{0}\gamma E)^{\perp}=\gamma^{-1}(a_{0}E)^{\perp}, we obtain that for n0loga0tn_{0}\geq\log\frac{\lVert a_{0}\rVert}{t}, there exists an e(12τ)n0e^{-(1-2\tau)n_{0}}-separated subset X0X_{0} contained in a single torus in YY such that

μ(nn0)δy0(N~bd(X0,e(1τ)n0))tC0.\mu^{*(n-n_{0})}*\delta_{y_{0}}\left(\operatorname{\tilde{N}bd}(X_{0},e^{-(1-\tau)n_{0}})\right)\geq t^{C_{0}}.

Increasing C0C_{0} if necessary, we may also assume that #X0eC0n0\#X_{0}\leq e^{C_{0}n_{0}}. Fix some large kk\in\mathbb{N} and then ε>0\varepsilon>0 such that 2kε<1.2k\varepsilon<1. Starting with

m0=τn02d,r0=e(12τ)n0,andρ0=e(1τ)n0,m_{0}=\lfloor\frac{\tau n_{0}}{2d}\rfloor,\quad r_{0}=e^{-(1-2\tau)n_{0}},\quad\text{and}\quad\rho_{0}=e^{-(1-\tau)n_{0}},

we apply Lemma 6.3 below kk times successively. This yields integers mim_{i}, and scales ri>ρir_{i}>\rho_{i}, defined inductively by

{ri+1=emi(1+ε)riρi+1=emi(1ε)ρimi+1=mi(12εd)\left\{\begin{array}[]{l}r_{i+1}=e^{-m_{i}(1+\varepsilon)}r_{i}\\ \rho_{i+1}=e^{-m_{i}(1-\varepsilon)}\rho_{i}\\ m_{i+1}=\lfloor m_{i}(1-\frac{2\varepsilon}{d})\rfloor\\ \end{array}\right.

and at each step, an rir_{i}-separated set XiX_{i} such that #Xi#X0\#X_{i}\leq\#X_{0} and

μ(nn0m0mi)δy0(N~bd(Xi,ρi))(tC02)di.\mu^{*(n-n_{0}-m_{0}-\dotsb-m_{i})}*\delta_{y_{0}}\bigl{(}\operatorname{\tilde{N}bd}(X_{i},\rho_{i})\bigr{)}\geq\left(\frac{t^{C_{0}}}{2}\right)^{d^{i}}.

Notice that by induction on ii, one always has e(d+1)miρirie^{(d+1)m_{i}}\rho_{i}\leq r_{i}, so that Lemma 6.3 may indeed be applied. Moreover, choosing n0loga0tn_{0}\asymp\log\frac{\lVert a_{0}\rVert}{t} large enough (the involved constant will depend on kk, τ\tau, C0C_{0}, etc.), we may ensure that all mim_{i} are large enough so that the error term ecmie^{-cm_{i}} from that lemma is always small compared to (tC02)di\left(\frac{t^{C_{0}}}{2}\right)^{d^{i}}. Set n1=n0+m0++mkn_{1}=n_{0}+m_{0}+\dots+m_{k} and ρ=ρk\rho=\rho_{k}. One has #Xk#X0\#X_{k}\leq\#X_{0} and

μ(nn1)δy0(N~bd(Xk,ρ))(tC02)dk\mu^{*(n-n_{1})}*\delta_{y_{0}}\bigl{(}\operatorname{\tilde{N}bd}(X_{k},\rho)\bigr{)}\geq\left(\frac{t^{C_{0}}}{2}\right)^{d^{k}}

so that for some yXky\in X_{k},

μ(nn1)δy0(B~(y,ρ))1#X0(tC02)dkeC0n0(tC02)dk\mu^{*(n-n_{1})}*\delta_{y_{0}}\left(\tilde{\operatorname{B}}(y,\rho)\right)\geq\frac{1}{\#X_{0}}\left(\frac{t^{C_{0}}}{2}\right)^{d^{k}}\geq e^{-C_{0}n_{0}}\left(\frac{t^{C_{0}}}{2}\right)^{d^{k}}

Now, since m0++mkkm03m_{0}+\dots+m_{k}\geq\frac{km_{0}}{3} we may choose kk large enough so that m0++mk3C0n0ηm_{0}+\dots+m_{k}\geq\frac{3C_{0}n_{0}}{\eta}, and then n0loga0tn_{0}\asymp\log\frac{\lVert a_{0}\rVert}{t} large enough to ensure that

ρ=ρke(1ε)(m0++mk)e2C0n0ηeC0n0η(tC02)dkη.\rho=\rho_{k}\leq e^{-(1-\varepsilon)(m_{0}+\dots+m_{k})}\leq e^{-\frac{2C_{0}n_{0}}{\eta}}\leq e^{-\frac{C_{0}n_{0}}{\eta}}\left(\frac{t^{C_{0}}}{2}\right)^{\frac{d^{k}}{\eta}}.

The proposition follows. ∎

After using Corollary 5.5, we can now forget how W0W_{0} is constructed from a0a_{0}. All what we need is that W0W_{0} is a GG^{\circ}-invariant rational subspace containing V0V_{0}.

We now prove the lemma that was used in the above proof. The notion of rr-separated sets in YY are with respect to the quasi-distance on YY.

Lemma 6.3.

For any ε>0\varepsilon>0 there exist c>0c>0 and m0m_{0}\in\mathbb{N} depending only on μ\mu and ε\varepsilon such that the following holds for any Borel probability measure ν\nu on YY and any mm0m\geq m_{0}.
Let r>0r>0 and ρ>0\rho>0 be such that e(d+1)mρ<re^{(d+1)m}\rho<r. Set

r1=em(1+ε)randρ1=em(1ε)ρ.r_{1}=e^{-m(1+\varepsilon)}r\quad\text{and}\quad\rho_{1}=e^{-m(1-\varepsilon)}\rho.

If XX is an rr-separated subset contained in a single torus in YY, then there is an r1r_{1}-separated subset X1YX_{1}\subset Y, contained in a single torus, with cardinality #X1#X\#X_{1}\leq\#X and such that

ν(N~bd(X1,ρ1))(μmν)(N~bd(X,ρ))decm.\nu\bigl{(}\operatorname{\tilde{N}bd}(X_{1},\rho_{1})\bigr{)}\geq(\mu^{*m}*\nu)\bigl{(}\operatorname{\tilde{N}bd}(X,\rho)\bigr{)}^{d}-e^{-cm}.
Proof.

In this proof, we write X(ρ)X^{(\rho)} for N~bd(X,ρ)\operatorname{\tilde{N}bd}(X,\rho). By Jensen’s inequality and the definition of μmν\mu^{*m}*\nu, (see [11, Lemma 7.6] for details),

(μmν)(X(ρ))dg1,,gdΓμm(g1)μm(gd)ν(g11X(ρ)gd1X(ρ)).(\mu^{*m}*\nu)(X^{(\rho)})^{d}\leq\sum_{g_{1},\dots,g_{d}\in\Gamma}\mu^{*m}(g_{1})\dots\mu^{*m}(g_{d})\nu(g_{1}^{-1}X^{(\rho)}\cap\dots\cap g_{d}^{-1}X^{(\rho)}).

This implies that the set of dd-tuples (gi)1id(g_{i})_{1\leq i\leq d} such that

(6.3) ν(g11X(ρ)gd1X(ρ))(μmν)(X(ρ))decm\nu(g_{1}^{-1}X^{(\rho)}\cap\dots\cap g_{d}^{-1}X^{(\rho)})\geq(\mu^{*m}*\nu)(X^{(\rho)})^{d}-e^{-cm}

has (μm)d(\mu^{*m})^{\otimes d}-measure at least ecme^{-cm}. Using the fact that the large deviation estimates Theorem 3.111 and 2 are valid under the only assumption that the action is irreducible, one readily checks that [25, Lemma 5.5] and its proof are also valid under this assumption. Applying this lemma in each irreducible subrepresentation of d\mathbb{R}^{d}, one obtains that if cc is chosen small enough, there must exist g1,,gdΓg_{1},\dotsc,g_{d}\in\Gamma satisfying (6.3) and moreover,

(6.4) vd/W0{0},e(1ε)mmax1id|giv|\scaleobj0.7|v|\scaleobj0.7\forall v\in\mathbb{R}^{d}/W_{0}\setminus\{0\},\quad e^{(1-\varepsilon)m}\leq\max_{1\leq i\leq d}\frac{\lvert g_{i}v\rvert^{\scaleobj{0.7}{\sim}}}{\lvert v\rvert^{\scaleobj{0.7}{\sim}}}

and (using the large deviation estimates again and the fact that λdimW(μ,W)(dimW1)λ1(μ,W)-\lambda_{\dim W}(\mu,W)\leq(\dim W-1)\lambda_{1}(\mu,W) for any GG-invariant WVW\subset V)

(6.5) i{1,,d},|gi|\scaleobj0.7e(1+ε)mand|gi1|\scaleobj0.7e(d1+ε)m.\forall i\in\{1,\dotsc,d\},\quad\lvert g_{i}\rvert^{\scaleobj{0.7}{\sim}}\leq e^{(1+\varepsilon)m}\quad\text{and}\quad\lvert g_{i}^{-1}\rvert^{\scaleobj{0.7}{\sim}}\leq e^{(d-1+\varepsilon)m}.

We fix such elements g1,,gdg_{1},\dots,g_{d} for the rest of the proof.

Since XX is contained in a single torus, all the gig_{i}’s are contained in the same class in G/GG/G^{\circ}.

We claim that the set g11X(ρ)gd1X(ρ)g_{1}^{-1}X^{(\rho)}\cap\dots\cap g_{d}^{-1}X^{(\rho)} is included in a union of at most #X\#X balls of radius ρ1=e(1ε)mρ\rho_{1}=e^{-(1-\varepsilon)m}\rho. Indeed, from (6.5) and the fact that e(d+1)mρ<re^{(d+1)m}\rho<r, we find, for a given xXx\in X and i1i\geq 1, that the set g11B~(x,ρ)g_{1}^{-1}\tilde{\operatorname{B}}(x,\rho) meets at most one component gi1B~(y,ρ)g_{i}^{-1}\tilde{\operatorname{B}}(y,\rho), yXy\in X. Therefore, there are at most #X\#X non-empty intersections g11B~(x1,ρ)gd1B~(xd,ρ)g_{1}^{-1}\tilde{\operatorname{B}}(x_{1},\rho)\cap\dotsb\cap g_{d}^{-1}\tilde{\operatorname{B}}(x_{d},\rho), for x1,,xdXx_{1},\dotsc,x_{d}\in X.

If x,yx,y lie inside such an intersection, then, for each ii, d~(gix,giy)ρ\tilde{d}(g_{i}x,g_{i}y)\leq\rho. Then (6.4) and (6.5) implies that d~(x,y)e(1ε)mρ=ρ1\tilde{d}(x,y)\leq e^{-(1-\varepsilon)m}\rho=\rho_{1}. Thus, each intersection g11B~(x1,ρ)gd1B~(xd,ρ)g_{1}^{-1}\tilde{\operatorname{B}}(x_{1},\rho)\cap\dots\cap g_{d}^{-1}\tilde{\operatorname{B}}(x_{d},\rho) is included in a ball of radius ρ1\rho_{1}.

Finally, using (6.5), we see that these ρ1\rho_{1}-balls are separated by at least r1=e(1+ε)mrr_{1}=e^{-(1+\varepsilon)m}r. Moreover they are contained in the same torus in YY because the gig_{i}’s are in the same GG^{\circ} coset. This finishes the proof of the proposition. ∎

6.2. A diophantine property

From the high concentration property obtained in the previous paragraph, we want to infer that μ(nn1)δy0\mu^{*(n-n_{1})}*\delta_{y_{0}} is concentrated near a proper Γ\Gamma-invariant subset. The argument relies on a diophantine property of the random walk, coming from the fact that μ\mu is supported on GLd()\operatorname{GL}_{d}(\mathbb{Z}).

Proposition 6.4 (Concentration near a closed invariant subset).

Given β>0\beta>0, there exists C>0C>0 such that the following holds.

Assume (6.1). Then there exist n1n_{1}\in\mathbb{N}^{*} such that 1Cloga0tn1Cloga0t\frac{1}{C}\log\frac{\lVert a_{0}\rVert}{t}\leq n_{1}\leq C\log\frac{\lVert a_{0}\rVert}{t} and ρ[eCn1,en1C]\rho\in[e^{-Cn_{1}},e^{-\frac{n_{1}}{C}}] and QρβQ\leq\rho^{-\beta} such that

μ(nn1)δy0(N~bd(ZQ,ρ))ρβ.\mu^{*(n-n_{1})}*\delta_{y_{0}}\bigl{(}\operatorname{\tilde{N}bd}(Z_{Q},\rho)\bigr{)}\geq\rho^{\beta}.

This proposition is an immediate consequence of Proposition 6.2 and of a diophantine property of the random walk, given by the following lemma.

Lemma 6.5 (Diophantine property).

For every β>0\beta>0, there exist constants CC and η>0\eta>0 depending on μ\mu and β\beta such that for any y0,yYy_{0},y\in Y, if for nC|logρ|n\geq C\lvert\log\rho\rvert, one has

(μnδy0)(B~(y,ρ))ρη(\mu^{*n}*\delta_{y_{0}})(\tilde{\operatorname{B}}(y,\rho))\geq\rho^{\eta}

then d~(y,ZQ)ρ1β\tilde{d}(y,Z_{Q})\leq\rho^{1-\beta} for some QρβQ\leq\rho^{-\beta}.

Proof.

Consider the action of GG on VF=γFV/γW0×{γ}V_{F}=\bigsqcup_{\gamma\in F}V/\gamma W_{0}\times\{\gamma\}. Since V/W0V/W_{0} contains no GG^{\circ}-invariant vector, for every non-zero (v1,v2)(v_{1},v_{2}) in VF×VFV_{F}\times V_{F}, the set {g|gv1=v2}\{\,g\,|\,gv_{1}=v_{2}\,\} is a linear subvariety in GG of dimension less than dimG\dim G. Using the spectral gap property modulo prime numbers [35] (or applying the first step of the proof of Proposition 3.3), we obtain that for mm large enough, for some c0c_{0} independent of (v1,v2)(v_{1},v_{2}),

(6.6) μm({(g1,,gm)|gmg1v1=v2})ec0m.\mu^{\otimes m}(\{\,(g_{1},\dots,g_{m})\,|\,g_{m}\dotsm g_{1}v_{1}=v_{2}\,\})\leq e^{-c_{0}m}.

Fix mm such that

ec0m2ρη>ec0m.e^{-\frac{c_{0}m}{2}}\geq\rho^{\eta}>e^{-c_{0}m}.

If CC is large enough, the condition nC|logρ|n\geq C\lvert\log\rho\rvert ensures that nmn\geq m. From the assumed inequality, it follows that there exists some y1Yy_{1}\in Y such that

(μmδy1)(B~(y,ρ))ρη,(\mu^{*m}*\delta_{y_{1}})(\tilde{\operatorname{B}}(y,\rho))\geq\rho^{\eta},

which implies that the set

Am={(g1,,gm)(suppμ)m|d~(gmg1y1,y)ρ}A_{m}=\{\,(g_{1},\dots,g_{m})\in(\operatorname{supp}\mu)^{m}\,|\,\tilde{d}(g_{m}\dots g_{1}y_{1},y)\leq\rho\,\}

satisfies

μm(Am)ρη>ec0m.\mu^{\otimes m}(A_{m})\geq\rho^{\eta}>e^{-c_{0}m}.

Using the finite exponential moment of μ\mu and reducing the set AmA_{m}, we can assume further that for all (gi)Am(g_{i})\in A_{m}, gmg1eC0m\lVert g_{m}\dotsm g_{1}\rVert\leq e^{C_{0}m} for some C0=C0(μ)C_{0}=C_{0}(\mu). Since all matrices have integer coefficients, the set

Am={gmg1;(g1,,gm)Am}A^{\prime}_{m}=\{g_{m}\dots g_{1}\ ;\ (g_{1},\dots,g_{m})\in A_{m}\}

is then finite.

Write y1=(x1,γ1)y_{1}=(x_{1},\gamma_{1}) and y=(x,γ)y=(x,\gamma). Recalling (6.6) above, one infers that the linear map

θ:V/γ1W0×V/γW0(V/γW0)Am(v1,v2)(gv1v2)gAm\begin{array}[]{lccc}\theta\colon&V/\gamma_{1}W_{0}\times V/\gamma W_{0}&\to&(V/\gamma W_{0})^{A_{m}^{\prime}}\\ &(v_{1},v_{2})&\mapsto&(gv_{1}-v_{2})_{g\in A_{m}^{\prime}}\end{array}

is injective. Moreover, in the canonical bases, its matrix has coefficients in \mathbb{Z} bounded by eC0me^{C_{0}m}, so its inverse has coefficients in 1Q\frac{1}{Q}\mathbb{Z} for some QeC1mQ\leq e^{C_{1}m}, and bounded by eC1me^{C_{1}m}. Therefore, any solution (v1,v2)(v_{1},v_{2}) in V/γ1W0×V/γW0V/\gamma_{1}W_{0}\times V/\gamma W_{0} to

(g1,,gm)Am,gmg1v1v2B~(0,ρ)+dmodγW0\forall(g_{1},\dots,g_{m})\in A_{m},\quad g_{m}\dots g_{1}v_{1}-v_{2}\in\tilde{\operatorname{B}}(0,\rho)+\mathbb{Z}^{d}\;\mathrm{mod}\;\gamma W_{0}\\

can be written, for some w1,w2dw_{1},w_{2}\in\mathbb{Z}^{d} and u1,u2u_{1},u_{2} in B~(0,eC1mρ)\tilde{\operatorname{B}}(0,e^{C_{1}m}\rho),

v1=1Qw1+u1modγ1W0andv2=1Qw2+u2modγW0.v_{1}=\frac{1}{Q}w_{1}+u_{1}\;\mathrm{mod}\;\gamma_{1}W_{0}\quad\text{and}\quad v_{2}=\frac{1}{Q}w_{2}+u_{2}\;\mathrm{mod}\;\gamma W_{0}.

This applies in particular to representatives of (x1,x)(x_{1},x) in V/γ1W0×V/γW0V/\gamma_{1}W_{0}\times V/\gamma W_{0}. It follows that

d~(y,ZeC1m)eC1mρ.\tilde{d}(y,Z_{e^{C_{1}m}})\leq e^{C_{1}m}\rho.

If η>0\eta>0 is chosen so small that 2C1c0η<β\frac{2C_{1}}{c_{0}}\eta<\beta, one has

eC1m=e2C1c0c0m2ρ2C1c0ηρβe^{C_{1}m}=e^{\frac{2C_{1}}{c_{0}}\frac{c_{0}m}{2}}\leq\rho^{-\frac{2C_{1}}{c_{0}}\eta}\leq\rho^{-\beta}

so the lemma is proved. ∎

6.3. Unstability of closed invariant subsets

To conclude the proof of Theorem 6.1, we use a variant of the argument given in [27, §3]. It is based on Foster’s exponential recurrence criterion, applied to a well-chosen function associated to a closed invariant subset. This technique has been used extensively in homogeneous dynamics since the work of Eskin and Margulis [19], in particular by Benoist and Quint for their study of stationary measures [3, 5, 6, 4].

Lemma 6.6 (Margulis inequality).

For every λ(0,1)\lambda\in(0,1), there exist constants C,α>0C,\alpha>0 depending only on μ\mu such that the following holds. For Q2Q\geq 2, define a function φQ:Y{+}\varphi_{Q}\colon Y\to\mathbb{R}\cup\{+\infty\} by

φQ(y)={d~(y,ZQ)αifd~(y,ZQ)>0+otherwise.\varphi_{Q}(y)=\left\{\begin{array}[]{ll}\tilde{d}(y,Z_{Q})^{-\alpha}&\mbox{if}\ \tilde{d}(y,Z_{Q})>0\\ +\infty&\mbox{otherwise}.\end{array}\right.

For all yYy\in Y and all integers n1n\geq 1,

φQ(gy)dμn(g)eλαnφQ(y)+QC.\int\varphi_{Q}(gy)\,\mathrm{d}\mu^{*n}(g)\leq e^{-\lambda\alpha n}\varphi_{Q}(y)+Q^{C}.

The proof of such inequalities is an application of Furstenberg’s law of large numbers [7, Theorem 4.28], using also the exponential moment assumption on μ\mu. Since it is rather standard, we leave it to the reader, and turn to the proof of Theorem 6.1.

Proof of Theorem 6.1.

Let C,α>0C,\alpha>0 be the parameters given by Lemma 6.6 applied with λ=1+λ2\lambda^{\prime}=\frac{1+\lambda}{2} instead of λ\lambda. Then set β=αC+2\beta=\frac{\alpha}{C+2}.

Proposition 6.4 shows that for some n1βloga0tn_{1}\asymp_{\beta}\log\frac{\lVert a_{0}\rVert}{t} and some ρ[eC1n1,ec1n1]\rho\in[e^{-C_{1}n_{1}},e^{-c_{1}n_{1}}], there exist QρβQ\leq\rho^{-\beta} such that

μ(nn1)δy0(N~bd(ZQ,ρ))ρβ.\mu^{*(n-n_{1})}*\delta_{y_{0}}\bigl{(}\operatorname{\tilde{N}bd}(Z_{Q},\rho)\bigr{)}\geq\rho^{\beta}.

Applying Lemma 6.6 yields

ρα+β\displaystyle\rho^{-\alpha+\beta} φQ(gy0)dμ(nn1)(g)\displaystyle\leq\int\varphi_{Q}(gy_{0})\,\mathrm{d}\mu^{*(n-n_{1})}(g)
eλα(nn1)φQ(y0)+QC.\displaystyle\leq e^{-\lambda^{\prime}\alpha(n-n_{1})}\varphi_{Q}(y_{0})+Q^{C}.

Note that QCρCβ12ρα+βQ^{C}\leq\rho^{-C\beta}\leq\frac{1}{2}\rho^{-\alpha+\beta} and therefore

φQ(y0)=d~(y0,ZQ)αeλα(nn1)ρα+βeλαneC1α(11C+2)n1.\varphi_{Q}(y_{0})=\tilde{d}(y_{0},Z_{Q})^{-\alpha}\gg e^{\lambda^{\prime}\alpha(n-n_{1})}\rho^{-\alpha+\beta}\gg e^{\lambda^{\prime}\alpha n}e^{-C_{1}\alpha(1-\frac{1}{C+2})n_{1}}.

Since n1loga0tn_{1}\asymp\log\frac{\lVert a_{0}\rVert}{t} and λ=λ+1λ2\lambda^{\prime}=\lambda+\frac{1-\lambda}{2}, taking n11λloga0tn\gg\frac{1}{1-\lambda}\log\frac{\lVert a_{0}\rVert}{t} yields

d~(y0,ZQ)eλn\tilde{d}(y_{0},Z_{Q})\leq e^{-\lambda n}

and the theorem is proved. ∎

Acknowledgements

It is a pleasure to thank Yves Benoist and Elon Lindenstrauss for several useful discussions, in particular on the existence of satellite measures in the presence of compact factors. The authors are also grateful to the anonymous referee for numerous corrections and helpful comments. While this research was conducted, W.H. was supported by ERC 2020 grant HomDyn (grant no. 833423), KIAS Individual Grant (no. MG080401) and the National Natural Science Foundation of China (No. 12288201).

References

  • [1] R. Aoun. Random subgroups of linear groups are free. Duke Math. J., 160(1):117–173, 2011.
  • [2] T. Bénard. Equidistribution of mass for random processes on finite-volume spaces. Isr. J. Math., 255(1):417–422, 2023.
  • [3] Y. Benoist and J.-F. Quint. Mesures stationnaires et fermés invariants des espaces homogènes. Ann. of Math. (2), 174(2):1111–1162, 2011.
  • [4] Y. Benoist and J.-F. Quint. Random walks on finite volume homogeneous spaces. Invent. Math., 187(1):37–59, 2012.
  • [5] Y. Benoist and J.-F. Quint. Stationary measures and invariant subsets of homogeneous spaces (II). J. Amer. Math. Soc., 26(3):659–734, 2013.
  • [6] Y. Benoist and J.-F. Quint. Stationary measures and invariant subsets of homogeneous spaces (III). Ann. of Math. (2), 178(3):1017–1059, 2013.
  • [7] Y. Benoist and J.-F. Quint. Random walks on reductive groups, volume 62 of Ergebnisse der Mathematik und ihrer Grenzgebiete. 3. Folge. A Series of Modern Surveys in Mathematics. Springer, Cham, 2016.
  • [8] P. Bougerol and J. Lacroix. Products of random matrices with applications to Schrödinger operators, volume 8 of Progress in Probability and Statistics. Birkhäuser Boston, Inc., Boston, MA, 1985.
  • [9] J. Bourgain. On the Erdős-Volkmann and Katz-Tao ring conjectures. Geom. Funct. Anal., 13(2):334–365, 2003.
  • [10] J. Bourgain. The discretized sum-product and projection theorems. J. Anal. Math., 112:193–236, 2010.
  • [11] J. Bourgain, A. Furman, E. Lindenstrauss, and S. Mozes. Stationary measures and equidistribution for orbits of nonabelian semigroups on the torus. J. Amer. Math. Soc., 24(1):231–280, 2011.
  • [12] J. Bourgain and A. Gamburd. On the spectral gap for finitely-generated subgroups of SU(2)\text{SU}(2). Invent. Math., 171(1):83–121, 2008.
  • [13] J. Bourgain and A. Gamburd. Uniform expansion bounds for Cayley graphs of SL2(𝔽p){\rm SL}_{2}(\mathbb{F}_{p}). Ann. of Math. (2), 167(2):625–642, 2008.
  • [14] J. Bourgain, A. A. Glibichuk, and S. V. Konyagin. Estimates for the number of sums and products and for exponential sums in fields of prime order. J. Lond. Math. Soc., II. Ser., 73(2):380–398, 2006.
  • [15] J. Bourgain and P. P. Varjú. Expansion in SLd(𝐙/q𝐙),qSL_{d}({\bf Z}/q{\bf Z}),\,q arbitrary. Invent. Math., 188(1):151–173, 2012.
  • [16] J.-B. Boyer. On the affine random walk on the torus. arXiv e-prints, page arXiv:1702.08387, Feb 2017.
  • [17] T. Bénard and W. He. Multislicing and effective equidistribution for random walks on some homogeneous spaces, 2024. Preprint arXiv:2409.03300.
  • [18] P. Erdős and E. Szemerédi. On sums and products of integers. In P. Erdős, L. Alpár, G. Halász, and A. Sárközy, editors, Studies in Pure Mathematics: To the Memory of Paul Turán, pages 213–218. Birkhäuser Basel, 1983.
  • [19] A. Eskin and G. Margulis. Recurrence properties of random walks on finite volume homogeneous manifolds. In Random walks and geometry. Proceedings of a workshop at the Erwin Schrödinger Institute, Vienna, June 18 – July 13, 2001. In collaboration with Klaus Schmidt and Wolfgang Woess. Collected papers., pages 431–444. Berlin: de Gruyter, 2004.
  • [20] K. J. Falconer. Hausdorff dimension and the exceptional set of projections. Mathematika, 29:109–115, 1982.
  • [21] H. Furstenberg. Noncommuting random products. Trans. Amer. Math. Soc., 108:377–428, 1963.
  • [22] Y. Guivarc’h and A. N. Starkov. Orbits of linear group actions, random walks on homogeneous spaces and toral automorphisms. Ergodic Theory Dynam. Systems, 24(3):767–802, 2004.
  • [23] W. He. Discretized sum-product estimates in matrix algebras. J. Anal. Math., 139(2):637–676, 2019.
  • [24] W. He and N. de Saxcé. Sum-product for real Lie groups. J. Eur. Math. Soc. (JEMS), 23(6):2127–2151, 2021.
  • [25] W. He and N. De Saxcé. Linear random walks on the torus. Duke Math. J., 171(5):1061–1133, 2022.
  • [26] W. He and N. de Saxcé. Trou spectral dans les groupes simples. Enseign. Math., 70(3/4):425–477, 2024.
  • [27] W. He, T. Lakrec, and E. Lindenstrauss. Affine Random Walks on the Torus. International Mathematics Research Notices, 01 2021. rnaa322.
  • [28] W. He, T. Lakrec, and E. Lindenstrauss. Equidistribution of affine random walks on some nilmanifolds. In Analysis at large. Dedicated to the life and work of Jean Bourgain, pages 131–171. Cham: Springer, 2022.
  • [29] W. Kim. Effective equidistribution of expanding translates in the space of affine lattices, 2021. To appear in Duke Math. J.
  • [30] E. Le Page. Théorèmes limites pour les produits de matrices aléatoires. In Probability measures on groups (Oberwolfach, 1981), volume 928 of Lecture Notes in Math., pages 258–303. Springer, Berlin-New York, 1982.
  • [31] J. Li. Discretized sum-product and Fourier decay in n\mathbb{R}^{n}. J. Anal. Math., 143(2):763–800, 2021.
  • [32] E. Lindenstrauss and A. Mohammadi. Polynomial effective density in quotients of 3\mathbb{H}^{3} and 2×2\mathbb{H}^{2}\times\mathbb{H}^{2}. Invent. Math., 231(3):1141–1237, 2023.
  • [33] E. Lindenstrauss, A. Mohammadi, and Z. Wang. Effective equidistribution for some one parameter unipotent flows, 2022. Preprint arXiv:2211.11099.
  • [34] R. Muchnik. Semigroup actions on 𝕋n\mathbb{T}^{n}. Geom. Dedicata, 110:1–47, 2005.
  • [35] A. Salehi Golsefidy and P. P. Varjú. Expansion in perfect groups. Geom. Funct. Anal., 22(6):1832–1891, 2012.
  • [36] W. M. Schmidt. Diophantine approximation., volume 785. Springer, Cham, 1980.
  • [37] T. Tao. Product set estimates for non-commutative groups. Combinatorica, 28(5):547–594, 2008.
  • [38] T. Tao and V. H. Vu. Additive combinatorics, volume 105 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, 2010.
  • [39] B. L. Van der Waerden. Modern Algebra. Volume II. Based in part on lectures by E. Artin and E. Noether. Frederick Ungar Publishing Co., New York, transl. from the 3rd german edition, 1950.
  • [40] L. Yang. Effective version of Ratner’s equidistribution theorem for SL3()\mathrm{SL}_{3}(\mathbb{R}), 2022. Preprint arXiv:2208.02525.