Publication Cover
Dynamical Systems
An International Journal
Volume 39, 2024 - Issue 2
461
Views
0
CrossRef citations to date
0
Altmetric
Research Article

Aspects of convergence of random walks on finite volume homogeneous spaces

ORCID Icon
Pages 243-267 | Received 10 Feb 2023, Accepted 10 Oct 2023, Published online: 24 Oct 2023

Abstract

We investigate three aspects of weak* convergence of the n-step distributions of random walks on finite volume homogeneous spaces G/Γ of semisimple real Lie groups. First, we look into the obvious obstruction to the upgrade from Cesàro to non-averaged convergence: periodicity. We give examples where it occurs and conditions under which it does not. In a second part, we prove convergence towards Haar measure with exponential speed from almost every starting point. Finally, we establish a strong uniformity property for the Cesàro convergence towards Haar measure for uniquely ergodic random walks.

2010 Mathematics Subject Classifications:

1. Introduction

Let G be a real Lie group and Γ a lattice in G, that is, a discrete subgroup of G such that the homogeneous space X=G/Γ admits a G-invariant Borel probability measure mX. This measure mX is unique and we refer to it as the (normalized) Haar measure on X. A good example to have in mind is G=SLd(R) and Γ=SLd(Z).

The objects of study in this paper are random walks on X, given by probability measures µ on G: A step corresponds to randomly choosing a group element gG according to µ and then moving from the current location Xx to gx. Starting at x0X, the distribution of the location after n steps is given by the convolution (1) μnδx0,(1) which is the push-forward of the product measure μnδx0 under the multiplication map Gn×X(gn,,g1,x)gng1xX.

The broader context in which the study of these random walks originated is that of subgroup actions on homogeneous spaces. After Ratner's treatment of the rigidity and asymptotic properties of unipotent actions in her celebrated series of articles [Citation21–24], a new approach was needed to understand the dynamics of non-unipotent actions. Passing from a deterministic to a probabilistic point of view turned out to be a particularly fruitful angle. Still, understanding the long-term behaviour of random walks on homogeneous spaces and the limiting behaviour of the n-step distributions (Equation1) is a notoriously difficult problem. Major contributions to this line of study were made e.g. by Eskin–Margulis in their work on non-divergence [Citation15], and by Benoist–Quint in their breakthrough series of articles [Citation4,Citation6–8]. We reproduce one of the main results of [Citation8] as motivating example. For the statement, recall that a probability measure ν on X is called homogeneous if there exists a closed subgroup H of G and a point xX such that supp(ν)=Hx is a closed orbit and ν is H-invariant.

Theorem 1.1

Benoist–Quint [Citation8]

Let µ be a compactly supported probability measure on G. Denote by S and G the closed subsemigroup and subgroup of G generated by supp(μ), respectively, and suppose that the Zariski closure of Ad(G) in Aut(g) is Zariski connected, semisimple, and has no compact factors. Then for every x0X there is a homogeneous probability measure νx0 on X with supp(νx0)=Sx0¯=Gx0¯ and such that (2) 1nk=0n1μkδx0νx0(2) as n in the weak* topology.

Here the weak* convergence (Equation2) more explicitly means that for every compactly supported continuous function fCc(X) we have 1nk=0n1Xfd(μkδx0)=1nk=0n1Gkf(gkg1x0)dμk(g1,,gk)Xfdνx0as n. Recently, it was shown by Bénard–de Saxcé [Citation3] that the compact support assumption on µ in Theorem 1.1 can be relaxed to a finite first moment assumption; see Remark 2.7. Another recent generalization of the theorem above in joint work of the author with Sert and Shi [Citation19] replaces the algebraic assumption on the support of µ by a certain expansion condition, which allows for cases in which µ is e.g. supported on a parabolic subgroup of a semisimple group.

Some questions left open by Theorem 1.1 are listed by Benoist–Quint at the end of their survey [Citation5]. A major one is the following.

Question 1.2

In the setting of Theorem 1.1, is it also true that (3) μnδx0νx0(3) as n?

Answers are available only in special cases: Breuillard [Citation11] established (Equation3) for certain measures µ supported on unipotent subgroups, Buenger [Citation12] proved it for some sparse solvable measures, and in previous work the author dealt with the case of spread out measures [Citation18]. Very recently, Bénard [Citation2] observed that (Equation3) holds for aperiodic measures µ under the assumption that µ has two convolution powers which are not mutually singular.

The purpose of this article is to discuss three (largely independent) aspects of random walk convergence related to Theorem 1.1 and Question 1.2, mainly having in mind the case that G is a semisimple real Lie group. We are going to use the following terminology.

Definition 1.3

Let ν be a probability measure on X and x0X. We say that the random walk on X given by µ converges to ν on average (resp. converges to ν) from the starting point x0 if 1nk=0n1μkδx0ν (resp. μnδx0ν) as n in the weak* topology.

Convergence on average is also commonly referred to as Cesàro convergence. We use the two terms interchangeably.

The article is organized as follows.

In Section 2, we look into the obvious obstruction to the upgrade from Cesàro convergence to (non-averaged) convergence: periodicity. We show in Example 2.1 how (Equation3) can fail when x0 has finite orbit under S. Using a product construction, we can also produce a counterexample in which the orbit closure Sx0¯ has positive dimension (Example 2.2). In both cases, the periodic behaviour occurs at the level of the connected components of the orbit closure. As it turns out, this is no coincidence: If, in the setting of Theorem 1.1, the orbit closure Sx0¯ is connected, there can be no periodicity (Theorem 2.5) and we can show that the Cesàro convergence (Equation2) also holds along arithmetic progressions (Corollary 2.8).

In Section 3, we establish effective convergence of random walks to the normalized Haar measure mX for typical starting points x0: When supp(μ) generates a Zariski dense subgroup of a semisimple real Lie group G without compact factors, for any fixed L2-function f on X the convergence Xfd(μnδx0)nXfdmXnot only holds but is in fact exponentially fast for mX-almost every x0X (Theorem 3.2, Proposition 3.4). The proof relies on an L2-spectral gap of the convolution operator π(μ):f(xGf(gx)dμ(g))acting on measurable functions on X. Taking into account regularity of the function f, the above can be further strengthened to the statement that almost every xX is exponentially generic (Definition 3.12): Up to a constant factor depending on derivatives of f, the exponential speed of convergence holds uniformly over all compactly supported smooth functions (Theorem 3.13). Key to this upgrade are the definition of suitable Sobolev norms and a functional analytic argument involving relative traces, first exploited in a dynamical context by Einsiedler–Margulis–Venkatesh [Citation13].

Finally, in Section 4 we prove that convergence on average to mX happens locally uniformly in x0 in a strong way when the random walk is uniquely ergodic and admits a Lyapunov function (Theorem 4.13). For example, this is the case when G is a connected semisimple real algebraic group and supp(μ) generates a non-discrete Zariski dense subgroup, and also in the setup of Simmons–Weiss [Citation27], which has connections to Diophantine approximation problems on fractals. To this end, we introduce the new concept of (Kn)n-uniform recurrence (Definition 4.10), which refines recurrence properties of random walks previously studied in [Citation6,Citation15].

1.1. Standing assumptions & notation

As many of our arguments work in greater generality, in the remainder of the article we will relax the assumptions stated at the beginning of this introduction. The following setup shall be in place whenever nothing else is specified: G is a locally compact σ-compact metrizable group acting ergodically on a locally compact σ-compact metrizable space X endowed with a G-invariant probability measure mX; and µ is a Borel probability measure on G.

2. Periodicity

In this section, we start with two simple counterexamples to (Equation3), which illustrate ways in which a random walk may exhibit periodic behaviour (Section 2.1). Analysing these examples for their common feature, we are led to a simple condition ensuring aperiodicity, stated and proved in Section 2.2.

2.1. Examples

The first example with periodicity is on finite periodic orbits. In the following, for d2 we denote by 1d the d×d-identity matrix.

Example 2.1

Consider the principal congruence lattice Γ=Γ(2)={gSL2(Z)g12mod2}in G=SL2(R). Being the kernel of the reduction homomorphism from SL2(Z) to SL2(Z/2Z), we recognize Γ(2) as a finite-index normal subgroup of SL2(Z). In particular, Γ(2) is a lattice in G. Let μ=12(δh1+δh2) with h1=(1101),h2=(1011).Then the closed subgroup G generated by supp(μ)={h1,h2} is G=SL2(Z), which is Zariski dense in G. The G-orbit of x0=12ΓG/Γ is O={x0,h1x0,h2x0,h2h1x0=(1112)x0,h1h2x0=(2111)x0,h1h2h1x0=h2h1h2x0=(2110)x0},with transitions as shown in the following diagram:

Consequently, we see that the random walk with starting point x0 alternates between the two sets O1={x0,h1h2x0,h2h1x0}andO2={h1x0,h2x0,h1h2h1x0}.The 2-step random walks on these sets constitute irreducible, aperiodic, finite state Markov chains, so that μ2nδx013pO1δp,μ(2n+1)δx013pO2δp,as n in the weak* topology.

In the example above, the support of µ generates a Zariski dense subgroup of G and the lattice Γ in G is irreducible. (Recall that, loosely speaking, ‘irreducibility’ of Γ means that it does not arise from a product construction, cf. [Citation20, Definition 5.20]). By the work of Benoist–Quint [Citation8, Corollary 1.8], these properties force any orbit closure Sx0¯ to be either finite or all of X. As soon as intermediate orbit closures are possible, however, one can also construct examples with periodic behaviour on non-discrete orbit closures.

Example 2.2

Let G, Γ, X=G/Γ, h1,h2, x0 and G be as in Example 2.1 and choose a diagonal matrix aSL2(R) such that the diagonal entries of a2 are irrational. We are going to consider the random walk on the product space X×X=(G×G)/(Γ×Γ)given by the probability measure μ=14i=14δgi on G×G with g1=(h1,ah1a1),g2=(h1,12),g3=(h2,ah2a1),g4=(h2,12).The (closed) subgroup generated by the support of this measure µ is given by G×aGa1=SL2(Z)×aSL2(Z)a1. Indeed, the correct entry in the second copy of G can be arranged using a finite product of g1±1,g3±1, and then the entry in the first copy can be corrected using g2±1,g4±1. By Theorem 1.1 we thus know that for the starting point (x0,x0)X×X we have the weak* convergence 1nk=0n1μkδ(x0,x0)ν(x0,x0)as n, where ν(x0,x0) is the homogeneous probability measure on the closure of the G×aGa1-orbit of (x0,x0). (Recall that it makes no difference for the closure whether one considers the orbit under the generated subgroup or subsemigroup.)

Let us identify this orbit closure. In the first copy of X, we recognize the finite orbit O from Example 2.1. In the second copy, we see the action of irrational conjugates of h1,h2. As the acting group has product structure, the orbit closure in question is the product of these two orbit closures in the components: (G×aGa1)(x0,x0)¯=O×aGa1x0¯.Since the orbit aGa1x0 is infinite by our choice of the matrix a, it follows from [Citation8, Corollary 1.8] that aGa1x0¯=X, so that (G×aGa1)(x0,x0)¯=O×Xandν(x0,x0)=mOmXfor the normalized counting measure mO on O and the normalized Haar measure mX on X. However, in analogy to Example 2.1, the random walk is found to alternate between the sets O1×XandO2×X,in the sense that supp(μ2nδ(x0,x0))O1×X and supp(μ(2n+1)δ(x0,x0))O2×X for all nN. Hence, we conclude that the random walk starting from (x0,x0) does not converge to ν(x0,x0).

Remark 2.3

The same behaviour as in the previous example can be arranged inside a homogeneous space X=G/Γ that is the quotient of a semisimple real Lie group G by an irreducible lattice Γ. Indeed, this is only a matter of choosing suitable embeddings G×GG and X×XX, where G and X are as in Example 2.2. Concretely, one can e.g. consider the 4×4-congruence lattice Γ=Γ(2)={gSL4(Z)g14mod2}in G=SL4(R) and the diagonal embeddings G×GG,X×XX,(g,h)(gh),(gΓ,hΓ)(gh)Γ.We therefore see that Example 2.2, i.e. periodic behaviour on a non-discrete orbit closure, can be realized inside X=G/Γ. Of course, after applying this embedding, the subgroup generated by the support of µ will no longer be Zariski dense in G.

2.2. An aperiodicity criterion

Inspecting the examples above, one may notice that their common salient feature is that the orbit closure Sx0¯ is disconnected. This naturally raises the question whether periodic behaviour can also occur when this orbit closure is connected. In what follows, we answer this question in the negative. We shall use the following formalization of periodicity.

Definition 2.4

Assume that the random walk on X given by µ converges on average to a probability measure ν on X from the starting point x0X. We say that this convergence is periodic if there exists an integer d2 and pairwise disjoint measurable subsets D0,,Dd1X with ν(Di)=0 for 0i<d and such that (μnδx0)(Dnmodd)=1 for every nN. Otherwise, we call the convergence aperiodic.

The requirement on the boundaries of the sets Di is needed to ensure that the cyclic behaviour is witnessed by the limit measure ν. Without a condition of this sort, one could try to artificially define Di as the set of all points in X that can be reached from x0 precisely in nimodd steps. Indeed, this construction is possible for example when µ is finitely supported with the property that its support freely generates a discrete subsemigroup S of G and the starting point x0X has a free S-orbit. The latter is the case e.g. for X=SL2(R)/SL2(Z), μ=12(δh1+δh2) with h1=(1201) and h2=(1021), and x0=aSL2(Z) for a diagonal matrix aSL2(R) such that the diagonal entries of a2 are irrational.

We are now ready to state the announced aperiodicity theorem.

Theorem 2.5

Retain the notation and assumptions from Theorem 1.1 and let x0X be such that the orbit closure Sx0¯ is connected. Then the Cesàro convergence to νx0 of the random walk on X given by µ starting from x0 is aperiodic.

For the proof we need the following simple lemma.

Lemma 2.6

Let H be a Zariski connected real algebraic group and S a subset of H generating a Zariski dense subsemigroup. Then for every dN, also the d-fold product set Sd={gdg1g1,,gdS} generates a Zariski dense subsemigroup of H. In particular, if supp(μ) generates a Zariski dense subsemigroup for some probability measure µ on H, the same is true for supp(μd).

Proof.

Let UH be a non-empty Zariski open subset and consider the map ϕ:HH,ggd. Since ϕ is Zariski continuous, ϕ1(U) is Zariski open. Moreover, this preimage is non-empty because U is dense in the Lie group topology and ϕ is a diffeomorphism near the identity. By the assumption that S generates a Zariski dense subsemigroup, we can thus find an element gϕ1(U) that is the product of finitely many elements of S. It follows that ϕ(g)=gd lies in the intersection of U with the subsemigroup generated by Sd.

The second claim involving µ immediately follows from the above together with the inclusion supp(μd)supp(μ)d.

Proof

Proof of Theorem 2.5

Suppose dN is an integer such that there are pairwise disjoint D0,,Dd1X with νx0(Di)=0 for all 0i<d and such that (μnδx0)(Dnmodd)=1 for all nN as in the definition of periodicity. We have to show that d = 1.

First note that from Theorem 1.1 and the properties of the sets Di it follows that (4) νx0(D0)=limn1nk=0n1(μkδx0)(D0)=1d,(4) where the application of weak* convergence to the set D0 is justified since it has negligible boundary with respect to the limit measure νx0. In view of Lemma 2.6, Theorem 1.1 also applies to the d-step random walk given by μd. Assuming for the moment that the limit measure for this d-step random walk starting from x0 coincides with νx0, we deduce that (5) νx0(D0)=limn1nk=0n1(μdkδx0)(D0)=1.(5) Together, (Equation4) and (Equation5) imply d = 1, the desired conclusion.

It thus remains to show that the d-step random walk starting from x0 does indeed have the same limit measure as the 1-step random walk. Denoting by S and Sd the closed subsemigroups of G generated by supp(μ) and supp(μd), respectively, this statement is equivalent to the equality Sx0¯=Sdx0¯ of orbit closures. To prove this, let gsupp(μ) be arbitrary. We claim that Sx0¯=k=0d1gkSdx0¯.Indeed, since Sx0¯ is homogeneous, it is invariant under the group generated by S. As Sx0¯ clearly contains Sdx0¯, the inclusion ‘’ follows. For the reverse inclusion let gn,,g1supp(μ) for some nN. Choose 0k<d such that n+k0modd. Then gkgng1x0Sdx0¯ and hence gng1x0gkSdx0¯, giving the claim.

We already noted that Theorem 1.1 applies to μd. In particular, the orbit closure Sdx0¯ and its translates by gk, 0k<d, are submanifolds of Sx0¯. Necessarily, all these translates have the same dimension, and since together they make up Sx0¯ by the claim above, their shared dimension coincides with that of Sx0¯. This implies that Sdx0¯ is open in Sx0¯. However, it is also closed, so that the assumed connectedness of Sx0¯ forces Sx0¯=Sdx0¯. This completes the proof.

Remark 2.7

It was recently shown by Bénard–de Saxcé [Citation3] that the compact support assumption on µ in Theorem 1.1 can be relaxed. Indeed, their [Citation3, Theorem C] establishes the same conclusion under the substantially weaker assumption that µ has a finite first moment, meaning that GlogAd(g)dμ(g)<.Relying on this stronger result, also our Theorem 2.5 above and Corollary 2.8 below are seen to hold under a finite first moment assumption on µ, instead of requiring compact support as in Theorem 1.1.

We end this section by recording a corollary of the proof above.

Corollary 2.8

Retain the notation and assumptions from Theorem 1.1 and suppose that Sx0¯ is connected. Let dN and denote by Sd the closed subsemigroup of G generated by supp(μd). Then Sx0¯=Sdx0¯, and for the homogeneous probability measure νx0 on this orbit closure we have for arbitrary rN0 that (6) 1nk=0n1μ(dk+r)δx0νx0(6) as n in the weak* topology.

Proof.

The statement about orbit closures was established as part of the proof of Theorem 2.5. From Theorem 1.1 we thus get the weak* convergence (7) 1nk=0n1μdkδx0nνx0,(7) which is (Equation6) for r = 0. Given fCc(X), the general case follows by applying (Equation7) to the compactly supported continuous function fr defined by fr(x):=Gf(gx)dμr(g)=Grf(grg1x)dμr(g1,,gr)for xX.

This corollary sharpens the convergence statement in Theorem 1.1 in the case of a connected orbit closure: The Cesàro convergence to νx0 holds along arbitrary arithmetic progressions. Although this does not provide an answer to Question 1.2, it at least allows the following conclusion to be drawn: If (ni)i is a sequence of indices such that μniδx0 converges to a weak* limit different from νx0 as i, then (ni)i cannot contain a density 1 subset of an infinite arithmetic progression.

3. Spectral gap

In this section, we will explain how a spectral gap of the convolution operator π(μ) associated to a random walk entails the convergence of μnδx towards mX for mX-a.e. xX. In its simplest form, the involved argument works in great generality and also produces an exponential rate of convergence from almost every starting point when the test function f is fixed. This is done in Section 3.1. The following Sections 3.23.4 are dedicated to a substantial refinement of this spectral gap argument for random walks on homogeneous spaces of real Lie groups, making the exponentially fast convergence uniform over smooth test functions.

3.1. Generic points

Recall that π(μ):L(X,mX)L(X,mX) is defined by π(μ)f(x):=Xfd(μδx)=Gf(gx)dμ(g)for fL(X,mX) and xX, and that it extends to a continuous contraction on each Lp-space (see [Citation9, Corollary 2.2]). We shall study its behaviour on L2(X,mX). By ergodicity, the G-fixed functions are the constant functions, so we restrict our attention to their orthogonal complement L02(X,mX) of L2-functions with mean 0.

Definition 3.1

We say that µ has a spectral gap on X if the associated convolution operator π(μ) restricted to L02(X,mX) has spectral radius strictly less than 1.

We are going to use the notation ρ(T) to denote the spectral radius of an operator T. Then by the spectral radius formula, µ having a spectral gap on X can be reformulated as the requirement that ρ(π(μ)|L02)=limnπ(μ)|L02nopn<1.Given the existence of a spectral gap, we obtain an almost everywhere convergence result in a quite general setup.

Theorem 3.2

Suppose that µ has a spectral gap on X. Then mX-a.e. xX is generic for the random walk on X given by µ, meaning that μnδxmXas n in the weak* topology. This convergence is exponentially fast in the sense that for every fixed fL2(X,mX) we have (8) lim supn|Xfd(μnδx)fdmX|1/nρ(π(μ)|L02)(8) for mX-a.e. xX.

Proof.

By separability of Cc(X), for the statement about weak* convergence it suffices to prove mX-a.s. convergence for one fixed function fCc(X). Consequently, it is enough to prove the second assertion of the theorem. To this end, fix a function fL2(X,mX) and a rational number ρ(π(μ)|L02)<α<1, and consider the L02-function f0=ffdmX. Then in view of the spectral radius formula we have π(μ)nffdmXL2=π(μ)nf0L2π(μ)|L02nopf0L2αnf0L2for sufficiently large nN.

Fix in addition a rational number ε(0,1). By Chebyshev's inequality, the above implies that for large n we have mX({xX||π(μ)nf(x)fdmX|αn(1ε)f0L2})π(μ)nffdmXL22α2n(1ε)f0L22α2εn.By Borel–Cantelli it follows that for all x in a full measure set Aα,ε, the inequality |π(μ)nf(x)fdmX|αn(1ε)f0L2holds only for finitely many nN. Since π(μ)nf(x)=fd(μnδx), we conclude that (Equation8) holds for all x in a countable intersection of the sets Aα,ε over rational numbers α approaching ρ(π(μ)|L02) and ε approaching 0 from above.

Remark 3.3

In the second conclusion of Theorem 3.2, how long it takes for the exponential rate of convergence to kick in depends on the point x. However, the measure of sets on which one has to wait for a long time can be controlled as follows: Given ρ(π(μ)|L02)<α<1, choose NN such that π(μ)|L02nopαn for all nN. Then if we additionally take ε(0,1) and denote Bα,ε,n,f={xX||π(μ)nf(x)fdmX|αn(1ε)f0L2forsomenn},the proof above gives the bound mX(Bα,ε,n,f)α2εn1α2εfor every nN. In particular, the measure of the set on which the exponential convergence does not start during the first n steps decays exponentially in n.

We now demonstrate that the previous result covers the case announced in Section 1.

Proposition 3.4

Let G be a connected semisimple real Lie group without compact factors and with finite centre, ΓG a lattice, and X the homogeneous space G/Γ endowed with the Haar measure mX. Suppose that the closed subsemigroup S generated by supp(μ) has the property that Ad(S) is Zariski dense in Ad(G). Then µ has a spectral gap on X.

Proof.

Consider the regular representation of G on L02(X,mX). By Bekka [Citation1, Lemma 3] it doesn't weakly contain the trivial representation. From this, in view of [Citation25, Theorem C], the result follows if we can argue that the projection of µ to any simple factor of G is not supported on a closed amenable subgroup. However, since amenability passes to the Zariski closure (see e.g. [Citation28, Theorem 4.1.15]) the latter would imply that one of the simple factors of Ad(G) is amenable, hence compact by a classical result of Furstenberg (see e.g. [Citation28, Proposition 4.1.8]).

3.2. Good height functions

Inspecting the proof of Theorem 3.2, one observes that every step is effective, with explicit bounds and good control over the measure of exceptional sets, except for the very first one: separability of the space Cc(X) of compactly supported continuous functions. In the remainder of this section, we aim to also make effective this step, the goal being exponentially fast convergence μnδxmX from almost every starting point, uniformly over functions f on X. As merely continuous functions can behave arbitrarily badly (with respect to the convergence problem at hand), there is no hope of achieving this feat for all fCc(X). We shall therefore restrict our attention to smooth functions of compact support, and take into account their regularity by considering not just their L2, but also certain Sobolev norms. Built into the definition of these norms will be what we call a good height function, the concept of which is introduced in this subsection.

Our setup is as follows: Let G be a real Lie group with Lie algebra g. We endow g with a scalar product, which we use to define a right-invariant metric dG on G. Given a lattice ΓG, this metric descends to a metric dX on X=G/Γ such that the projection GX is locally an isometry. Moreover, we fix an orthonormal basis of g, using which we will identify g with Rdimg. Here is the crucial definition.

Definition 3.5

We call a measurable function ht:X(0,) a good height function if there exists 0<R1 and a function r:X(0,R] with the following properties:

  1. The restriction of the exponential map exp:(R,R)dimgG is a diffeomorphism onto its image and we have exp((r/2,r/2)dimg)BrG(e) for all rR, where BrG(e) denotes the open ball of radius r around the identity eG with respect to the metric dG on G.

  2. For all xX, the projection GBr(x)G(e)X,ggx is injective.

  3. There exist constants c,κ>0 such that r(x)cht(x)κ for all xX.

  4. There exists a constant σ>1 such that ht(x)σht(gx) for all xX and all gBr(x)G(e).

The definition suggests to think of a good height function as reciprocal of the injectivity radius. And indeed, this viewpoint allows their construction on any homogeneous space X=G/Γ.

Proposition 3.6

Let G be a real Lie group and Γ a lattice in G. Then X=G/Γ admits a good height function.

Proof.

Choose R>0 such that condition (i) of the definition is satisfied and set r(x)=min{R,rinj(x)}, where rinj(x) is the injectivity radius at xX, i.e. the maximal radius such that (ii) holds at x. Define ht(x)=r(x)1.Then the only thing that needs to be verified is the validity of (iv). We claim that it holds with σ=2. This will follow if we can show that (9) rinj(gx)2rinj(x)(9) whenever gBr(x)G(e). To this end, let r>rinj(x). Then by definition, there are distinct g1,g2BrG(e) such that g1x=g2x. As gBr(x)G(e), right-invariance of the metric implies dG(gig1,e)=dG(gi,g)dG(gi,e)+dG(g,e)<r+r(x)<2rfor i = 1, 2, and we also have (g1g1)gx=(g2g1)gx. This shows that rinj(gx)2r, and as r>rinj(x) was arbitrary, we see that (Equation9) holds.

Often, however, one might want to work with different, naturally occurring height functions. The flexibility in our definition of a good height function accommodates this possibility.

In the examples below, we denote by λ1(Λ) the length of a shortest non-zero vector in a lattice ΛRd.

Example 3.7

Let G=SLd(R) and Γ=SLd(Z). Then X=G/Γ can be identified with the space of lattices in Rd with covolume 1 via XgSLd(Z)gZdRd.Then the function ht=λ11, defined on X via the above identification, is a good height function. Indeed, one can first choose R>0 such that (i) is satisfied, and then set r(x)=min{R,rinj(x)} as in the proof of Proposition 3.6. Then (ii) is automatically satisfied, and (iv) is valid for a suitable choice of σ due to the inequality λ1(gx)gλ1(x) for gG and xX, where denotes any matrix norm. To see that also (iii) holds, let x=gΓ and suppose that hx = x for some hG with he. Then for all γSLd(Z), the matrix ()1h() fixes the lattice Zd but is not the identity, so that κ1he()1(he)()=()1h()ec1for some constants c1,κ1>0. For a basis change γSLd(Z) such that consists of a reduced basis of the lattice x we have c2λ1(x)κ2 for some c2,κ2>0 (cf. e.g. [Citation26, Chapter III]). With this choice, the above inequality implies hecλ1(x)κfor c=c1/c2 and κ=κ1κ2. Since near the identity, the metric dG on G is Lipschitz-equivalent to the distance induced by , this establishes (iii).

A similar construction is possible in a more general context.

Example 3.8

[Citation13]

Let G=G(R) be the group of real points of a semisimple Q-group G and Γ an arithmetic lattice in G. Choose a rational Ad(Γ)-stable lattice gZg. Then, using similar reasoning as in the previous example, the function ht on X=G/Γ defined by ht(x)=λ1(Ad(g)gZ)1for x=gΓX is seen to be a good height function (cf. [Citation13, Section 3.6]).

3.3. Sobolev norms

Given a good height function ht on X, the associated Sobolev norm of degree 0 of a compactly supported smooth function fCc(X) is defined by S(f)2=degDht()DfL22,where the sum runs over differential operators D given by monomials of degree at most ℓ in elements of the fixed orthonormal basis of g in the universal enveloping algebra.

In other words, the differential operators D appearing above are v1vk for any k-tuple (v1,,vk) of elements of the fixed basis of g, 0k, where v for vg is defined by vf(x)=limt0f(exp(tv)x)f(x)tfor fCc(X) and xX.

Here are two immediate observations.

Lemma 3.9

Let ht be a good height function on X and S the associated Sobolev norms.

  1. The norms S are induced by inner products , on Cc(X).

  2. Given 001, there exists a constant c~>0 such that S0c~S1.

Proof.

Part (i) is clear. Part (ii) is also immediate from the definition of the Sobolev norms, once we know that a good height function must be bounded away from 0. The latter, however, follows directly from property (iii) in the definition of a good height function, as the function r appearing there is assumed to be bounded.

The proof of our convergence result in Section 3.4 will depend on the following proposition.

Proposition 3.10

[Citation13]

For the Sobolev norms associated to a good height function on X, there exists a non-negative integer 00 and a constant C>0 with the following properties:

  1. (Sobolev embedding estimate [Citation13, (3.9)]) For every fCc(X) it holds that fCS0(f).

  2. (Finite relative traces [Citation13, (3.10)]) For all integers 0 the relative trace Tr(S2|S+02) is finite, meaning that for any orthogonal basis (e(k))k in the completion of Cc(X) with respect to S+0 Tr(S2|S+02):=kS(e(k))2S+0(e(k))2<.

We refer to Bernstein–Reznikov [Citation10] for a systematic treatment of relative traces. In particular, it is proved in this reference that the above expression is independent of the choice of orthogonal basis.

The proofs in [Citation13] of the statements in the above proposition are given for the height function from Example 3.8. However, the only properties used are those in our definition of a good height function. In fact, the arguments only depend on validity of the second statement in [Citation13, Lemma 5.1], which holds in our context, as we demonstrate below.

Lemma 3.11

Let ht be a good height function on X. Then there exists a non-negative integer 00 and a constant C>0 such that for every non-negative integer 0 and every differential operator D given by a monomial of degree at most ℓ in elements of the fixed basis of g we have |ht(x)Df(x)|CS+0(f)for every fCc(X) and xX.

Proof.

We inspect the function F=Df in a chart around x given by the exponential map: We set ε=r(x)/2, where r:X(0,R] is the function from the definition of a good height function, d=dimg, and consider F~:(ε,ε)dR,vF(exp(v)x).Then by the first statement of [Citation13, Lemma 5.1], which is simply a Sobolev embedding estimate on Rd, we know (10) |F(x)|=|F~(0)|C12dr(x)dSd,ε(F~),(10) where C1>0 is a constant depending only on the dimension d of g and Sd,ε is the standard degree d Sobolev norm on the open subset (ε,ε)d of Rd, i.e. Sd,ε(F~)2=|α|dαF~L2((ε,ε)d)2,where the sum is over all multi-indices α of degree at most d and αF~ is the corresponding standard partial derivative of F~. Using property (iii) in the definition of a good height function, (Equation10) implies that (11) |ht(x)F(x)|C2ht(x)+0Sd,ε(F~),(11) where C2>0 is another constant and we used that ht is bounded away from 0 to replace κd appearing in the exponent by 0=max{κd,d}. Using properties (i) and (ii) in the definition of a good height function, we find C3>0 such that (12) Sd,ε(F~)C3degDdDF|Br(x)X(x)L22.(12) To see this, one needs to note two things: firstly, that by the chain rule the partial derivatives of F~ at a point v(ε,ε)d in the chart can be expressed as linear combinations of derivatives DF appearing on the right-hand side in (Equation12) evaluated at the corresponding point x=exp(v)x, with fixed coefficient functions depending only on finitely many derivatives of the exponential map on (ε,ε)d; and secondly, that the Haar measure mX is a smooth measure, meaning that it has a smooth and nowhere vanishing density w.r.t. Lebesgue measure in the chart.

Combining (Equation11), (Equation12), condition (iv) in the definition of a good height function, and plugging back in the definition of F, we finally arrive at |ht(x)Df(x)|C4degDdht()+0DDf|Br(x)X(x)L22C4S+0(f),for yet another constant C4>0, which is the one appearing in the lemma.

3.4. Exponentially generic points

Now we are ready to define the notion of effective genericity we wish to establish, and to prove the main convergence result of this section.

Until the end of this section, we fix a good height function ht on X. Moreover, given a bounded measurable function f on X and nN we will use the notation Dn(f)(x)=π(μ)nf(x)fdmXfor xX. We refer to Dn(f) as the time n discrepancy for the function f.

Definition 3.12

We say that a point xX is (,β)-exponentially generic if 0 is a non-negative integer and β a real number in (0,1) satisfying lim supnsupfCc(X){0}(|Dn(f)(x)|S(f))1/nβ,where S is the degree ℓ Sobolev norm associated to ht.

With this terminology, we have the following result, which quantifies the dependence on the function f in the effective part of Theorem 3.2.

Theorem 3.13

Let G be a real Lie group, ΓG a lattice and X=G/Γ endowed with the Haar measure mX. Suppose that µ has a spectral gap on X. Then there exists a non-negative integer 10 such that mX-almost every point xX is (1,ρ(π(μ)|L02))-exponentially generic.

Our argument uses ideas from the proof of [Citation13, Proposition 9.2]. Recall that , denotes the inner product associated to the Sobolev norm S.

Proof.

Set 1=20 with 0 from Proposition 3.10. We denote by H the completion of Cc(X) with respect to S1.

The first step of the proof is to argue that H admits an orthonormal basis (e(k))k with respect to S1 that is also orthogonal with respect to S0. To this end, let us endow H with the scalar product ,1 associated to S1. This makes H into a Hilbert space. As a consequence of Lemma 3.9(ii), ,0 defines a bounded positive definite Hermitian form on (H,,1). Using Riesz representation it follows that there is a bounded positive self-adjoint operator T on (H,,1) such that v,w0=Tv,w1for all v,wH. Finiteness of the relative trace Tr(S02|S12) from Proposition 3.10(ii) then translates into the statement that T is a trace-class operator on (H,,1) (cf. [Citation14, Proposition 6.44]); in particular, the operator T is compact (cf. [Citation14, Proposition 6.42]). By the spectral theorem, T is thus diagonalizable. Hence, an orthonormal basis (e(k))k of (H,,1) consisting of eigenvectors of T is a basis with the desired properties.

Next, fix rational numbers ρ(π(μ)|L02)<α<1 and ε(0,1). As in the proof of Theorem 3.2, using Chebyshev's inequality we find that for every k0 and large enough n we have (13) mX({xX||Dn(e(k))(x)|αn(1ε)S0(e(k))})e0(k)L22S0(e(k))2α2εne(k)L22S0(e(k))2α2εn,(13) where e0(k)=e(k)e(k)dmX. Since the relative trace Tr(S02|S02) is finite by Proposition 3.10, the terms on the right-hand side of (Equation13) are summable over k,n0. Borel–Cantelli thus implies that lim supk,n0{xX||Dn(e(k))(x)|αn(1ε)S0(e(k))}is a null set. Let Aα,ε be the complement of this null set. We claim that any xAα,ε is (1,α1ε)-exponentially generic. Fix such a point x. Then we know that there are only finitely many pairs (k,n) with |Dn(e(k))(x)|αn(1ε)S0(e(k)). Thus, there exists n0 such that for nn0 the inequality |Dn(e(k))(x)|<αn(1ε)S0(e(k)) holds for all k. Now let fCc(X){0} be arbitrary and write f=kfke(k) for the expansion of f in terms of the orthonormal basis (e(k))k. Then, using the triangle inequality, we can estimate the time n discrepancy for f as follows: (14) |Dn(f)(x)|k|fk||Dn(e(k))(x)|.(14) The exchange of integral and summation involved in the above estimate is justified by part (i) of Proposition 3.10: It ensures that the functions e(k) are defined pointwise and the series expansion of f converges uniformly. Next, for nn0 an application of the Cauchy–Schwarz inequality implies that the right-hand side of (Equation14) is strictly less than (15) αn(1ε)(k|fk|2)1/2(kS0(e(k))2)1/2=αn(1ε)S1(f)Tr(S02|S12)1/2.(15) Again by Proposition 3.10, the relative trace Tr(S02|S12) is finite. Hence, in view of our definition of exponential genericity and the fact that n0 does not depend on f, combining (Equation14) and (Equation15) establishes the claim. It follows that all x in a countable intersection of the sets Aα,ε over rational numbers α approaching ρ(π(μ)|L02) and ε approaching 0 from above are (1,ρ(π(μ)|L02))-exponentially generic, giving the theorem.

Remark 3.14

In analogy to Remark 3.3, we can control the measure of the set of points where exponentially generic behaviour is not observed for a given number of steps: If we define Bα,ε,n={xX||Dn(f)(x)|αn(1ε)S1(f)Tr(S02|S12)1/2forsomenn,fCc(X)}for ρ(π(μ)|L02)<α<1, ε(0,1) and nN, and NN is chosen such that π(μ)|L02nopαn for all nN, then for every nN it holds that mX(Bα,ε,n)Tr(S02|S02)α2εn1α2ε.Indeed, we have Bα,ε,nnn,k0{xX||Dn(e(k))(x)|αn(1ε)S0(e(k))}, as the proof of Theorem 3.13 demonstrates. Thus, again, the measure of the set of ‘bad points’, on which exponential genericity takes more than n steps to manifest, is itself exponentially small in n.

4. Uniform Cesàro convergence

In this last section, we explore the situation where the only possible limit in Theorem 1.1 is the normalized Haar measure mX. In this setting, by analogy with the case of unique ergodicity in classical ergodic theory, it is reasonable to expect the Cesàro convergence (Equation2) to hold (locally) uniformly in the starting point x0. We shall prove in Section 4.1 below that this indeed holds true. In Section 4.2, we conclude the article by showing that in many naturally occurring situations something even stronger than locally uniform can be achieved.

Before continuing with the pertinent definitions, let us recall that even though the setup of Theorem 1.1 is our motivation and useful to have in mind, formally we are working with the assumptions stated at the end of Section 1: (X,mX) is merely required to be a space with a G-action for which mX is invariant and ergodic.

Definition 4.1

A probability measure ν on X is called µ-stationary if μν=ν. The random walk on X induced by µ is called uniquely ergodic if mX is the unique µ-stationary probability measure on X.

In particular, for a random walk to be uniquely ergodic, there must be no finite G-orbits in X, where G denotes the closed subgroup of G generated by µ. In the case that X=G/Γ for a lattice Γ in G, this happens if and only if G is not virtually contained in a conjugate of Γ. (Recall that a subgroup H of G is said to be virtually contained in a subgroup L of G if HL has finite index in H.) In fact, in many cases of interest, finite orbits are the only obstruction to unique ergodicity: For example, this is true when G is a connected semisimple Lie group without compact factors, Γ is an irreducible lattice, X=G/Γ, and Ad(S) is Zariski dense in Ad(G) (see [Citation8, Corollary 1.8]); and also in the setting of [Citation27], a special case of which is reproduced below as Example 4.8.

4.1. Locally uniform convergence

The notion of unique ergodicity introduced above coincides with the classical property of unique ergodicity of the Markov operator π(μ). When the space X is compact, this is enough to guarantee that the Cesàro convergence 1nk=0n1μkδxmX as n is uniform in x (see e.g. [Citation16, Section 5.1]). Without compactness, we also need to assume a form of recurrence.

Definition 4.2

We say that the random walk on X given by µ is locally uniformly recurrent if for every compact subset KX and ε>0 there exists n0N and a compact subset MX with μnδx(M)1εfor all nn0 and xK. It is called locally uniformly recurrent on average if the above holds with the Cesàro averages 1nk=0n1μkδx in place of μnδx.

It is a simple exercise to check that locally uniform recurrence implies locally uniform recurrence on average. In concrete examples, recurrence properties such as these are typically established by constructing a Lyapunov function; see Section 4.2 below.

The following well-known fact explains why these properties are referred to as ‘non-escape of mass’.

Lemma 4.3

Let the sequence {xn}n of points in X be relatively compact and suppose that the random walk on X is locally uniformly recurrent (resp. on average). Then every weak* limit of the sequence (μnδxn)n (resp. (1nk=0n1μkδxn)n) is a probability measure.  □

The proof is immediate and left to the reader.

We are now ready to state and prove our first result on locally uniform Cesàro convergence.

Theorem 4.4

Suppose that the random walk on X induced by µ is uniquely ergodic and locally uniformly recurrent on average. Then for every fCc(X), every compact KX, and every ε>0, there exists n0N such that for every nn0 and xK we have |1nk=0n1Xfd(μkδx)XfdmX|<ε.Equivalently, considering the space of probability measures on X as endowed with the weak* topology, the sequence of functions Xx1nk=0n1μkδxconverges to mX uniformly on compact subsets of X as n.

Proof.

The equivalence of the two formulations is due to the definition of neighbourhoods in the weak* topology by finitely many test functions in Cc(X).

To prove the statement for individual functions, we proceed by contradiction. If the conclusion is false, then for some fCc(X), KX compact and ε>0 there exist indices n(j) and xjK with (16) |1n(j)k=0n(j)1Xfd(μkδxj)XfdmX|ε(16) for all jN. Let ν be a weak* limit point of the sequence (1n(j)k=0n(j)1μkδxj)j.Then ν is µ-stationary, and a probability measure because of our recurrence assumption and the fact that all xj lie in the fixed compact set K Lemma 4.3. But by unique ergodicity this forces ν=mX, contradicting (Equation16).

4.2. Lyapunov functions & stronger uniformity

Loosely speaking, (Foster–)Lyapunov functions are functions enjoying certain contraction properties with respect to the random walk, to the effect that (on average) its dynamics are directed towards the ‘centre’ of the space, where the function takes values below some threshold. They were introduced into the study of random walks on homogeneous spaces by Eskin–Margulis [Citation15], whose ideas were further developed by Benoist–Quint [Citation6].

Definition 4.5

A measurable function V:X[0,] is called a Lyapunov function for the random walk on X induced by µ if

  1. it is proper, in the sense that the sublevel sets V1([0,L]) are relatively compact for L[0,), and

  2. there exist constants α<1, β0 such that π(μ)VαV+β, where π(μ) is the convolution operator associated to µ introduced in Section 3.

The inequality in the second condition above is referred to as the contraction property of V.

Allowing Lyapunov functions to take the value ∞ is conceptually important for the proofs of results such as Theorem 1.1, in order to show that the random walk does not accumulate near a lower-dimensional homogeneous subspace. Also, affording the possibility of non-continuous Lyapunov functions is crucial in recent constructions given in the literature [Citation6,Citation19]. For the purposes of the discussion in this section, however, it is no big restriction to have in mind the case of a continuous Lyapunov function which is finite on all of X.

Remark 4.6

Let us collect some immediate observations about Lyapunov functions.

  1. If V is a Lyapunov function, then so are cV and V + c for any constant c>0. In particular, one may impose an arbitrary lower bound on V, so that it is no restriction to assume that a Lyapunov function takes values 1, say.

  2. Given a Lyapunov function V:X[0,] for the n0-step random walk (induced by the convolution power μn0), one can construct a Lyapunov function V for the random walk given by µ itself by setting V=k=0n01αn01kn0π(μ)kV.

  3. By enlarging α and using properness, the contraction property in the definition of a Lyapunov function V may be replaced by π(μ)VαV+β1Kfor some compact KX, where 1K denotes the indicator function of K (cf. [Citation17, Lemma 15.2.8]).

Two examples in which a Lyapunov function exists are the following.

Example 4.7

[Citation15]

Identify X=SL2(R)/SL2(Z) with the space of unimodular lattices in R2 as in Example 3.7 and recall that we denote by λ1(x) the length of a shortest non-zero vector in xX. Then for every compactly supported probability measure µ on G whose support generates a Zariski dense subgroup there exist ε,δ>0 such that V=1+ελ1δ is a finite continuous Lyapunov function for the n0-step random walk on X induced by μn0 for some n0N. This construction can be generalized to higher dimensions by taking into account the higher successive minima λ2,,λd of lattices in Rd. A more advanced construction also ensures existence of Lyapunov functions for Zariski dense probability measures with finite exponential moments when G=G(R) is the group of real points of a Zariski connected semisimple algebraic group G defined over R such that G has no compact factors.

Example 4.8

[Citation27]

Let G=SLd+1(R), Γ=SLd+1(Z) and X=G/Γ. For 0im let ci>1 be positive real numbers, yiRd vectors such that y0=0 and y1,,ym span Rd, OiSOd(R) and set gi=(ciOiyi0cid)G.Then for any choice of p0,,pm>0 with i=0mpi=1, the measure μ=i=0mpiδgi defines a uniquely ergodic random walk on X admitting a finite continuous Lyapunov function.

It is well known that existence of a Lyapunov function implies recurrence properties of the random walk.

Lemma 4.9

[Citation15, Lemma 3.1]

Suppose the random walk on X given by µ admits a finite continuous Lyapunov function V. Then this random walk is locally uniformly recurrent.

The intuitive reason for this behaviour is simple: The contraction property means that after a step of the random walk, the value of the Lyapunov function V on average gets smaller by a constant factor, at least when starting outside some compact set K (cf. Remark 4.6(iii) above), which one can think of as the ‘centre’ of the space. The set K can be chosen as (closure of) a sublevel set of V. By the contraction property, the number of steps required to reach it is uniform over starting points x in any given sublevel set of V, or in any given compact subset of X in the case that V is finite and continuous. This suggests that we might even let the starting points diverge, as long as this divergence is outcompeted by the geometric rate of contraction of V. We are led to the following notion of recurrence.

Definition 4.10

Let (Kn)n be a sequence of subsets of X. We say that the random walk on X given by µ is (Kn)n-uniformly recurrent if for every ε>0 there exists n0N and a compact subset MX with μnδx(M)1εfor all nn0 and xKn. It is called (Kn)n-uniformly recurrent on average if the above holds with the Cesàro averages 1nk=0n1μkδx in place of μnδx.

Remark 4.11

We point out that contrary to the locally uniform situation, for the two versions of this property (with/without average) it is generally not clear whether one implies the other.

We are now going to establish such recurrence properties for certain families (Kn)n of sublevel sets of Lyapunov functions, which can be chosen to be increasing and to exhaust the part of X where the Lyapunov function is finite. Recall that the Lyapunov exponent of a function φ:N[1,) is the exponential growth rate λ(φ)=lim supn1nlogφ(n).If λ(φ)=0, we say that φ has sub-exponential growth.

Proposition 4.12

Let φ:N[1,) be a function. Suppose that the random walk on X induced by µ admits a Lyapunov function V with contraction factor α<1 and set Kn=V1([0,φ(n)]).

  1. If φ has Lyapunov exponent λ(φ)<log(α1), then the random walk on X given by µ is (Kn)n-uniformly recurrent. The number n0 in the definition can be chosen independently of ε.

  2. If φ has sub-exponential growth, then the random walk on X given by µ is (Kn)n-uniformly recurrent on average.

The proof is a refinement of the methods in [Citation6,Citation15].

Proof.

Let α,β be the constants from the contraction property of V and define B=β1α. We are going to use the same set M for both parts of the proposition, namely M=V1([0,2B/ε])¯, which is compact since V is proper. Then for nN and xKn we find, by repeatedly using the contraction property of V, μnδx(Mc)ε2Bπ(μ)nV(x)ε2B(αnV(x)+B)ε2Bαnφ(n)+ε2.When the exponential growth rate of φ is less than log(α1), for some n0N we have αnφ(n)B for all nn0. This proves (i).

In order to prove (ii) we use a similar estimate, but have to ensure that the values μkδx(Mc) are small for a sufficiently large proportion of 0k<n. For xKn we find, as above, (17) μkδx(Mc)ε2Bαkφ(n)+ε2.(17) Using straightforward manipulations, we further see αkφ(n)B/2knlog(α1)1(1nlogφ(n)1nlog(B/2)),the right-hand side of which tends to 0 as n by sub-exponential growth of φ. Hence, with k(n)=εn/4, we may choose n0 large enough to ensure the above inequality holds for all kk(n) for nn0. For such n we conclude, using (Equation17), 1nk=0n1μkδx(Mc)=1nk=0k(n)1μkδx(Mc)+1nk=k(n)n1μkδx(Mc)k(n)n+3ε4ε,which ends the proof of (ii).

Theorem 4.4 can now be strengthened in the following way.

Theorem 4.13

In addition to the assumptions of Theorem 4.4, suppose that the random walk on X induced by µ admits a Lyapunov function V. Let φ:N[1,) have sub-exponential growth. Then for every fCc(X) we have limnsupV(x)φ(n)|1nk=0n1Xfd(μkδx)XfdmX|=0.

Proof.

Using (Kn)n-uniform recurrence on average for Kn=V1([0,φ(n)]) from Proposition 4.12(ii), the proof of Theorem 4.4 goes through with the obvious modifications.

Acknowledgments

The author would like to express his gratitude to Andreas Wieser for valuable comments on preliminary versions of the article, and to Manfred Einsiedler for explaining how relative traces can be used to make separability effective. Thanks also go to HIM Bonn and the organizers of the trimester program ‘Dynamics: Topology and Numbers’, in the course of which parts of this manuscript were completed, for hospitality and providing an excellent working environment. Finally, the author is grateful to the anonymous referee for pointing out a simple way to establish a better speed of convergence in Theorems 3.2 and 3.13.

Disclosure statement

No potential conflict of interest was reported by the author(s).

References

  • Bekka M.B., On uniqueness of invariant means, Proc. Amer. Math. Soc. 126 (1998), pp. 507–514.
  • Bénard T., Equidistribution of mass for random processes on finite-volume spaces, Israel J. Math.255 (2023), pp. 417–422.
  • Bénard T. and de Saxcé N., Random walks with bounded first moment on finite-volume spaces, Geom. Funct. Anal. 32 (2022), pp. 687–724.
  • Benoist Y. and Quint J.-F., Mesures stationnaires et fermés invariants des espaces homogènes, Ann. Math. (2) 174 (2011), pp. 1111–1162.
  • Benoist Y. and Quint J.-F., Introduction to random walks on homogeneous spaces, Jpn. J. Math. 7 (2012), pp. 135–166.
  • Benoist Y. and Quint J.-F., Random walks on finite volume homogeneous spaces, Invent. Math. 187 (2012), pp. 37–59.
  • Benoist Y. and Quint J.-F., Stationary measures and invariant subsets of homogeneous spaces (II), J. Amer. Math. Soc. 26 (2013), pp. 659–734.
  • Benoist Y. and Quint J.-F., Stationary measures and invariant subsets of homogeneous spaces (III), Ann. Math. (2) 178 (2013), pp. 1017–1059.
  • Benoist Y. and Quint J.-F., Random Walks on Reductive Groups, Springer, Cham, 2016.
  • Bernstein J. and Reznikov A., Sobolev norms of automorphic functionals, Int. Math. Res. Not. 2002 (2002), pp. 2155–2174.
  • Breuillard E. F., Equidistribution of random walks on nilpotent Lie groups and homogeneous spaces, PhD thesis, Yale University, 2004.
  • Davis Buenger C., Quantitative non-divergence, effective mixing, and random walks on homogeneous spaces, PhD thesis, The Ohio State University, 2016.
  • Einsiedler M., Margulis G., and Venkatesh A., Effective equidistribution for closed orbits of semisimple groups on homogeneous spaces, Invent. Math. 177 (2009), pp. 137–212.
  • Einsiedler M. and Ward T., Functional Analysis, Spectral Theory, and Applications, Springer, Cham, 2017.
  • Eskin A. and Margulis G., Recurrence properties of random walks on finite volume homogeneous manifolds, in Random walks and geometry, Vadim A. Kaimanovich, ed., De Gruyter, Berlin, 2004, pp. 431–444. Proceedings of a Workshop at the Erwin Schrödinger Institute, Vienna, June 18–July 13, 2001. Corrected version: https://www.math.uchicago.edu/eskin/return.ps.
  • Krengel U., Ergodic Theorems, de Gruyter, Berlin, 1985.
  • Meyn S. and Tweedie R. L., Markov Chains and Stochastic Stability, 2nd ed., Cambridge University Press, Cambridge, 2009.
  • Prohaska R., Spread out random walks on homogeneous spaces, Ergodic Theory Dynam. Syst. 41 (2021), pp. 3439–3473.
  • Prohaska R., Sert C., and Shi R., Expanding measures: Random walks and rigidity on homogeneous spaces, Forum Math. Sigma 11(e59) (2023), pp. 1–61.
  • Raghunathan M.S., Discrete Subgroups of Lie Groups, Springer, Berlin, 1972.
  • Ratner M., On measure rigidity of unipotent subgroups of semisimple groups, Acta Math. 165 (1990), pp. 229–309.
  • Ratner M., Strict measure rigidity for unipotent subgroups of solvable groups, Invent. Math. 101 (1990), pp. 449–482.
  • Ratner M., On Raghunathan's measure conjecture, Ann. Math. (2) 134 (1991), pp. 545–607.
  • Ratner M., Raghunathan's topological conjecture and distributions of unipotent flows, Duke Math. J. 63 (1991), pp. 235–280.
  • Shalom Y., Explicit Kazhdan constants for representations of semisimple and arithmetic groups, Ann. Inst. Fourier (Grenoble) 50 (2000), pp. 833–863.
  • Siegel C. L., Lectures on the Geometry of Numbers, Springer, Berlin, 1989.
  • Simmons D. and Weiss B., Random walks on homogeneous spaces and diophantine approximation on fractals, Invent. Math. 216 (2019), pp. 337–394.
  • Zimmer R. J., Ergodic Theory and Semisimple Groups, Birkhäuser, Boston, 1984.