<<

JOURNAL OF THE AMERICAN MATHEMATICAL SOCIETY 28, Number 2, April 2015, Pages 531–577 S 0894-0347(2014)00811-0 Article electronically published on June 4, 2014

ERGODICITY OF UNIPOTENT FLOWS AND KLEINIAN GROUPS

AMIR MOHAMMADI AND

Contents 1. Introduction 531 2. Ergodic properties of BMS and BR measures 537 BR 3. Weak convergence of the conditional of μE,s 542 4. PS density and its non-focusing property when δ>1 545 5. estimate and L2-convergence for the projections 548 6. Recurrence properties of BMS and BR measures 554 7. Window theorem for Hopf average 559 8. Additional invariance and of BR for δ>1 566 9. BR is not ergodic if 0 <δ≤ 1 573 Acknowledgments 576 References 576

1. Introduction In this paper we study dynamical properties of one-parameter unipotent flow for the frame bundle of a convex cocompact hyperbolic 3- M. When the critical exponent of the fundamental group π1(M) exceeds one, we show that this flow is conservative and ergodic for the Burger-Roblin mBR: almost all points enter to a given Borel of positive measure for an unbounded amount of time. Such a manifold admits a unique positive square-integrable eigenfunction φ0 of the Laplacian with eigenvalue. Our result implies that a randomly chosen unipotent , normalized by the time average of the eigenfunction φ0, becomes equidistributed with respect to the Burger-Roblin measure. To state our result more precisely, let G =PSL2(C), which is the group of orien- tation preserving isometries of the hyperbolic H3. Let Γ be a non-elementary, torsion-free, discrete subgroup of G which is convex cocompact,thatis,theconvex core of Γ is compact. Equivalently, Γ\H3 admits a finite sided fundamental domain with no cusps. Convex cocompact groups arise in topology as fundamental groups of compact hyperbolic 3- with totally boundary.

Received by the editors September 15, 2012 and, in revised form, February 23, 2014. 2010 Subject Classification. Primary 11N45, 37F35, 22E40; Secondary 37A17, 20F67. Key words and phrases. Geometrically finite hyperbolic groups, Ergodicity, Burger-Roblin measure, Bowen-Margulis-Sullivan measure. The first author was supported in part by NSF Grant #1200388. The second author was supported in part by NSF Grant #1068094.

c 2014 American Mathematical Society 531

License or copyright restrictions may apply to redistribution; see https://www.ams.org/journal-terms-of-use 532 AMIR MOHAMMADI AND HEE OH

The frame bundle of the manifold M =Γ\H3, which is a circle bundle over the unit tangent bundle T1(M), is identified with the homogeneous space X =Γ\G. We consider the unipotent flow on X given by the right translations of the one- parameter unipotent subgroup 10 (1.1) U = u := : t ∈ R . t t 1

This flow is called ergodic, with respect to a fixed locally finite Borel measure on X, if any invariant Borel subset is either null or co-null. We denote by δ the critical exponent of Γ, which is equal to the Hausdorff dimension of the limit set of Γ[34].Whenδ =2,X is compact [37] and the classical Moore’s theorem in 1966 [23] implies that this flow is ergodic with respect to the volume measure, i.e., the G- . When δ<2, the volume measure is not ergodic any more, and furthermore, Ratner’s measure classification theorem [30] says that there exists no finite U-ergodic invariant measure on X. This raises a natural question of finding a locally finite U-ergodic measure on X. Our main result in this paper is that when δ>1, the Burger-Roblin measure is conservative and ergodic, and is never ergodic otherwise. The conservativity means that for any subset S of positive measure, the U-orbits of almost all points in S spend an infinite amount of time in S. Any finite invariant measure is conservative by the Poincar´e recurrence theorem. For a general locally finite invariant measure, the Hopf decomposition theorem [14] says that any ergodic measure is either conservative or totally dissipative (i.e., for any Borel subset S, xut ∈/ S for all large |t|1anda.e. x ∈ S). For δ<2, there are many isometric embeddings of the real line in X,byt → xut, giving rise to a family of dissipative ergodic measures for U. We refer to the Burger-Roblin measure as the BR measure for short, and give its description using the Iwasawa decomposition G = KAN: K =PSU2, A = {as : s ∈ R}, N = {nz : z ∈ C} where es/2 0 10 as = and nz = . 0 e−s/2 z 1

Furthermore let M denote the centralizer of A in K. The groups A and N play important roles in dynamics as the right translation 1 by as on X is the frame flow, which is the extension of the geodesic flow on T (M) and N-orbits give rise to unstable horospherical foliation on X for the frame flow. 3 Fixing o ∈ H stabilized by K,wedenotebyνo the Patterson-Sullivan measure on the boundary ∂(H3), supported on the limit set of Γ, associated to o ([26], [34]), and refer to it as the PS measure. Sullivan showed that the PS measure coincides with the δ-dimensional Hausdorff measure of the limit set of Γ. Using the transitive 3 action of K on ∂(H )=K/M, we may lift νo to an M-invariant measure on K. BR Burger-Roblin measure Define the measurem ˜ on G as follows: for ψ ∈ Cc(G), BR −δs m˜ (ψ)= ψ(kasnz)e dνo(k)ds dz G where ds and dz are the Lebesgue measures on R and C respectively. It is left Γ-invariant and right N-invariant. The BR measure mBR is a locally finite measure

License or copyright restrictions may apply to redistribution; see https://www.ams.org/journal-terms-of-use BURGER-ROBLIN 533

on X induced bym ˜ BR.Whenδ =2,mBR is simply a G-invariant measure, but it is an infinite measure if δ<2. Roblin showed that the BR measure is the unique NM-invariant ergodic measure on X whichisnotsupportedonaclosedNM-orbit in X [31]. For Γ Zariski dense (which is the case if δ>1), Winter [36] proved that mBR is N-ergodic, and this implies that mBR is the unique N-invariant ergodic measure on X which is not supported on a closed N-orbit in X, by Roblin’s classification. We note that the analogous result for G =PSL2(R) was established earlier by Burger [5] when Γ is convex-compact with δ>1/2. The main result of this paper is: Theorem 1.1. Let Γ be a convex cocompact subgroup of G which is not virtually abelian. The U-flow on (X, mBR) is ergodic if and only if δ>1. We also show the conservativity of the BR-measure for δ>1, without knowing its ergodicity a priori. Indeed we establish the non-ergodicity in the case 0 <δ≤ 1 by proving the failure of “sufficient” recurrence; see Section 9. Remark 1.2. We remark that most of arguments in the proof of Theorem 1.1 works for a higher dimensional case as well. Namely, the same proof will show that if G is the group of orientation preserving isometries of the hyperbolic n-space, Γ is a Zariski dense, convex cocompact subgroup of G,andU is a k-dimensional connected unipotent subgroup of G, then the U action is ergodic with respect to the BR-measure on Γ\G if δ>n− k and not ergodic if δ

License or copyright restrictions may apply to redistribution; see https://www.ams.org/journal-terms-of-use 534 AMIR MOHAMMADI AND HEE OH

The BR measure on X projects down to the absolutely continuous measure on the manifold M and its Radon-Nikodym derivative with respect to the measure is given by φ0. We deduce the following from Theorem 1.1 and Hopf’s ratio theorem (1.3): Corollary 1.3. Let δ>1. (1) For mBR almost all x ∈ X, the projection of xU to M is dense. (2) For any ψ ∈ L1(X, mBR) and for almost all x ∈ X, T ψ(xut)dt lim 0 = ψdmBR. →∞ T T X 0 φ0(xut)dt We explain the proof of Theorem 1.1 in the case δ>1, in comparison with the finite measure case. This account makes our introduction rather lengthy but we hope that this will give a summary of the main ideas of the proof which will be helpful to the readers. The proof of Moore’s ergodicity theorem is based on the following equivalence for a finite invariant measure μ: μ is ergodic if and only if any U-invariant function of L2(X, μ) is constant a.e. Through this interpretation, his ergodicity theorem follows from a theorem in the unitary representation theory 2 that any U-invariant vector in the Hilbert space L (X, μG)isG-invariant for the volume measure μG. For an infinite invariant measure, its ergodicity cannot be understood merely via L2-functions, but we must investigate all invariant bounded measurable functions. This means that we cannot depend on a convenient theorem on the dual space of X, but rather have to work with the geometric properties of flows in the space X directly. We remark that as we are working with a unipotent flow as opposed to a hyperbolic flow, the Hopf argument using the stable and unstable foliations of flows, which is a standard tool in studying the ergodicity for hyperbolic flows, is irrelevant here. We use the polynomial divergence property of unipotent flows to establish that almost all U-ergodic components of mBR are invariant under the full horospherical subgroup N.TheN-ergodicity of the BR measure then implies the U-ergodicity as well. This approach has been noted by Margulis as an alternative approach to show the ergodicity of the volume measure μG in the finite volume case. However, carrying out this argument in an infinite measure case is subtler. In- deed the heart of the argument, as is explained below, lies in the study of two nearby orbits in the “intermediate range”. To the best of our knowledge, such questions of infinite measure have not been understood before. Let us present a sketch of the argument in the probability measure case. Let (X, μ) be a probability measure space. Then it is straightforward from (1.2) that for any generic point x,any0

License or copyright restrictions may apply to redistribution; see https://www.ams.org/journal-terms-of-use BURGER-ROBLIN 535

Let Nˇ and Uˇ denote the transpose of N and U respectively. We denote by NG(U) the normalizer of U in G. Choose sequences of generic points xk and yk inside a suitably chosen compact subset of X, moreover suppose that yk = xkgk with gk ∈/ NG(U)andgk → e.Put 1 it Vˇ = : t ∈ R , and assume that the Vˇ -component and the Uˇ-component1 01 of gk are of “comparable” size. −1 Flowing by ut,wecomparetheorbitsxkut and ykut = xkut(ut gkut). The divergence properties of unipotent flows (a simple computation in our case), in view of our above assumption on gk’s, imply that the divergence of the two orbits is −1 −1 comparable to ut gkut. Furthermore, the (2, 1)-matrix entry of ut gkut dominates −1 other matrix entries. Let p(t)denotethe(2, 1)-matrix entry of ut gkut. This is a polynomial of degree two whose leading coefficient has comparable real and imaginary parts. Therefore, the divergence of the two orbits is “essentially” in the direction of N − U. Choose a sequence of times Tk so that p(Tk)convergestoa non-trivial element v ∈ N − U. Letting ε>0 be small, since p(t)isapolynomial, ykut remains within an O(ε)-neighborhood of xkutv for any t ∈ [(1 − ε)Tk,Tk]. Hence the window theorem (1.4) applied to the sequence of windows [(1 − ε)Tk,Tk] implies that μ(ψ) − μ(v.ψ)=O(ε) and hence μ(ψ)=μ(v.ψ)asε>0 is arbitrary. Repeating this process for a sequence of vn → e, we obtain that the measure μ is invariant under N. We now turn our attention to an infinite measure case, assuming δ>1. There is a subtle difference for the average over the one-sided interval [0,T]andover the two sided [−T,T], and the average over [−T,T] is supposed to behave more typically in infinite . We first prove that the BR measure mBR is U-conservative based on a theorem of Marstrand [19], which allows us to write an BR ergodic decomposition m = x μx where μx is conservative for a.e. x. Letting x be a generic point for Hopf’s ratio theorem and IT =[−T,T], in order to deduce − ψ1(xut)dt IT IrT ∼ μx(ψ1) , it is sufficient to prove that there is some c>0 such that − ψ2(xut)dt μx(ψ2) IT IrT for all T  1,

(1.5) ψ2(xut)dt ≥ c ψ2(xut)dt. IT −IrT IT This type of inequality requires strong control on the recurrence of the flow, and seems unlikely that (1.5) can be achieved for a set of positive measure, see [1, Section 2.4]. Hence formulating a proper replacement of this condition (1.5) and its proof are simultaneously the hardest part and at the heart of the proof of Theorem 1.1. We call x ∈ X a BMS point if both the forward and backward endpoints of the geodesic determined by x belong to the limit set of Γ. These points precisely comprise the support of the Bowen-Margulis-Sullivan measure mBMS on X,whichis the unique measure of maximal for the geodesic flow, up to a multiplicative constant (see Section 2.3). We will call mBMS the BMS measure for simplicity. The support of mBMS is contained in the convex core of Γ, and in particular a compact subset. By a BMS box, we mean a subset of the form x0NˇρAρNρM where x0 ∈ X is a BMS point, ρ>0 is at most the injectivity radius at x0 and Sρ means the ρ-neighborhood of e in S for any S ⊂ G.

1 these components are well defined for all gk close enough to e.

License or copyright restrictions may apply to redistribution; see https://www.ams.org/journal-terms-of-use 536 AMIR MOHAMMADI AND HEE OH

Theorem 1.4 (Window Theorem). Let δ>1.LetE ⊂ X be a BMS box and ψ ∈ Cc(X) be a non-negative function with ψ|E > 0. Then there exist 0 1 such that for any T ≥ T0, rT T BR r BR m x ∈ E : ψ(xut)dt ≤ (1 − r) ψ(xut)dt > · m (E). −rT −T 2 We call x a good point for the window I − I if T rT

ψ(xut)dt ≤ (1 − r) ψ(xut)dt, IrT IT or equivalently if ψ(xut)dt ≥ r ψ(xut)dt. The window theorem says that IT −IrT IT the set of good points for the window IT − IrT has a positive proportion of E for all large T . It follows that for any ε>0, we can choose a sequence Tk = Tk(ε)such that the set Ek, of good points for the window [(1−ε)Tk,Tk](or[−Tk, −(1−ε)Tk]), has positive measure. Let xk,yk = xkgk ∈ Ek. To be able to use this in obtaining an additional invariance, we need to control the size as well as the direction of the divergence u−1g u . More precisely, we need to be able to choose our generic Tk k Tk 1 points yk = xkgk so that the size of gk is comparable with 2 and the size of its Tk Vˇ -component is comparable with that of Uˇ-component. We emphasize here that we work in the opposite order of a standard way of applying the pointwise ergodic theorem where one is usually given a sequence gk

and then find window ITk depending on gk (as the window theorem works for any Tk). In our situation, we cannot choose Tk, but rather have to work with given Tk (depending on ε). So, only after we know which Tk’s give good windows for the ε-width, we can choose good points xkgk for those windows. What allows us to carry out this process is that we have a good understanding of the structure of the generic set along contracting leaves. To be more precise, the PS-measures on the contracting leaves are basically δ-dimensional Hausdorff measures on R2,andthe assumption that δ>1 enables us to find gk for the “right scale”, see Section 4. Hoping to have given some idea about how the above window theorem 1.4 will be used, we now discuss its proof, which is based on the interplay between the BR measure and the BMS measure. We mention that the close relationship between the BR and the BMS measure is also the starting point of Roblin’s unique ergodicity theorem for NM-invariant measures. Unlike the finite measure case, mBR is not invariant under the frame flow, which is the right translation by as in X. However, as s → +∞, the normalized measure BR BR| BR| μs := (a−s)∗m E (the push-forward of the restriction m E by the frame flow BMS a−s)convergestom in the weak* topology. Under the assumption δ>1, the BMS measure turns out to be U-recurrent and hence almost all of its U-leafwise measures are non-atomic. This will imply that the analogue of (1.5) holds for “most” of the U-leafwise measures of mBMS. BR BMS The goal is to utilize this and the fact that μs weakly converges to m ,in order to deduce that many of the U-leafwise measures of mBR must also satisfy (1.5). We mention that, in general, it is unusual to be able to deduce “interesting” statements regarding leafwise measures from weak* convergence of measures. One possible explanation for this is that the leafwise measures of a sequence of mea- sures may change “very irregularly” as one moves in the transversal direction, e.g. approximation of Lebesgue measure by atomic measures.

License or copyright restrictions may apply to redistribution; see https://www.ams.org/journal-terms-of-use BURGER-ROBLIN 537

We succeed here essentially because we have a rather good understanding of the BR N-leafwise measures of μs . To be more precise, we can show (i) the N-leafwise BR measures of μs change rather regularly (see Section 3) furthermore, (ii) the pro- BR 2 jection of an N-leafwise measure of μs converges in the L -sense to its counterpart of mBMS in most directions, see Section 5.1. We emphasize that we established the L2-convergence of these measures, not merely the weak* convergence, and this is crucial to our proof; see the Key Lemma 5.12 and Section 7. The proof of this L2-convergence requires a certain control of BR  the energy of the conditional measures of μs which is uniform for all s 1. Our energy estimate is obtained using the following deep property of the PS measure: δ for all ξ in the limit set of Γ and for all small r>0, νo(B(ξ,r)) r (with the implied constant being independent of ξ and r), together with the Besicovitch covering lemma. Lastly, we remark that our proof of the window theorem makes use of the rich theory of entropy and is inspired by the low entropy method developed by Lindenstrauss in [15].

2. Ergodic properties of BMS and BR measures 2.1. Measures on T1(Γ\H3) associated to a pair of conformal densities. Let (H3,d) denote the hyperbolic 3-space and ∂(H3) its geometric boundary. We denote by T1(H3) the unit tangent bundle of H3 and by π the natural projection from T1(H3) → H3. Denote by {gs : s ∈ R} thegeodesicflow.Foru ∈ T1(H3), we set u+ := lim gs(u)andu− := lim gs(u) s→∞ s→−∞ which are respectively the forward and backward endpoints in ∂(H3)ofthegeodesic defined by u. Definition 2.1. (1) The Busemann function β : ∂(H3) × H3 × H3 → R is defined as follows: for ξ ∈ ∂(H3)andx, y ∈ H3,

βξ(x, y) = lim d(x, ξs) − d(y, ξs) s→∞ 3 where ξs is a geodesic ray tending to ξ as s →∞from a base point o ∈ H , fixed once and for all. ∈ 1 H3 H+ Hˇ (2) For u T ( ), the unstable horosphere u and the stable horosphere u denote respectively the

1 3 − − {v ∈ T (H ):v = u ,βu− (π(u),π(v)) = 0}; 1 3 + + {v ∈ T (H ):v = u ,βu+ (π(u),π(v)) = 0}.

Each element of the group PSL2(C)actsonCˆ = C ∪{∞}as a Mobius trans- formation and its action extends to an isometry of H3, giving the identification 3 of PSL2(C) as the group of orientation preserving isometries of H .Notethat (g(u))± = g(u±)forg ∈ G.ThemapT1(H3) → ∂(H3)givenbyu → u+ is called the visual . For discussions in this section, we refer to [31], [24] and [22]. Let Γ be a non- elementary (i.e., non virtually abelian) torsion-free discrete subgroup of G.Let 3 3 {μx : x ∈ H } be a Γ-invariant conformal density of dimension δμ > 0on∂(H ).

License or copyright restrictions may apply to redistribution; see https://www.ams.org/journal-terms-of-use 538 AMIR MOHAMMADI AND HEE OH

3 That is, each μx is a non-zero finite Borel measure on ∂(H ) satisfying for any x, y ∈ H3, ξ ∈ ∂(H3)andγ ∈ Γ, dμ y −δμβξ(y,x) γ∗μx = μγx and (ξ)=e , dμx −1 3 where γ∗μx(F )=μx(γ (F )) for any Borel subset F of ∂(H ). { } {  } H3 Let μx and μx be Γ-invariant conformal densities on ∂( )ofdimensionδμ μ,μ 1 3 and δμ respectively. Following Roblin [31], we define a measure m on T (Γ\H ) { } {  } ∈ H3 associated to the pair μx and μx . Note that, fixing o ,themap + − u → (u ,u ,βu− (o, π(u))) is a homeomorphism between T1(H3)with(∂(H3)×∂(H3)−{(ξ,ξ):ξ ∈ ∂(H3)})×R. Definition 2.2. Set  − μ,μ δμβu+ (o,π(u)) δμ βu− (o,π(u)) + dm˜ (u)=e e dμo(u )dμo(u )dt. { } {  } μ,μ It follows from the Γ-conformal properties of μx and μx thatm ˜ is Γ- invariant and that this definition is independent of the choice of o ∈ H3. Therefore, it induces a locally finite Borel measure mμ,μ on T1(Γ\H3). 2.2. BMS and BR measures on T1(Γ\H3). Two important densities we will consider are the Patterson-Sullivan density and the G-invariant density. We denote by δ the critical exponent of Γ, that is, the abscissa of convergence of P −sd(o,γ(o)) ∈ H3 the Poincare series Γ(s):= γ∈Γ e for o . As Γ is non-elementary, we have δ>0. The limit set Λ(Γ) is the set of all accumulation points of orbits Γ(z), z ∈ H3. As Γ acts properly discontinuously on H3,Λ(Γ)⊂ ∂(H3). Generalizing the work of Patterson [26] for n = 2, Sullivan [34] constructed a Γ-invariant conformal 3 3 density {νx : x ∈ H } of dimension δ supported on Λ(Γ). Fixing o ∈ H ,eachνx is the unique weak limit as s → δ+ of the family of measures on the 3 3 3 H := H ∪ ∂∞(H ): 1 ν := e−sd(x,γ(o))δ x,s e−sd(o,γ(o)) γ(o) γ∈Γ γ∈Γ

where δγ(o) is the dirac measure at γ(o). This family will be referred to as the PS density. When Γ is of divergence type, i.e., PΓ(δ)=∞, the PS-density is the unique Γ-invariant conformal density of dimension δ (up to a constant multiple) and -free [31, Cor. 1.8]. 3 We denote by {mx : x ∈ H } a G-invariant conformal density on the boundary 3 ∂(H ) of dimension 2, unique up to homothety. In particular, each mx is invariant under the maximal compact subgroup which stabilizes x. Definition 2.3. (1) The measure mν,ν on T1(Γ\H3) is called the Bowen- BMS Margulis-Sullivan measure m associated with {νx} [35]:

BMS δβ + (o,π(u)) δβ − (o,π(u)) + − m (u)=e u e u dνo(u )dνo(u )dt. (2) The measure mν,m on T1(Γ\H3) is called the Burger-Roblin measure mBR associated with {νx} and {mx} ([5], [31]):

BR 2β + (o,π(u)) δβ − (o,π(u)) + − m (u)=e u e u dmo(u )dνo(u )dt.

License or copyright restrictions may apply to redistribution; see https://www.ams.org/journal-terms-of-use BURGER-ROBLIN 539

We will refer to these measures as the BMS and the BR measures respectively, for short. It is worth mentioning that the Riemannian volume measure, in these coordinates, is mm,m. The quotient Γ\C(Λ(Γ)) of the convex hull C(Λ(Γ)) of the limit set modulo Γ is called the convex core of Γ, denoted by C(Γ). A discrete subgroup Γ of G is called geometrically finite if a unit neighborhood of the convex core C(Γ) has finite volume. It is equivalent to saying that Γ\H3 admits a finite sided fundamental domain. A geometrically finite group Γ is called convex cocompact if one of the following three equivalent conditions hold (cf. [4]): (1) C(Γ) is compact; (2) Γ\H3 admits a finite sided fundamental domain with no cusps; (3) Λ(Γ) consists only of radial limit points: ξ ∈ Λ(Γ) is radial if any geodesic ray ξt toward ξ returns to a compact subset for an unbounded sequence of t. The BMS measure is invariant under the geodesic flow. Sullivan showed that for Γ geometrically finite, it is ergodic and moreover the unique measure of maximal entropy ([35], [25]). For Γ convex cocompact, the support of the BMS measure is compact, as its projection is contained in C(Γ). Theorem 2.4. [9] If Γ is geometrically finite and Zariski dense, the PS density of any proper Zariski subvariety of ∂(H3) is zero. 2.3. BMS and BR measures on X =Γ\G. We fix a point o ∈ H3 whose stabilizer group is K := PSU(2). Then the map g → g(o) induces a G-equivariant isometry between G/K and H3.Set

iθ −iθ M := {mθ = diag(e ,e )}.

By choosing the unit tangent vector X0 based at o stabilized by M, G/M can 1 3 be identified with the unit tangent bundle T (H ) via the orbit map g → g(X0). This identification can also be lifted to the identification of the frame bundle of H3 with G. These identifications are all Γ-equivariant and induce identifications of the frame bundle of the manifold Γ\H3 with Γ\G.WesetX =Γ\G. Abusing the notation, we will denote by mBMS and mBR, respectively, the M-invariant lifts of the BMS and the BR measures to X.Forg ∈ G,wesetg± =(gM)± where gM ∈ G/M =T1(H3). For x =Γ\Γg,wewritex± ∈ Λ(Γ) if g± ∈ Λ(Γ); this is well-defined independent of the choice of g. With this notation, the supports of mBMS and mBR are given respectively by Ω:={x ∈ X : x+,x− ∈ Λ(Γ)} and − ΩBR := {x ∈ X : x ∈ Λ(Γ)}. The right translation action of the diagonal subgroup

s/2 −s/2 A := {as := diag(e ,e ):s ∈ R} on G is called the frame flow and it projection to G/M corresponds to the ge- odesic flow. For this action, mBMS is A-invariant and mBR is A-quasi-invariant: BR (2−δ)s BR (a−s)∗m = e m .

License or copyright restrictions may apply to redistribution; see https://www.ams.org/journal-terms-of-use 540 AMIR MOHAMMADI AND HEE OH

Set 10 1 z N := nz = : z ∈ C and Nˇ := nˇz = : z ∈ C z 1 01 and for g ∈ G, H(g):=gN and Hˇ (g):=gN.ˇ The restriction of the projection G → G/M induces a diffeomorphism from 1 3 H(g)(resp. Hˇ (g)) to the horosphere HgM (resp. HˇgM )inT(H ) and hence the ± 3 − visual maps u → u induce diffeomorphisms PH(g) : ∂(H ) −{g }→H(g)and H3 −{ +}→ ˇ ∈ PHˇ (g) : ∂( ) g H(g), respectively, for each g G. Definition 2.5. Let y ∈ G. (1) Set Leb 2βv+ (o,π(vM)) + ∈ dμH(y)(v)=e dmo(v )forv H(y). Leb Leb Leb The measure μH(y) is G-invariant: g∗μH(y) = μg(H(y)); in particular, it is an N-invariant measure on H(y). (2) Set

PS δβv+ (o,π(vM)) + dμH(y)(v)=e dνo(v ). { PS } We note that μH(y) is a Γ-invariant family. Fix a left G-invariant and right K-invariant metric on G which induces the hyperbolic distance d on G/K.

Notation 2.6. (1) For ρ>0 and a subset Y of G,wedenotebyYρ the intersection of Y and the ρ-ball centered at e in G. (2) The M-injectivity radius ρx at x ∈ X is the supremum of ρ such that for Bρ := NˇρAρMNρ,themapBρ → x0Bρ given by g → x0g is injective.

Definition 2.7. AboxinX (around x0) refers to a subset of the form

x0Bρ = x0NˇρAρMNρ ∈ ˇ with 0 <ρ<ρx0 for some x0 X.Notethatx0Bρ coincides with x0NρAρNρM. ± ∈ We call this box a BMS box if x0 Λ(Γ), i.e., if x0 belongs to the support of the BMS measure.

We fix a box x0Bρ.SetT˜ρ := NˇρAρ and Tρ := NˇρAρM. Since the measures mBMS and mBR have the same transverse measures for the unstable horospherical foliations, we have for any ψ ∈ C(x B ), 0 ρ BMS PS m (ψ)= ψ(ynm)dμH(y)(yn)dν˜x T˜ (y)dm ∈ ˜ ∈ ∈ 0 ρ y x0Tρ,m M n Nρ PS ⊗ = ψ(ymn)dμH(ym)(ymn)d(˜νx T˜ m)(ym); ∈ ˜ ∈ 0 ρ ym x0TρM n Nρ BR Leb m (ψ)= ψ(ynm)dμH(y)(yn)dν˜x T˜ (y)dm ∈ ˜ ∈ 0 ρ y x0Tρ,m M Nρ = ψ(ymn)dμLeb (ymn)d(˜ν ⊗ m)(ym) H(ym) x0T˜ρ ym∈x0T˜ρM n∈Nρ that is, dν := dν˜ ⊗ dm denotes the transverse measure of mBMS (and hence x0Tρ x0T˜ρ BR of m )onx0Tρ.

License or copyright restrictions may apply to redistribution; see https://www.ams.org/journal-terms-of-use BURGER-ROBLIN 541

The following easily follows from Theorem 2.4: Corollary 2.8. If Γ is geometrically finite and Zariski dense, and E is a box in X,thenmBR(∂(E)) = 0. 2.4. BR measure in the Iwasawa coordinates G = KAN. The canonical map ι : Nˇ → G/MAN = K/M has a diffeomorphic image S := ι(Nˇ)whichisK/M minus a single point. By abuse of notation, we use the same notation νo for the 2 measure on K which is the trivial extension of the PS measure νo on S = K/M: for ψ ∈ C(K), − ψdνo = ψ(sm) dνo(sX0 )dm K M S where dm is the probability Haar measure of M. The lift of the BR measurem ˜ BR on G can also be written as follows (cf. [24]): for ψ ∈ C (G), c BR −δs m˜ (ψ)= ψ(kasnz)e dνo(k)ds dz G ∈ R C where kasnz KAN, ds and dz are some fixed Lebesgue measures on and ∈ respectively. As usual, this means that for Ψ(Γg)= γ∈Γ ψ(γg) with ψ Cc(G), mBR(Ψ) =m ˜ BR(ψ). 2.5. BR measure associated to a general unipotent subgroup. A horo- spherical subgroup N0 is a maximal unipotent subgroup of G,orequivalently, n −n N0 = {g ∈ G : b gb → e as n →∞} for a non-trivial diagonalizable element b ∈ G.SinceA normalizes N, it follows from the Iwasawa decomposition G = KAN −1 ∈ that any horospherical subgroup N0 is of the form k0 Nk0 for some k0 K.The BR measure associated to N0 is defined to be BR BR mN0 (ψ):=m (k0.ψ) BR where ψ ∈ Cc(X)andk0.ψ(g)=ψ(gk0). As m is M = NK (U)-invariant (here NK (U) being the normalizer of U in K), this definition does not depend on the choice of k0 ∈ K.IfU0 is a one-parameter unipotent subgroup of G, its centralizer CG(U0)inG is a horospherical subgroup. The BR measure associated to U0 means mBR for N = C (U ). N0 0 G 0 2.6. of frame flow and its consequences. Some of important dynami- cal properties of flows on X have been established only under the finiteness assump- tion of the BMS measure. Examples of groups with finite BMS measure include all geometrically finite groups [34] but not limited to those (see [27]). Roblin showed that if |mBMS| < ∞, then Γ is of divergence type. In the following two theorems, BMS BMS we consider the groups Γ with |m | < ∞. We normalize νo so that |m | =1. The following two theorems were proved by [36], based on the the previous works of Babillot [2], Roblin [31], and Flaminio-Spatzier [9]. Theorem 2.9. [36] Suppose that Γ is Zariski dense and |mBMS| =1. (1) The frame flow on X is mixing with respect to mBMS, that is, for any ψ ,ψ ∈ L2(X, mBMS),ass →±∞, 1 2 BMS BMS BMS ψ1(xas)ψ2(x) dm (x) → m (ψ1)m (ψ2). X (2) The BR measure mBR on X is N-ergodic.

License or copyright restrictions may apply to redistribution; see https://www.ams.org/journal-terms-of-use 542 AMIR MOHAMMADI AND HEE OH

(3) If Γ is geometrically finite, mBR is the only N-ergodic measure on X which is not supported on a closed N-orbit. BMS Theorem 2.10. [36] Let Γ be Zariski dense and |m | =1. Then for all ψ1,ψ2 ∈ ⊂ Cc(X) or for ψ1 = χE1 ,ψ2 = χE2 where Ei X is a bounded Borel subset with mBMS(∂(E )) = 0, we have: as s → +∞, i BR BMS BR ψ1(xa−s)ψ2(x) dm (x) → m (ψ1)m (ψ2). X We note that by the quasi-invariance of the BR measure, BR (2−δ)s BR ψ1(xa−s)ψ2(x) dm (x)=e ψ1(x)ψ2(xas) dm (x). X X In particular, the above theorem implies that if δ<2, BR ψ1(x)ψ2(xas) dm (x) → 0ass → +∞. X Lemma 2.11. If Γ is a discrete subgroup of G with δ>1,thenΓ is Zariski dense in G.

Proof. Let G0 be the identity component of the Zariski closure of Γ. Suppose G0 is a proper subgroup of G. Being an algebraic subgroup of G, G0 is contained either in a parabolic subgroup of G or in a subgroup isomorphic to PSL2(R). In either case, the critical exponent of G0 is at most 1. This leads to a contradiction and hence G0 = G.  Weak convergence of the conditional of BR 3. μE,s In this section, we suppose that Γ is a Zariski dense discrete subgroup of G admitting a finite BMS measure, which we normalize so that |mBMS| =1. Fix a bounded M-invariant Borel subset E ⊂ X with mBR(E) > 0and mBR(∂(E)) = 0. BR For each s>0, define a Borel measure μE,s on X to be the normalization of the BR push-forward (a− ) ∗ m | :forΨ∈ C (X), s E c BR 1 BR − μE,s(Ψ) := BR Ψ(xa s) dm (x). m (E) E Equivalently, (2−δ)s BR e BR μE,s(Ψ) = BR Ψ(x)χE(xas)dm (x). m (E) X BR Note that μE,s is a probability measure supported in the set Ea−s. The following is immediate from Theorem 2.10: → ∞ BR BMS Theorem 3.1. As s + , μE,s weakly converges to m , that is, for any Ψ ∈ Cc(X), lim μBR (Ψ) = mBMS(Ψ). s→+∞ E,s For simplicity, we will write for x ∈ X, Leb PS PS dλx(n)=dμH(x)(xn)anddμx (n)=dμH(x)(xn) PS BR BMS so that λx and μx are respectively the conditional measures of m and m on xN.

License or copyright restrictions may apply to redistribution; see https://www.ams.org/journal-terms-of-use BURGER-ROBLIN 543

Recall that ρx denotes the injectivity radius at x. ∈ Definition 3.2. Fix x X.Fors>0, define a Borel measure λE,x,s on xNρx as ∈ follows: for ψ Cc(xNρx ), e(2−δ)s λE,x,s(ψ):= BR ψ(xn)χE(xnas)dλx(n). m (E) ∈ n Nρx ˇ ∈ ≤ Recall the notation Tρ = NρAρM for ρ>0. Let x0 X and let 0 <ρ ρx0 . For any box x0Bρ = x0TρNρ and Ψ ∈ C(x0Bρ), we have (2−δ)s BR e BR μE,s(Ψ) = BR Ψ(x)χE(xas)dm (x) m (E) ∈ x X e(2−δ)s = BR Ψ(xn)χE(xnas)dλx(n)dνx0Tρ (x) m (E) ∈ ∈ x x0Tρ n Nρ | = λE,x,s(Ψ xNρ )dνx0Tρ (x). x∈x0Tρ BR Hence λE,x,s is precisely the conditional measure of μE,s on xNρ. The aim of this section is to prove: − Theorem 3.3. Suppose that x ∈ Λ(Γ) and 0 <ρ<ρx. For any ψ ∈ Cc(xNρ), −→ PS → ∞ λE,x,s(ψ) μx (ψ) as s + . − The condition x ∈ Λ(Γ) is needed to approximate the measure λE,x,s by its thickening in the transverse direction. For a function Ψ on X and ε>0, we define functions on X as follows: + − Ψε (y):=supΨ(yg)andΨε (y):= inf Ψ(yg) g∈O g∈Oε ε

where Oε is a symmetric ε-neighborhood of e in G.Wealsoset + O − ∩ Eε := E ε and Eε = u∈Oε Eu.

Lemma 3.4. Let x ∈ X and 0 <ρ<ρx. For all small ε>0,thereexistsε1 > 0 ∈ ∈ such that for any non-negative Ψ C(xTε1 Nρ) and any t Tε1 , we have −ε − ≤ ≤ ε + e λx(Ψε ) λxt(Ψ) e λx(Ψε ).

Proof. Let 0 <ε<ρx−ρ. Consider the map φt : xN → xtN given by φt(xn)=xtn, ∗ −1 so that φt λxt = λx.Sinceφt is a translation by n tn,thereexistsε1 > 0such ∈ ∈ −1 ∈O that for all n Nρ and t Tε1 , n tn ε and the Radon-Nikodym derivative − satisfies e ε ≤ dλxt (n) ≤ eε. dλx Therefore −1 ≤ ε + ε + λxt(Ψ) = Ψ(xn(n tn))dλxt(n) e Ψε (xn)dλx(n)=e λx(Ψε ).

The other inequality follows similarly. 

Lemma 3.5. Let x ∈ X and 0 <ρ<ρx. For any ε>0,thereexistsε1 > 0 such ∈ ∈ that for any non-negative Ψ C(xTε1 Nρ),anyt Tε1 and any s>0, −ε − ε + e λ − (Ψ ) ≤ λ (Ψ) ≤ e λ + (Ψ ). Eε ,x,s 2ε E,xt,s Eε ,x,s 2ε

License or copyright restrictions may apply to redistribution; see https://www.ams.org/journal-terms-of-use 544 AMIR MOHAMMADI AND HEE OH

O −1 ⊂O Proof. Let ε1 > 0 be as in Lemma 3.4. We may also assume that n ε1 n ε ∈ for all n Nρ. αw − ∈ ˇ ∈  For t = 0 α 1 NAM and nz N with α + zw = 0, define z2w + αz − α−1z α + zw w ψt(z)= and bt,z = −1 . α + zw 0 ψt(z)w + α − zw

Then by a direct computation, we verify that

(3.1) tnz = nz+ψt(z)bt,z. ∈ Therefore we may assume that ε1 > 0 is small enough so that for all t Tε1 and ∈ { ∈ }⊂ ∈ nz Nρ,wehave nψt(z) : nz Nρ Nε, bt,z Tε, and the absolute value of the | Jacobian of the map ψt Nρ is at most ε/2.

We observe that nzas = nz+ψt(z)as(a−sbt,zas) and since the conjugation by a−s ˇ ∈ O contracts NA for s>0, we have nzas nz+ψt(z)as εM. ≤ + Since E is M-invariant, we deduce that χE (xtnzas) χE (xnz+ψt(z)as) for all ∈ ∈ ε ∈ t Tε1 and nz Nρ. Together with Lemma 3.4, we now obtain that for any t Tε1 , (2−δ)s λE,xt,s(Ψ) = e Ψ(xtnz)χE(xtnzas)dλxt(z) ∈ nz Nρ (2−δ)s + ≤ + e Ψε (xnz)χE (xnz+ψt(z)as)dλxt(z) ∈ ε nz Nρ ε (2−δ)s + ≤ + e e Ψε (xnz)χE (xnz+ψt(z)as)dλx(z) ∈ ε nz Nρ 2ε (2−δ)s + ≤ + e e Ψε (xnz−ψt(z))χE (xnzas)dλx(z) ∈ ε nz Nρ+ε 2ε (2−δ)s + ≤ e e Ψ (xn )χ + (xn a )dλ (z) 2ε z Eε z s x nz ∈Nρ+3ε 2ε + = e λ + (Ψ ) Eε ,x,s 2ε ∩ + where the last inequality follows since Nρ+3ε contains xN supp(Ψ2ε). The other inequality can be proven similarly. 

Theorem 3.3 follows from:

− Theorem 3.6. Let x ∈ Λ(Γ) and ρ<ρx.Letψ ∈ C(xNρ) be a non-negative function. For ε>0,thereexistss0  1 such that for any s>s0, −4ε ≤ PS ≤ 4ε e λE,x,s(ψ) μx (ψ) e λE,x,s(ψ). Moreover, if x+ ∈ Λ(Γ) and ψ is positive, then the above integrals are all non-zero. − ∈ Proof. Let ε1 be as in Lemma 3.5. We note that as x Λ(Γ), ν(xTε1 ) > 0. Hence ∈ there exists a non-negative continuous function φ C(xTε1 ) with ν(φ) = 1. Define ∈ Ψ C(xTε1 Nρ)by ∈ Ψ(xtn):=ψ(xn)φ(xt)forxtn xTε1 Nρ.

License or copyright restrictions may apply to redistribution; see https://www.ams.org/journal-terms-of-use BURGER-ROBLIN 545

+ − ∈ Set ψε (xn)=supu∈Nε ψ(xnu)andψε (xn) = infu Nε ψ(xnu). Then by Lemma 3.5, BR μE,s(Ψ) = λE,xt,s(Ψ)dνxTρ (xt) xT ε1 ε + ≤ e λ + (ψ )φ(xt)dν (xt) Eε ,x,s 2ε xTρ xTε1 ε + = e λ + (ψ ). Eε ,x,s 2ε We can prove the other inequality similarly and hence −ε − BR ε + (3.2) e λ − (ψ ) ≤ μ (Ψ) ≤ e λ + (ψ ). Eε ,x,s 2ε E,s Eε ,x,s 2ε → PS Since the map t μxt is continuous by [31, Lemma 1.16], we have −ε PS ≤ BMS ≤ ε PS e μx (ψ) m (Ψ) e μx (ψ)

by replacing ε1 by a smaller one if necessary. BR → BMS Since μE,s(Ψ) m (Ψ) by Theorem 3.1, we deduce from (3.2) that there exists s0 > 1 such that for all s>s0, −2ε − PS 2ε + (3.3) e λ − (ψ ) ≤ μ (ψ) ≤ e λ + (ψ ). Eε ,x,s 2ε x Eε ,x,s 2ε We claim that −4ε ± 4ε (3.4) e λ (ψ) ≤ λ ± (ψ ) ≤ e λ (ψ) E,x,s Eε ,x,s 2ε E,x,s which will complete the proof of the theorem by (3.3). We can deduce from (3.2) that −ε BR − − + ε BR + ≤ − ≤ ≤ e μ − (Ψ4ε) λE ,x,s(ψ2ε) λE+,x,s(ψ2ε) e μ + (Ψ4ε). E2ε,s ε ε E2ε,s Since it follows from Theorem 3.1 that −ε BR ≤ BR ± ≤ ε BR  e μE,s(Ψ) μ ± (Ψ4ε) e μE,s(Ψ) for all large s 1, E2ε,s we have −2ε BR − + 2ε BR e μ (Ψ) ≤ λ − (ψ ) ≤ λ + (ψ ) ≤ e μ (Ψ). E,s Eε ,x,s 2ε Eε ,x,s 2ε E,s This implies + 4ε − λ + (ψ ) ≤ e λ − (ψ ) Eε ,x,s 2ε Eε ,x,s 2ε and hence (3.4) follows. 

4. PS density and its non-focusing property when δ>1 Let Γ be a (non-elementary) convex cocompact subgroup of G. The assumption on Γ being convex cocompact is crucial for the following theo- rem:

Theorem 4.1. For any compact subset F0 of X,thereexistsc0 = c0(F0) > 1 such + that for any x ∈ F0 with x ∈ Λ(Γ) and for all 0

where xNr = {xnz : |z|

License or copyright restrictions may apply to redistribution; see https://www.ams.org/journal-terms-of-use 546 AMIR MOHAMMADI AND HEE OH

− Similarly, for any x ∈ F0 with x ∈ Λ(Γ) and for all 0

for xNˇr = {xnˇz : |z| 0 for Γ convex cocompact. 

Lemma 4.2. Let δ>1 and F0 ⊂ X be a compact subset. For every ε>0,there − exists a positive integer d = d(ε, F0) such that for any x ∈ F0 with x ∈ Λ(Γ) and for all small 0

Bd(x, r):={xnˇz : |z| 1 is an absolute constant independent of d and r. Let d = d(ε)  1 be such that c d1−δ

Lemma 4.3. There exists b0 > 1 such that for all small 0 <η 1

−1 −1 ˇ −1 ⊂ ˇ ⊂ ˇ N A N M NηAηNηM Nb0ηAb0ηNb0ηM. b0 η b0 η b0 η Proof. The claim follows since the product maps Nˇ × A × N × M → G by (ˇn, a, n, m) → nanmˇ and N × A × Nˇ × M → G by (n, a, n,ˇ m) → nanmˇ are local diffeomorphisms at the identity.  We will use the above results to prove the following proposition 4.4. The proof is elementary and is based on the fact we have a good control of the conditional measures on contracting leaves, i.e., Nˇ-orbits. However, the fact that this statement holds is quite essential to our approach. Indeed, as we explained in the introduction, one major difficulty we face is that the return times for our U-flow do not have the regularity one needs in order to get the required ergodic theorem on the nose. In our version of the window theorem, the set where a window estimate holds depends on time; see Section 7, and in particular Theorem 7.7 below. Usually, in arguments with a similar structure as ours, this fact is fatal as one has very little control on the structure of the “generic” set for the measure in question. In our case, however, the following proposition saves the day and provides us with a rather strong control. ˇ ± ∈ In the following proposition we fix a BMS box E = x0NρAρNρM with x0 Λ(Γ) 1 and 0 <ρ< infx∈Ω ρx where b0 is as in Lemma 4.3 and ρx is the injectivity radius b0 at x.

Proposition 4.4. Let δ>1.Fix0 1 and s0  1 such that for any Borel subset F ⊂ E with m (F ) >

License or copyright restrictions may apply to redistribution; see https://www.ams.org/journal-terms-of-use BURGER-ROBLIN 547

BR r · m (E) and any s ≥ s0, there exists a pair of elements xs,ys ∈ F satisfying ∈ ˇ (1) xs = ysnˇws for nˇws N, 1 d0 (2) ≤|ws|≤ and d0s s |(ws)| (3) |(ws)|≥ . d0

Proof. Let c0 > 1 be as in Theorem 4.1 where F0 is the 2ρ-neighborhood of Ω. We ˇ ∈ − − will write B(z,ρ)=zNρ in this proof. For all x x0NρAρM, x = x0 and hence x− ∈ Λ(Γ). Hence by Theorem 4.1, −1 δ ≤ PS ≤ δ (4.1) c0 η μHˇ (x)(B(x, η)) c0η for all 0 <η<1.

ν(x0Nb0ρAb0ρM) BR Set d1 := mBR(E) where ν denotes the transverse measure of m on x N A M. We claim that there exists z ∈ x N A M with μPS (B(z,b ρ)∩ 0 ρx0 ρx0 0 b0ρ b0ρ Hˇ (z) 0 r F ) > d . Suppose not; then 1 BR ≤ PS ∩ m (F ) μHˇ (z)(B(z,b0ρ) F ) dν(z) ∈ z x0Nb0ρAb0ρM ≤ r · BR ν(x0Nb0ρAb0ρM)=r m (E) d1 which contradicts the assumption on F . Set Q := B(z,b ρ) ∩ F ∩ supp (μPS )andforeachs>1, consider the covering 0 Hˇ (z) {B(x, s−1) ⊂ Hˇ (z):x ∈ Q} of Q. By the Besicovitch covering lemma (cf. [20]), there exist κ>0 (independent of s) and a finite subset Qs such that the corre- −1 sponding finite subcover {B(x, s ):x ∈ Qs} of Q is of multiplicity at most κ. Note that for q>1, by (4.1), PS 1 2 −δ 2 PS 1 2 −δ 3 δ δ μ ∪ ∈ B x, ≤ κ q c μ ∪ ∈ B x, ≤ κ q c b ρ . Hˇ (z) x Qs qs 0 Hˇ (z) x Qs s 0 0 ≥ 2 −δ 3 δ δ r Hence by taking q 1 large so that κ q c0b0ρ < 3d ,wehave 1 PS 1 r ∪ ∈ μHˇ (z) x Qs B x, < . qs 3d1 If we set 1 |(w)| R(s, d):=∪ ∈ w ∈ B x, : |(w)|≤ , x Qs s d

it follows from Lemma 4.2 that there exist d2 > 1ands0 > 1 such that for any s>s0, PS r μHˇ (z)(R(s, d2)) < . 3d1 Hence for any s>s0,theset 1 Q − ∪ ∈ B x, ∪ R(s, d ) x Qs qs 2

PS r has a positive μ measure (at least ). In particular, there exists xs ∈ Q such Hˇ (z) 3d1 ∩ 1 − 1 ∪ PS that (Q B(xs, s )) (B(xs, qs ) R(s, d2)) has a positive μHˇ (z) measure. Picking ys from this set, we have found a desired pair xs,ys from F with d0 =max(q, d1). 

License or copyright restrictions may apply to redistribution; see https://www.ams.org/journal-terms-of-use 548 AMIR MOHAMMADI AND HEE OH

5. Energy estimate and L2-convergence for the projections Let Γ be a convex cocompact subgroup of G with δ>1 and fix a BMS box E ⊂ X (see 2.7 for its definition). We have mBR(E) > 0 and by Lemma 2.11 and Corollary 2.8, mBR(∂(E)) = 0.

± 1 In the entire section, we fix x ∈ X with x ∈ Λ(Γ) and 0 <ρ<√ ρx. 2 ∈ Recall the definition of the measure λE,x,s on xNρx from (3.2): for ψ C(xNρx ), e(2−δ)s λE,x,s(ψ)= BR ψ(xn)χE(xnas)dλx(n). m (E) ∈ n Nρx

PS 5.1. Projections of μH(x) and λE,x,s. The N-orbit of x can be identified with R2 via the visual map xn → (xn)+ ∈ ∂(H3) −{x−} and the identification of ∂(H3) −{x−} with R2 by mapping x− to the point at infinity. Therefore, we may PS R2 consider λE,x,s and μH(x) as measures on . Let 10 10 U = { : t ∈ R}; V = { : t ∈ R}. t 1 it 1 In the sequel by a measure on [0, 2π] we mean the normalized Lebesgue measure. ∈ −1 −1 For each θ [0, 2π), we set Uθ = mθUmθ and Vθ = mθVmθ .Wemayidentify 2 Uθ as the line in R in the θ-direction and Vθ as the line in the θ + π/2 direction. We denote by pθ : UθVθ → Vθ the projection parallel to the line Uθ.Forτ>0, set τ { ∈ − } τ { ∈ − } Uθ := t exp(iθ):t [ τ,τ] and Vθ := it exp(iθ):t [ τ,τ] . Definition 5.1. Fix 0 <θ<π,0<τ≤ ρ and s>1. We define the measures on τ τ xV as follows: for ψ ∈ Cc(xV ), θ θ τ PS σx,θ(ψ):= ψ(pθ(y)) dμH(x)(y), ρ τ xVθ Uθ and τ σx,θ,s(ψ):= ψ(pθ(y)) dλE,x,s(y). ρ τ xVθ Uθ τ τ PS That is, σ and σ are respectively the push-forwards of μ | ρ τ and x,θ x,θ,s H(x) xVθ Uθ λE,x,s| ρ τ via the map pθ. xVθ Uθ 5.2. Energy and Sobolev norms of the projections. Consider the Schwartz 2 α (β) 2 (β) space S := {f ∈ L (xVθ):t f ∈ L (xVθ)}, where α, β ∈ N ∪{0} and f is the β-th derivative of f.DenotebyS the dual space of S with the strong dual topology, which is the space of tempered distributions. For r>0, we consider the following Sobolev space r  r ˆ 2 H (xVθ):={f ∈S :(1+|t|) f ∈ L (xVθ)} with the norm r | | ˆ 2 f 2,r := (1 + t ) f L (xVθ ) where fˆ denotes the Fourier transform of f.

License or copyright restrictions may apply to redistribution; see https://www.ams.org/journal-terms-of-use BURGER-ROBLIN 549

We recall the notion of α-energy: Definition 5.2 (α-energy). For α>0 and a μ on R2,theα-energy of μ is given by 1 Iα(μ):= α dμ(x)dμ(y). R2 R2 |x − y| It is a standard fact that I (μ) can be written as α ∞ μ(B(x, )) (5.1) Iα(μ)=α 1+α ddμ(x) R2 0  where B(x, ) is the Euclidean disc around x of radius . The α-energy of a measure μ is a useful tool in studying the projections of μ in various directions. See [28, Proposition 2.2] or [21, Theorem 4.5] for the following theorem: Theorem 5.3. Let ν be a Borel probability measure on R2 with compact support. If the 1-energy of ν is finite, i.e., I1(ν) < ∞, then the following hold:

(1) pθ∗ν is absolutely continuous with respect to the Lebesgue measure for al- most all θ; (2) there exists c>1 (independent of ν) such that for any 0

where D(pθ∗ν) is the Radon-Nikodym derivative of pθ∗ν with respect to the Lebesgue measure. Lemma 5.4. Let Q ⊂ R2 be a compact subset, c>0 and β>0 be fixed. Let M be a collection of Borel measures on Q such that (5.2) μ(B(x, )) 0. Then for any 0 <α<β, sup Iα(μ) < ∞. μ∈M Proof. Fix 0 <α<β. We use (5.1). Note that since μ(B(x, )) ≤ μ(Q), (5.2) has meaning only when  is not too big. We use (5.2) only for 0 <<1 and use the upper bound of μ(Q)for ≥ 1. We have 1 1 μ(B(x, )) ∞ μ(B(x, )) Iα(μ)= 1+α ddμ(x)+ 1+α ddμ(x) α Q 0  Q 1  ≤ · β−α|=1 · −α|=∞ · 2 c  =0 μ(Q)+ =1 μ(Q) = μ(Q)(c + μ(Q)). ∞ Now, since Q is compact, the assumption implies that supμ∈M μ(Q) < . Hence, Iα(μ) is uniformly bounded for all μ ∈M.  Corollary 5.5. Fix 0 <τ≤ ρ. The following holds for almost all θ: τ τ (1) σx,θ is absolutely continuous with respect to the Lebesgue measure on xVθ ; (2) its support has a positive Lebesgue measure; τ ∈ r τ (3) its Radon-Nikodym derivative satisfies D(σx,θ) H (xVθ ) for any 0

License or copyright restrictions may apply to redistribution; see https://www.ams.org/journal-terms-of-use 550 AMIR MOHAMMADI AND HEE OH

Proof. It follows from Theorem 4.1 and Lemma 5.4 that for any 0 <α<δ, I (μPS | ) < ∞. α H(x0) x0Nρ Now, the fact that the support of the projection has positive measure follows from [19, Theorem I]. The other two claims follow from Theorem 5.3. 

δ − 1 We fix 0

5.3. Uniform bound for the energy of λE,x,s, s ≥ 1. In this subsection, we set † | λE,x,s := λE,x,s xNρ . M { † ≥ } We will show that the collection = λE,x,s : s 1 of measures on xNρ satisfies † the hypothesis of Lemma 5.4 with β = δ. We may consider λE,x,s as a measure on R2 supported on the ρ-ball around the origin. − Since E is a BMS box, E is of the form x0Nr0 Ar0 Nr0 M for some 0

Proof. Suppose xn ∈ Ea−s,sothatxn = x0nˇwatnzmθa−s with |z|

xn = x0nˇwat−smθne−se2πiθz. + + + ∈ If we set y := x0nˇwat−smθ,theny = x0 . Hence y Λ(Γ). Since xn = −s 2πiθ −s yne−se2πiθz and |e e z| 0 such that for all s 1, y supp(λE,x,s) and any >0, † · δ λE,x,s(B(y, ))

where B(y, )={ynz : |z| <}.

Proof. Since B(y, 2ρ) contains xNρ, it suffices to show the above for 0 <<2ρ. Since supp(λE,x,s) ⊂ Ea−s ∩ xNρ, it follows from Lemma 5.7 that for each z ∈ † −s −s ∈ + ∈ supp(λE,x,s), B(z,3ρe ) contains B(w, ρe )forsomew H(z) with w Λ(Γ). Hence by Theorem 4.1 we have PS −s ≥ PS −s ≥ −1 δ −δs μH(z)(B(z,3ρe )) μH(z)(B(w, ρe )) c0 (3ρ) e

where c0 is as in Theorem 4.1 with F0 being the ρ-neighborhood of Ω. Consider † −s ∈ † the covering of supp(λE,x,s) given by the balls B(z,3ρe ), z supp(λE,x,s). By ⊂ † the Besicovitch covering lemma we can choose a finite set Js supp(λE,x,s)such −s that the corresponding finite collection {B(z,3ρe ):z ∈ Js} has multiplicity at † most κ (independent of s) and covers supp(λE,x,s).

License or copyright restrictions may apply to redistribution; see https://www.ams.org/journal-terms-of-use BURGER-ROBLIN 551

Now we consider two cases for . Case 1. 0 <≤ e−s. ∈ † In this case, for any y supp(λE,x,s), we have † ≤ (2−δ)s 2 ≤ δ λE,x,s(B(y, )) πe  π . −s −s Case 2. e <<2ρ.LetJy,s = {z ∈ Js : B(z,3ρe ) ⊂ B(y, 3)}.Wehave † ≤ { −s −s ∩  ∅} λE,x,s(B(y, )) λE,x,s(B(z,3ρe )) : B(z,3ρe ) B(y, ) = ∈ z Js −s ≤ λE,x,s(B(z,3ρe )) z∈J y,s ≤ e(2−δ)s(3ρ)2e−2s z∈J y,s ≤ 2−δ PS −s c0(3ρ) μH(y)(B(z,3ρe )) z∈Jy,s ≤ 2−δ PS κc0(3ρ) μH(y)(B(y, 3)) ≤ δ 2 2−δ δ 3 κc0(3ρ)  .

Hence for all 0 <<2ρ and y ∈ supp(λE,x,s), † ≤ δ λE,x,s(B(y, )) c1

for some constant c1 > 0 independent of s  1.  Therefore by Lemma 5.4, we deduce: Corollary 5.9. For any 0 <α<δ, † ∞ sup Iα(λE,x,s) < . s1 5.4. L2-convergence of projected measures. τ τ Recall the notation σx,θ,si and σx,θ from Definition 5.1. Theorem 3.3 is used crucially in the following proposition: Proposition 5.10. Fix 0 <τ≤ ρ,aPL-direction θ ∈ M for (x, τ) and a sequence s → +∞.Ifsup D(στ ) < ∞,then i i x,θ,si 2,r

2 τ −L−−−−(xVθ→) τ →∞ D(σx,θ,si ) D(σx,θ) as i . ± Proof. By Theorem 3.3 and the assumption of x ∈ Λ(Γ), λE,x,s | ρ τ weakly i xUθ Vθ PS τ τ converges to μ | ρ τ as si →∞. Therefore, σ weakly converges to σ H(x) xUθ Vθ x,θ,si x,θ as i →∞. Hence, it suffices to show that the collection { τ ∈ 2 τ } D(σx,θ,si ) L (xVθ ) 2 τ is relatively compact in L (xVθ ). Since this collection is uniformly bounded in the r τ Sobolev space H (xVθ ) by the assumption, the claim follows from the fact that r τ 2 τ  H (xVθ ) embeds compactly in L (xVθ ) for any r>0 (see [16, Theorem 16.1]). Recall that by a measure on [0, 2π), we mean the Lebesgue measure normalized to be the probability measure.

License or copyright restrictions may apply to redistribution; see https://www.ams.org/journal-terms-of-use 552 AMIR MOHAMMADI AND HEE OH

Theorem 5.11. Let si → +∞ be a fixed sequence. For any ε>0 and any finite subset {τ1,...,τn} of (0,ρ], there exists a Borel subset Θε(x) ⊂ [0, 2π), of measure at least 1 − ε, such that

(1) every θ ∈ Θε(x) is a PL direction for (x, τ) for each 1 ≤  ≤ n; ∈ { } (2) for each θ Θε(x), there exists an infinite subsequence sji (depending on (x, θ)) such that for each 1 ≤  ≤ n,

L2(xV ) D(στ ) −−−−−θ→ D(στ ). x,θ,sji x,θ Proof. Recall that we fixed some 0 1 such that sup D(στ ) 2 dθ ≤ L for 1 ≤  ≤ n. x,θ,si 2,r i Hence using Corollary 5.5 and Chebyshev’s inequality, we deduce that for any ε>0, there exists some L0 > 0 such that if we let Θτ = {θ : θ is a PL direction for (x, τ )and D(στ ) 2 0, we have m(Θτ ) > 1− ε . Let Θ = ∩ Θτ and let Θ = lim sup Θ . si 2n i  si i i − ∈ ∈ Then m(Θ) > 1 ε.Forθ Θ, θ lies in infinitely many of Θi’s, i.e., θ Θji for some infinite subsequence {ji}. Hence, the claim follows from Proposition 5.10 { }  applied to sji . † 5.5. Key lemma on the projections of λE,x,s. The following is the key technical lemma in the proof of the window theorem.

Lemma 5.12 (Key Lemma). Fix 0 <τ<ρandasequencesi → +∞. For any ε>0, there exists a Borel subset Θε(x) ⊂ [0, 2π) of measure at least 1 − ε such that ∈ −1 ⊂ if θ Θε(x) and Eimθ X is a sequence of Borel subsets satisfying − −1 → λE,x,si (xNτ Eimθ ) 0, { } ⊂ then there is an infinite subsequence sji such that for any Borel subset Oθ(x) ρ xVθ , lim sup σρ (O (x)−p (E m−1 ∩ xN )) ≤ σρ {t ∈ O (x):D(στ )(t)=0}. x,θ,sj θ θ i θ τ x,θ θ x,θ i i By Theorem 5.11, the Key Lemma follows from the following lemma. Observe that this is a rather strong control on the conditional measures, as one can easily construct counter-examples in a general setting. Here our L2- convergence result of the projection measures to a “rich” measure is crucially used. Lemma 5.13. Fix 0 <τ<ρand a PL direction θ ∈ [0, 2π), simultaneously for −1 ⊂ { } (x, τ) and (x, ρ).LetWimθ xNτ be a sequence of Borel subsets and si be a sequence tending to infinity. Assume the following holds as i →∞: L2(xV ) (1) D(στ ) −−−−−θ→ D(στ ); x,θ,si x,θ L2(xV ) (2) D(σρ ) −−−−−θ→ D(σρ ); x,θ,si x,θ − −1 → (3) λE,x,si (xNτ Wimθ ) 0. ⊂ ρ Then for any Borel subset Oθ(x) xVθ , lim sup σρ (O (x)−p (W m−1)) ≤ σρ {t ∈ O (x):D(στ )(t)=0}. x,θ,si θ θ i θ x,θ θ x,θ i

License or copyright restrictions may apply to redistribution; see https://www.ams.org/journal-terms-of-use BURGER-ROBLIN 553

Pτ { ∈ τ } Proof. Set := t Oθ(x):D(σx,θ)(t) > 0 and Lτ τ ∩ −1 ∩ −1 i := pθ(Oθ(x)Uθ Wimθ )=Oθ(x) pθ(Wimθ ).

For n0 > 1, define ∈Pτ τ ≥ 1 ρ Σn0 = t : D(σx,θ)(t) ,D(σx,θ)(t)

Let ε>0 be arbitrary. There exists n0 = n0(ε) > 1 such that ρ − ρ Pτ σx,θ(Σn0 ) > (1 ε)σx,θ( ). Since D(στ ) → D(στ )inL2(xV ), denoting by λ the Lebesgue measure on x,θ,si x,θ θ xVθ,wehave | τ −Lτ − τ −Lτ |≤ | τ − τ | σx,θ(Oθ(x) i ) σx,θ,si (Oθ(x) i ) D(σx,θ) D(σx,θ,si ) dλ xVθ ≤ τ − τ · 1/2 → D(σx,θ) D(σx,θ,si ) 2 λ(xVθ) 0. Since στ (O (x)−Lτ ) ≤ λ (xN − W m−1) → 0 by the assumption on x,θ,si θ i E,x,si ρ i θ Wimθ, it follows now that there is some i0 = i0(n0) such that for all i ≥ i0, τ −Lτ ε σx,θ(Oθ(x) i ) < 2 . n0 ⊂ τ ε ρ Note that for any set Υ Σn0 with σx,θ(Υ) < 2 ,wehaveσ (Υ) <ε.Tosee n0 x,θ this, note that if 1 ≤ τ ≤ ε λ(Υ) D(σx,θ,t)dλ(t) 2 , n0 Υ n0 ≤ ε then λ(Υ) n and hence 0 ρ ρ ≤ ≤ ε ≤ σx,θ(Υ) = D(σx,θ)(t)dλ(t) n0λ(Υ) ε. Υ n0 Therefore, we have ρ −Lτ ≤ ρ −Lτ ∩ ρ −Lτ ∩ − σx,θ(Oθ(x) i ) σx,θ((Oθ(x) i ) Σn0 )+σx,θ((Oθ(x) i ) (Oθ(x) Σn0 )) ≤ · ρ Pτ ρ −Pτ ε + ε σx,θ( θ )+σx,θ(Oθ(x) θ ). Since ε>0 is arbitrary, ρ −Lτ ≤ ρ −Pτ (5.3) lim sup σx,θ(Oθ(x) i ) σx,θ(Oθ(x) θ ). i Now since D(σρ ) → D(σρ )inL2(xV ρ), we have x,θ,si x,θ θ |σρ (O (x)−Lτ ) − σρ (O (x)−Lτ )|≤ |D(σρ ) − D(σρ )|dλ → 0. x,θ θ i x,θ,si θ i x,θ,si x,θ xVθ Combined with (5.3), this implies that lim sup σρ (O (x)−Lτ ) ≤ σρ (O (x) −Pτ ). x,θ,si θ i x,θ θ θ i 

License or copyright restrictions may apply to redistribution; see https://www.ams.org/journal-terms-of-use 554 AMIR MOHAMMADI AND HEE OH

6. Recurrence properties of BMS and BR measures 6.1. Theorems of Marstrand on Hausdorff measures. Let Λ ⊂ R2.The s-dimensional Hausdorff measure of Λ is defined to be Hs Hs (Λ) = inf η(Λ), η↓0 Hs { s ⊂∪∞ ≤ } where η(Λ) := i d(Wi) :Λ i=1Wi,d(Wi) η and d(Wi) denotes the diameter of Wi. The Hausdorff dimension of Λ is dim(Λ) = sup{s : Hs(Λ) > 0} =inf{s : Hs(Λ) = ∞}. A set Λ is called an s-set if 0 < Hs(Λ) < ∞. Following Marstrand [19], a point ξ ∈ Λ is called a condensation point for Λ if ξ is a limit point from (ξ,θ) ∩ Λfor almost all θ where (ξ,θ) denotes the ray through ξ lying in the direction θ. Let Λ be an s-set in the following three theorems: Theorem 6.1. [19, Theorem 7.2] If s>1, Hs-almost all points in Λ are conden- sation points for Λ. Theorem 6.2. [19, Lemma 19] If s>1, almost every lines L through Hs-almost all points in Λ intersect Λ in a set of dimension s − 1. Theorem 6.3. [19, Theorem II] If s ≤ 1, then the projections of Λ have Hausdorff dimension s for almost all directions. 6.2. U-conservativity of mBR. In the rest of this section, we assume that Γ is convex cocompact. ∈ PS Theorem 6.4. [35] For x G, the measure μH+(x) on xN is a δ-dimensional Hausdorff measure supported on the set {xn ∈ H+(x):(xn)+ ∈ Λ(Γ)}. Further- more, this is a positive and locally finite measure on xN. { 10 ∈ R} For U = ut = t 1 : t , we recall the definition of a conservative action: Definition 6.5 (Conservative action). Let μ be a locally finite U-invariant measure on X.TheU-action on X is conservative for μ if one of the following equivalent conditions holds: (1) for every positive Borel function ψ of X,

ψ(xut)dt = ∞ for a.e. x ∈ X; t∈R (2) for any Borel subset B of X with μ(B) > 0,

χB(xut)dt = ∞ for a.e. x ∈ B; t∈R The following is Maharam’s recurrence theorem (cf. [1, 1.1.7]).

Lemma 6.6. If there is a measurable subset B ⊂ X with 0 1,thenU is conservative for mBR.

License or copyright restrictions may apply to redistribution; see https://www.ams.org/journal-terms-of-use BURGER-ROBLIN 555

BMS BR Proof. Recall the notation Ω = supp(m )andΩBR = supp(m ). Set − F := {x ∈ X : x ∈ Λ(Γ),xut ∈/ Ω for all large t x 1}.

+ Hence x ∈F means (xut) ∈/ Λ(Γ) for all large t x 1. We claim that

(6.1) mBR(F)=0.

BR Suppose not. Then by the Fubini theorem, there is a set O⊂ΩBR with m (O) > 0 such that for all x ∈O, xmθ ∈Ffor a positive measurable subset of θ’s. Note − that νo({x ∈ Λ(Γ) : x ∈O}) > 0whereνo is the PS measure on Λ(Γ). Fix 3 2 ∈ H −{ } R | 3 ξ0 / Λ(Γ) and identify ∂( ) ξ0 with .Sinceνo ∂(H )−{ξ0} is equivalent to the δ-dimensional Hausdorff measure Hδ on Λ(Γ) ⊂ R2 by Theorem 6.4, we have δ − + 2 3 H {x ∈ Λ(Γ) : x ∈O}> 0. Note that Lθ(x):={(xmθut) ∈ R = ∂(H ) −{ξ} : t ≥ 0} is the line segment connecting x+ (at t =0)andx− (at t = ∞). Hence x ∈O − implies that x is not a limit point of the intersection Lθ(x) ∩ Λ(Γ) for a positive set of directions θ. This contradicts Theorem 6.1 and proves the claim (6.1). Let O be an r-neighborhood of Ω for some small r>0. If x ∈ X −F,then ∈ ∈O | | ∈ xut Ωand xut+s for all s

6.3. Leafwise measures. Let W be a closed connected subgroup of N.Let M∞(W ) denote the space of locally finite measures on W with the smallest topology ∗ so that the map ν → ψdνis continuous for all ψ ∈ Cc(W ) (the weak topology). A locally finite Borel measure μ on X gives rise to a system of locally finite mea- W ∈M sures [μx ] ∞(W ), unique up to normalization, called the leafwise measures or conditional measures on W -orbits. There is no canonical way of normalizing these W ∩ measures. For our purpose here, we fix a normalization so that μx (N1 W )=1. → W With this normalization, the assignment x μx , is a Borel map, furthermore, for a  W W ∈  full measure subset X of X, μxu = u.μx for every x, xu X ; for a comprehensive account on leafwise measures we refer the reader to [7]. PS PS Leb InthecasewhenW = N,wehaveμH(x) = μx and μH(x) = λx, which are precisely the N-leafwise measures of the BMS and BR measures respectively, up to normalization. We will be considering the U leafwise measures of mBMS as well as BR of μE,s. We will use the following simple lemma.

Lemma 6.8. Let μ be a locally finite M-invariant measure on X. For any 0 < τ  1,andany0 ≤ θ<π, we have τ U τ μU = μ θ xmθ x for μ a.e. x ∈ X.

Proof. Since μ is M-invariant, for Tρ = AρNˇρMρ,wehave

U τ − − | θ | τ 1 1 μx μ(xmθU mθ (mθVρTρmθ )) τ = lim =1. |μU | ρ→0 μ(xm U τ (V T )) xmθ θ ρ ρ 

License or copyright restrictions may apply to redistribution; see https://www.ams.org/journal-terms-of-use 556 AMIR MOHAMMADI AND HEE OH

6.4. Recurrence for mBMS. Since the frame flow is mixing by Theorem 2.9 with respect to mBMS,wehave: Proposition 6.9. For any non-trivial a ∈ A, mBMS is a-ergodic. ∈ BMS U Theorem 6.10. Let δ>1. For a.e. x Ω, (m )x is atom-free. F { ∈ BMS U } Proof. Setting := x Ω:(m )x has an atom , we first claim that mBMS(F) = 0. Suppose not. Fix any non-trivial a ∈ A.SinceU is normal- ized by a, F is a-invariant. Hence mBMS(F) = 1 by Proposition 6.9. Using the Poincare recurrence theorem, it can be shown that F  { ∈ BMS U } := x Ω:(m )x is the dirac measure at e has a full measure in Ω (cf. [13], [15, Theorem 7.6]). ∈ PS { ∈ Since for any x Ω, μx is a positive δ-dimensional Hausdorff measure on xn + ∈ } BMS U PS U ∈ H(x):(xn) Λ(Γ) by Theorem 6.4 and (m )x =(μH(x))x for a.e. x Ω, ∈ PS U it follows that for a.e. x Ω, (μH(x))x is the Dirac measure at e. By the Fubini ∈ ⊂ PS theorem, there exist x Ω and a measurable subset D0 H(x) with μH(x)(D0) > 0 such that for each y ∈ D ,(μPS )U is the dirac measure at ym for a positive 0 H(ymθ ) ymθ θ s measurable subset of mθ’s. For s ≥ 0, denote by H the s-dimensional Hausdorff −{ −} Hδ PS R2 measure on Λ(Γ) x ;so = μH(x). In the identification of H(x) with via → +  ⊂ −{ −}⊂R2 the map y y , this implies that there is a subset D0 Λ(Γ) x with Hδ  ∈  (D0) > 0 such that for all ξ D0, there is a positive measurable subset of lines L through ξ such that 0 < H0((Λ(Γ) −{x−}) ∩ L) < ∞. This contradicts Theorem 6.2 which implies that (Λ(Γ) −{x−}) ∩ L has dimension δ − 1 > 0 for almost all lines L through ξ.  Corollary 6.11. If δ>1, mBMS is U-recurrent, i.e., for any measurable subset B of X, {t : xut ∈ B} is unbounded for a.e. x ∈ B. BMS U Proof. By [7, Theorem 7.6], Theorem 6.10 implies that (m )x is infinite for a.e. x. [7, Theorem 6.25] implies the claim.  BMS U | BMS| 6.5. Doubling for the (μ )x . As before, we assume m =1.SinceΩisa compact subset, we have 1 (6.2) ρ := inf{ρ : x ∈ Ω} > 0. 2 x Fix a small number ε>0. It follows from Theorem 6.10 that there exist 0 <   ⊂ BMS  − ε2 β = β(ε) ρ and a compact subset Ωε Ω with m (Ωε) > 1 2 such that 1 (6.3) (mBMS)U [−3β,3β] < (mBMS)U [−(ρ − β),ρ− β] for all x ∈ Ω . x 2 x ε

Since the covering {xBτ : x ∈ Ω,τ > 0} admits a disjoint subcovering of Ω ∈  with full BMS measure (see [20, Theorem 2.8]), there exist x0 = x0(ε) Ωε and 0 <τ<β(ε) such that for B (τ):=x Nˇ A MN , x0 0 τ τ τ ε2 (6.4) mBMS(B (τ) ∩ Ω ) > 1 − · mBMS(B (τ)). x0 ε 2 x0

∈  We fix x0 Ωε and τ>0 for the rest of this section.

License or copyright restrictions may apply to redistribution; see https://www.ams.org/journal-terms-of-use BURGER-ROBLIN 557

ˇ Recall the notation Tτ = Nτ Aτ M, so that Bx0 (τ)=x0Tτ Nτ .Setν = νx0Tτ for simplicity. Using Theorem 6.10, we will prove:

Theorem 6.12. Let δ>1.Letc0 > 1 be as in Theorem 4.1 where F0 is the 2ρ- PS ⊂ neighborhood of Ω. Then there exists a Borel subset Ξε (x0) x0Tτ which satisfies the following properties: PS ≥ − 4 · (i) ν(Ξε (x0)) (1 c0 ε)ν(x0Tτ ); ∈ PS (ii) for any xmθ Ξε (x0), with θ a PL direction for (x, τ), there exists a Borel { ∈ τ } subset Oθ(x) of the set t xVθ : D(σx,θ)(t) > 0 such that

δ PS ρ ≥ PS 2τ ≥ τ μx (Oθ(x)Uθ ) 2μx (Oθ(x)Uθ ) . 4c0 Despite the rather complicated formulation of this theorem, which is tailored

towards our application later, the theorem is intuitively clear. Indeed, Bx0 (τ)is chosen so that for “most” BMS points, we have (6.3). On the other hand, in view τ of Corollary 5.5, for a PL direction θ, the Radon-Nikodym derivative D(σx,θ)is τ positive σx,θ-almost everywhere. Therefore, by Fubini’s theorem, for “most” BMS ∈ points x Bx0 (τ), “most” points in xNρ satisfy both (6.3) and the non-vanishing of the Radon-Nikodym derivative implies Theorem 6.12. The precise treatment of the above sketch of the proof is given in the rest of this subsection. ∈  ⊂ ∩  −1 Lemma 6.13. Let xmθ x0Tτ . For any Borel subset Oθ(x) pθ(xNτ Ωεmθ ), PS  ρ ≥ PS  2τ μx (Oθ(x)Uθ ) 2μx (Oθ(x)Uθ ). ∈ ∩  −1 ∈ −1 Proof. If t pθ(xNτ Ωεmθ ) and hence t = xvθ for vθ Vθ where ztmθ = xvθuθ ∈  −1 −1 ∈ τ for zt Ωε, we can write it as tmθ = ztmθ uθ mθ = ztu0 for some u0 U . Hence, by (6.3),

ρ ρ−τ 3τ 2τ |μU |≥|μU |≥2|μU |≥2|μU |. tmθ zt zt tmθ  ⊂ ∩  −1 Using this and since Oθ(x) pθ(xNτ Ωεmθ ), we have  U ρ ρ μPS(O (x)U ρ)= |μ θ |dσρ (t)= |μU |dσρ (t) x θ θ t x,θ tmθ x,θ O (x) O (x) θ θ 2τ U 2τ ≥ | U | ρ | θ | ρ 2 μtm dσ (t)=2 μt dσ (t) θ x,θ x,θ Oθ (x) Oθ (x) PS  2τ =2μx (Oθ(x)Uθ ). 

⊂  BMS − 2 Lemma 6.14. There exists a compact subset Ωε Ωε with m (Ωε) > 1 ε such that BMS ∩ − 2 · BMS m (Bx0 (τ) Ωε) > (1 ε ) m (Bx0 (τ)) and that ∩ −1 ⊂{ ∈ τ } pθ(xNτ Ωεmθ ) t xVθ : D(σx,θ)(t) > 0

for all xmθ ∈ x0Tτ with θ a PL-direction for (x, τ).

License or copyright restrictions may apply to redistribution; see https://www.ams.org/journal-terms-of-use 558 AMIR MOHAMMADI AND HEE OH

 { ∈ τ } Proof. Setting xVθ := t xVθ : D(σx,θ,t)=0 ,wehave BMS ∪  τ PS  τ m ( xmθ ∈x0Tτ xVθ Uθ )= μx (xVθ Uθ )dν(xmθ) ∈ xmθ x0Tτ U τ | PS θ | τ = (μx )x dσx,θ(t)dν(xmθ)=0. ∈ ∈ xmθ x0Tτ t xVθ O Hence, there exists an open subset ε of Bx0 (τ) which contains the subset  τ BMS ε2 BMS  ∪ ∈ O ≤ · −O xmθ x0Tτ xVθ Uθ and m ( ε) 2 m (Bx0 (τ)). Now set Ωε := Ωε ε.It is easy to check that this Ωε satisfies the claim. 

We set PS { ∈ PS ∩ − PS } Ξε (x0):= x x0Tτ : μx (xNτ Ωε) > (1 ε)μx (xNτ ) . Lemma 6.15. We have PS ≥ − 4 · ν(Ξε (x0)) (1 c0 ε)ν(x0Tτ ).

μPS(x N ) μPS(x N ) x0 0 τ x0 0 τ + ∈ Proof. Set b1 := infx∈x0Tτ PS and b2 := supx∈x T PS .Sincex μx (xNτ ) 0 τ μx (xNτ ) 0 + + + ∈ Λ(Γ) and x0 = x ,wehavex Λ(Γ) and hence it follows from Theorem 4.1 that − c 4 < b2

By the M-invariance of mBMS, by Lemma 6.8, μPS (Ω ∩ xm N )=μPS (Ω ∩ xN m )=μPS(Ω m−1 ∩ xN ). xmθ ε θ τ xmθ ε τ θ x ε θ τ ∈ PS Note that for any xmθ Ξε (x0), we have δ PS ∩ −1 − ≥ τ μx (xNτ Ωεmθ ) > (1 ε)μx(xNτ ) . 2c0

License or copyright restrictions may apply to redistribution; see https://www.ams.org/journal-terms-of-use BURGER-ROBLIN 559

∩ −1 2τ ∩ By setting Oθ(x):=pθ(xNρ Ωεmθ ), we note that Oθ(x)Uθ contains xNτ −1  Ωεmθ and hence for all small 0 <ε 1, δ PS 2τ ≥ − PS ∩ −1 ≥ τ μx (Oθ(x)Uθ ) (1 ε)μx (xNτ Ωεmθ ) . 2c0 Therefore the above three lemmas prove Theorem 6.12.

7. Window theorem for Hopf average We will combine the results from previous sections and prove the window theorem 7.7 in this section. We first show that the disintegration along U of λs, notation as in Section 3, has certain doubling properties, see Theorem 7.1. This is done by applying results in Section 5, in particular the key lemma, to λE,x,s and the PS limiting measure μx , in combination with Theorem 6.12, which gives a rather strong doubling property for the disintegration of the PS measure. As we mentioned in the introduction, in general, the weak* convergence of measures does not give control on the corresponding conditional measures, e.g., one should recall the well- known discontinuity of the entropy. However, here the key lemma gives a good PS control both on the prelimiting measures λE,x,s and the limit measure μx , and helps us to draw some connection between the conditionals. In order to obtain the Window Theorem 7.3, we flow by a−s for a suitably big − − BR s and bring [ T,T]tosize[ ρ, ρ]. We are now working with mE,s rather than 1 BR| mBR(E) m E, and the desired estimate follows from Theorem 7.1

7.1. Window theorem for χE. Let Γ be a convex cocompact subgroup with δ>1. Let E be a BMS box. For simplicity, we set BR μs := μE,s and λx,s = λE,x,s defined in section 5. For 0 0 as in (6.2), we put U − (μs)x [ 2ρr, 2ρr] E (r):= x ∈ Ea− : > 1 − r . s s U − (μs)x [ 2ρ, 2ρ]

Theorem 7.1. There exist 0 1 (depending on E) such that for all s>s0, we have BR − μE,s(Es(r0)) < 1 r0.

Proof. Suppose not; then there are a subsequence ri → 0 and a subsequence si → +∞ such that μBR (E (r )) ≥ 1 − r .Set E,si si i i BR μi = μE,si and Ei := Esi (ri). ∈  PS Fix ε>0. Let x0 Ωε,0<τ<ρ, c0 > 1andΞε (x0) be as in Theorem 6.12. BMS Set q0 := m (Bx0 (τ)) > 0and ∈ ∩ − 2ri x0Ti := y x0Tτ : λsi,y(Ei yNτ ) > 1 λsi,y(yNτ ) . q0

Recall the measure ν = νx0Tρ from Theorem 6.12. We claim that for all large i  1, we have √ (7.1) ν(x0Ti) ≥ (1 − 4 ri)ν(x0Tτ ).

License or copyright restrictions may apply to redistribution; see https://www.ams.org/journal-terms-of-use 560 AMIR MOHAMMADI AND HEE OH

We will first show that for all large i  1, ∩ ≥ − 2ri (7.2) μi(Ei Bx0 (τ)) 1 μi(Bx0 (τ)). q0 If this does not hold, by passing to a subsequence, we have that − ∩ − 1 ri <μi(Ei)=μi(Ei Bx0 (τ)) + μi(Ei Bx0 (τ)) ≤| |−2ri μi μi(Bx0 (τ)). q0

Since |μi| = 1, it follows that 2 ≤ μi(Bx0 (τ)) 1. q0 BMS On the other hand, since μi weakly converges to m by Theorem 3.1, we have μ (B (τ)) i x0 → 1, q0 which gives a contradiction. This shows (7.2). Now, by the same type of argument as the proof of Lemma 6.15, we can show (7.2) implies (7.1). Passing√ to a subsequence, which we continue to denote by ri, we assume that 4 i ri <ε/2and2ri/q0 <εfor all i.Ifweset ∗ Ξ (x0):=∩ix0Ti, then it follows that ∗ ν(Ξ (x0)) > (1 − ε)ν(x0Tτ ). Hence for all sufficiently small ε>0, PS ∩ ∗ − 4 ν(Ξε (x0) Ξ (x0)) > (1 (1 + c0)ε) ν(x0Tτ ) > 0.

Let Θε(x) be given as in the Key Lemma 5.11 for {ρ, τ} applied to the set − E and the sequence si. Since supp(ν) ⊂{x ∈ X : x ∈ Λ(Γ)}, we can find ∈ ∗ ∩ PS − − ∈ xmθ Ξ (x0) Ξε (x0) (depending on ε>0) with (xmθ) = x Λ(Γ) and ∈ + + ± ∈ θ Θε(x). Since x = x0 ,wehavex Λ(Γ). ∈ By the M-invariance of the measure μsi and as xmθ x0Ti,wehave −1 λx,s (Eim ∩ xNτ )=λxm ,s (Ei ∩ xmθNτ ) i θ θ i 2r ≥ 1 − i λ (xm N ) q xmθ ,si θ τ 0 − 2ri = 1 λx,si (xNτ ), q0 and hence − −1 → (7.3) λx,si (xNτ Eimθ ) 0. { } Let sji be the corresponding subsequence given by Lemma 5.11 depending on

(x, θ). By passing to that subsequence, we set si := sji . τ ∩ −1 For Oθ(x) as in Theorem 6.12, we consider Li := pθ(Oθ(x)Uθ Eimθ ). By Lemma 7.2 below, ρ ≤ 1 2τ λx,si (LiUθ ) λx,si (LiUθ ). 1 − ri

License or copyright restrictions may apply to redistribution; see https://www.ams.org/journal-terms-of-use BURGER-ROBLIN 561

−1 ∩ { ∈ τ } ∅ Since Li = pθ(Eimθ ) Oθ(x)and t Oθ(x):D(σx,θ,t)=0 = , it follows from the key Lemma 5.12 and (7.3) that for all large i  1, (7.4) λ (O (x)U ρ)=σρ (O (x)) x,si θ θ x,θ,si θ (7.5) ≤ (1 + ε)σρ (L )=(1+ε)λ (xL U ρ). x,θ,si i x,si i θ Therefore, we have

ρ ≤ (1 + ε) 2τ λx,si (Oθ(x)Uθ ) λx,si (Oθ(x)Uθ ) (1 − ri) ≤ 2τ (1+2ε)λx,si (Oθ(x)Uθ ). { } Recall we chose si = sji so that Theorem 5.11 holds for ρ, τ . Hence, by sending si →∞, the above implies PS ρ ≤ PS 2τ μx (Oθ(x)Uθ ) (1+2ε)μx (Oθ(x)Uθ ). Together with Theorem 6.12, this gives · PS 2τ ≥ PS 2τ 2ε μx (Oθ(x)Uθ ) μx (Oθ(x)Uθ ), PS 2τ τ δ however, μ (Oθ(x)U ) ≥ > 0. This gives a contradiction and finishes the x θ 8c0 proof of Theorem 7.1. 

Lemma 7.2. Let xmθ ∈ x0Tτ for x0 ∈ Ω and s>0. For any Borel subset ⊂ ∩ −1  Lθ(x) pθ(xNτ Es(r)mθ ) and for all sufficiently small 0

U 2ρ 1 U 2rρ |(μs) |≤ |(μs) | zt 1 − r zt and hence

U ρ U ρ+τ U 2ρ 1 U 2rρ 1 U 2rρ+τ |(μs) |≤|(μs) |≤|(μs) |≤ |(μs) |≤ |(μs) |. tmθ zt zt 1 − r zt 1 − r tmθ Therefore, since μBR is M-invariant, by Lemma 6.8, E,s U ρ ρ | θ | ρ λx,s(Lθ(x)Uθ )= (μs)t dσs,x,θ(t) ∈ t Lθ (x) ρ = |(μ )U | dσρ (t) s tmθ x,θ,s ∈ t Lθ (x) 1 2τ ≤ |(μ )U | dσρ (t) s tmθ x,θ,s 1 − r ∈ t Lθ (x) 1 U 2τ = |(μ ) θ | dσρ (t) − s t x,θ,s 1 r t∈Lθ (x) 1 = λ (L (x)U 2τ ). 1 − r x,s θ θ 

License or copyright restrictions may apply to redistribution; see https://www.ams.org/journal-terms-of-use 562 AMIR MOHAMMADI AND HEE OH

We deduce the following from Theorem 7.1:

Theorem 7.3. There exists 0 1, depending on E, such that for all T>T0 rT T BR BR m x ∈ E : χE(xut)dt < (1 − r) · χE(xut)dt ≥ r · m (E). −rT −T Proof. Setting ⎧ ⎫ ⎨ 2rρes ⎬ − s χE(xut)dt E(s, r):= x ∈ E : 2rρe ≥ 1 − r , ⎩ 2ρes ⎭ −2ρes χE(xut)dt it suffices to prove that for some 0

We note that (μBR )U [−2rρes, 2rρes] E(s, r)= x ∈ E : 0,E x ≥ 1 − r , BR U − s s (μ0,E )x [ 2ρe , 2ρe ]

BR BR U BR m |E where (μE,0)x denotes the leafwise measure of μE,0 = mBR(E) . Note that (7.7) follows from Theorem 7.1 if we show BR BR · BR m (E(s, r)) = m (E) μE,s(Es(r)) for all 0

Hence a−sE(s, r)coincideswithEs(r), up to a BR null set. This implies the claim BR  using the definition of μE,s. 7.2. Ergodic decomposition and the Hopf ratio theorem. In this subsection, let μ be a locally finite U-invariant conservative measure on X.LetM∞(X)denote the space of locally finite measures on X with weak∗ topology. Let A denote a countably generated σ-algebra equivalent to the σ-algebra of all U-invariant subsets of X.ThereexistaA-measurable conull set X of X, a family { A ∈ } μx = μx : x X of conditional measures on X and a probability measure μ∗ on X which give rise to the ergodic decomposition of μ:

μ = μx dμ∗(x)

License or copyright restrictions may apply to redistribution; see https://www.ams.org/journal-terms-of-use BURGER-ROBLIN 563

 where the map X →M∞(X), x → μx, is Borel measurable, μx is a U-invariant, ergodic and conservative measure on X and for any ψ ∈ L1(X, μ),

μ(ψ)= μx(ψ) dμ∗(x); x∈X see [7, 5.1.4]. The following is the Hopf ratio theorem in a form convenient for us ([11], see also [38]). Theorem 7.4. Let ψ, φ ∈ L1(mBR) with φ ≥ 0. Furthermore suppose that ψ and φ are compactly supported. Then T 0 ψ(xu )dt − ψ(xut)dt μ (ψ) 0 t T x lim T = lim 0 = T T μx(φ) 0 φ(xut)dt −T φ(xut)dt ∈{ ∈ T } 2 for μ-a.e. x x X :supT 0 φ(xut)dt > 0 . Lemma 7.5. Fix a compact subset E ⊂ X with μ(E) > 0.Letφ be a non-negative compactly supported Borel function on X such that φ|E > 0. For any ρ>0 there exists a compact subset Eρ(φ) ⊂ E with μ(E − Eρ(φ)) <ρ· μ(E) satisfying:

(1) the map x → μx is continuous for all x ∈ Eρ(φ);

(2) infx∈Eρ μx(φ) > 0; (3) for any ψ ∈ Cc(X) the convergence T ψ(xu )dt μ (ψ) 0 t → x T μx(φ) 0 φ(xut)dt

is uniform on Eρ(φ).  ⊂ −  Proof. By Lusin’s theorem, there exists a compact subset E E with μ(E E ) < ρ →  T → ∞ 3 μ(E)andthemapx μx is continuous on E . Since 0 φ(xut)dt + for a.e. x ∈ E by the conservativity of μ,wehaveμx(φ) > 0 almost all x ∈ E.Since x → μx(φ) is a measurable map, it follows again by Lusin’s theorem that there  ⊂   −  ρ  ∞ ∞ exists a compact subset E E with μ(E E ) < 3 μ(E ), φ(xut)dt = for  0 all x ∈ E , and infx∈E μx(φ) > 0. We claim that for any ε>0 and any compact subset Q of X, there exists a    compact subset E0 = E0(Q, ε) ⊂ E such that μ(E − E0) <εμ(E )andforall ψ ∈ C(Q), the convergence T ψ(xu )dt μ (ψ) 0 t → x T μx(φ) 0 φ(xut)dt

is uniform on E0.LetB = {ψj } be a countable dense subset of C(Q) which includes the constant function χQ. We can deduce from the Hopf ratio Theorem 7.4 and   Egorov’s theorem that there is a compact subset E1 ⊂ E such that μ(E − E1) < ε μ(E), sup μ (Q) < ∞,andforeachψ ∈B, the convergence 3 x∈E1 x j T ψ (xu )dt μ (ψ ) 0 j t → x j (7.8) T μx(φ) 0 φ(xut)dt

2 note that for a.e. x in this set μx(φ) > 0, see [38, Page 3].

License or copyright restrictions may apply to redistribution; see https://www.ams.org/journal-terms-of-use 564 AMIR MOHAMMADI AND HEE OH

is uniform on E1. We will show the uniform convergence in E1 for all ψ ∈ C(Q). For any η>0, there exists ψj ∈Bsuch that ψj − ψ ∞ <η.Let T0  1besuch that T T ψ (xu )dt μ (ψ ) χ (xu )dt μ (χ ) 0 j t − x j ≤ 0 Q t − x Q ≤ T η, T η μx(φ) μx(φ) 0 φ(xut)dt 0 φ(xut)dt for all x ∈ E1 and T ≥ T0.Nowforanyx ∈ E1 and T ≥ T0,wehave T T ψ(xu )dt μ (ψ) |ψ(xu ) − ψ (xu )|dt 0 t − x ≤ 0 t j0 t T T φ(xut)dt μx(φ) φ(xut)dt 0 0 T ψj (xut)dt μ (ψ ) μ (ψ ) μ (ψ) + 0 0 − x j0 + x j0 − x T φ(xut)dt μx(φ) μx(φ) μx(φ) 0 T 0 χQ(xut)dt μx(Q) ≤ − ∞ − ∞ T ψ ψj + η + ψ ψj μx(φ) 0 φ(xut)dt μ (Q) μ (Q) ≤ x η + η2 + η + x η μx(φ) μx(φ) ≤ η(2a0 + η +1)

μx(Q) where a0 := sup ∈ < ∞. This proves the claim. Let Q1 ⊂ Q2 ⊂··· be an x E1 μx(φ) ∩ ρ exhaustion of X by compact sets. Then Eρ(φ):= iE0(Qi, 4i+1 ) satisfies all the desired properties. 

7.3. Window theorem for ψ ∈ Cc(X) with ψ|E > 0. Let Γ be a convex cocom- pact subgroup with δ>1. Let A denote a countably generated σ-algebra which is equivalent to the σ-algebra of all U-invariant subsets of X, as before. Since mBR is U-conservative by Theorem 6.7, we may write an ergodic decom- position BR BR m = μx dm∗ (x) x∈X  BR where X is a A-measurable conull set of X, m∗ is a probability measure on X, ∈  A and for all x X , μx = μx is a U-invariant ergodic conservative measure.

Lemma 7.6. Let E and 0 0. For any ρ>0,thereexistss ≥ 1 such that for all s>s , E 0 0 rs s BR BR m x ∈ E : ψ(xut)dt ≤ (1 − r + ρ) ψ(xut)dt ≥ (r − 2ρ) · m (E). −rs −s Proof. For simplicity, set rs s Fρ(s):= x ∈ E : ψ(xut)dt ≤ (1 − r + ρ) ψ(xut)dt . −rs −s ⊂ |  Let Eρ(χE) E be as in Lemma 7.5. Since ψ E > 0, there is a subset Eρ of Eρ(χE) BR  BR μx(ψ) such that m (E − E ) < 2ρ · m (E) and infx∈E > 0. Then for all large ρ ρ μx(E) s (uniformly for all x ∈ Eρ(χE)), s s μx(ψ) ψ(xut)dt = + ax(ψ, s) χE(xut)dt −s μx(E) −s

where |ax(ψ, s)|≤a(s) → 0ass →∞by Lemma 7.5.

License or copyright restrictions may apply to redistribution; see https://www.ams.org/journal-terms-of-use BURGER-ROBLIN 565

Setting rs s E˜(s, r)= x ∈ E : χE(xut)dt ≥ (1 − r) · χE(xut)dt , −rs −s we claim that  ∩ − ˜ ⊂ Eρ (E E(s, r)) Fρ(s) for all large s, ∈  ∩ − ˜ from which the lemma follows by Theorem 7.3. For any x Eρ (E E(s, r)), rs μ (ψ) rs ψ(xu )dt = x + a (ψ, rs) χ (xu )dt t μ (E) x E t −rs x −rs μ (ψ) s ≤ x + |a (ψ, rs)| (1 − r) χ (xu )dt μ (E) x E t x −s s s ≤ (1 − r) ψ(xut)dt +(|ax(ψ, s)| + |ax(ψ, rs)|)(1 − r) χE (xut). −s −s ≥ ∈  Let s1 > 1 be such that for s s1 and for all x Eρ, (|a (ψ, s)| + |a (ψ, rs)|)(1 − r) x x ≤ ρ; μx(ψ) | + ax(ψ, s)| μx(E)

this is possible since μx(ψ) is uniformly bounded from below by a positive number. μx(E) Then the claim holds. 

By taking ρ = r/4 and replacing 3r/4byr in the above lemma, we now obtain:

Theorem 7.7 (Window Theorem). Let ψ ∈ Cc(X) be a non-negative function such that ψ|E > 0.Then,thereexists0 1 such that for any T ≥ T0, rT T BR r BR m x ∈ E : ψ(xut)dt < (1 − r) ψ(xut)dt > · m (E). −rT −T 2

It is worth mentioning that r obtained here may be rather small. The following lemma demonstrates how the window estimates for a sequence will be used.

Lemma 7.8. Let ε>0 and a sequence sk → +∞ be given. Let E and ψ be as in Theorem 1.4.Fixρ>0.Letxk ∈ Eρ(ψ) be a sequence satisfying sk sk ψ(xkut)dt ≥ cε ψ(xkut)dt (1−ε)sk 0

for some c>0 independent of k. Then for any f ∈ Cc(X),ask →∞, sk − f(xkut)dt μ (f) (1 ε)sk ∼ xk sk . ψ(xkut)dt μxk (ψ) (1−ε)sk Proof. By the Hopf ratio theorem, and Lemma 7.5, we have s s s μxk (f) f(xkut)dt = ψ(xkut)dt + axk (s) ψ(xkut)dt 0 μxk (ψ) 0 0

License or copyright restrictions may apply to redistribution; see https://www.ams.org/journal-terms-of-use 566 AMIR MOHAMMADI AND HEE OH

with lim →∞ a (s)=0, uniformly in {x }. Therefore s xk k sk sk μxk (f) f(xkut)dt = ψ(xkut)dt (1−ε)s μxk (ψ) (1−ε)s k k sk (1−ε)sk − − + axk (sk) ψ(xkut)dt axk ((1 ε)sk) ψ(xkut)dt. 0 0 Since − sk (1 ε)sk a (s ) ψ(x u )dt − a ((1 − ε)s ) ψ(x u )dt xk k k t xk k k t 0 0 sk ≤| − |· axk (sk)+axk ((1 ε)sk) ψ(xkut)dt 0 | − | sk axk (sk)+axk ((1 ε)sk) − ψ(xkut)dt ≤ (1 ε)sk , c · ε we obtain that sk − f(xkut)dt μ (f) |a (s )+a ((1 − ε)s )| (1 ε)sk xk xk k xk k sk = + O . ψ(xkut)dt μxk (ψ) cε (1−ε)sk − → { }  Since axk (sk)+axk ((1 ε)sk) 0, uniformly in xk , the lemma follows. 8. Additional invariance and ergodicity of BR for δ>1 Let Γ be a convex cocompact subgroup with δ>1.  BR BR A ∗ 8.1. Reduction. Let ,X and m = x∈X μxd(m ) (x) be the decomposition of mBR into U-ergodic components, see Section 7.2. Our strategy in proving the U-ergodicity of mBR is to show that for a.e. x ∈ X, μx is N-invariant. Fix a BMS box E and a non-negative function ψ ∈ Cc(X) with ψ|E > 0. Let r ⊂ 0

Proof. Since N is abelian and U

By Lemma 8.2, the characteristic function χF is an N-invariant measurable func- tion. Since mBR is N-ergodic by Theorem 2.9 and mBR(F ) > 0 by Theorem 8.1, BR BR it follows that m (X − F ) = 0. That is, μx = m for a.e. x, and since μx’s are U-ergodic components of mBR, the claim follows. 

License or copyright restrictions may apply to redistribution; see https://www.ams.org/journal-terms-of-use BURGER-ROBLIN 567

8.2. Proof of Theorem 8.1. As we explained in the introduction, we will flow two nearby points in the generic set and study their divergence in the “interme- diate range”. We first need to prove a refinement of the window theorem, see Propositions 8.5 and 8.6 below. ∈ ∩ BR Fix x0 Er0 (ψ) supp(m ). Proposition 8.4. There is a Borel subset E ⊂ E such that mBR(E − E)=0and for any x ∈ E and all integers m ≥ 1, 1 xN ∩ B x , ∩ E (ψ) = ∅. 0 m ρ

BR Proof. Set Nk := {nz : |z|

1 Since inf ∈ μx(ψ) > 0andx → μx is continuous on Er (ψ), there x Er0 (ψ) μx(ψ) 0 exists a symmetric neighborhood O such that |μ (gψ)| |μ (gψ)| (8.1) 0 < inf x ≤ sup x < ∞. g∈O,x∈Er (ψ) μx(ψ) ∈O ∈ μx(ψ) 0 g ,x Er0 (ψ) O  ∩ ≥ Set Kψ := supp(ψ) and Kψ := g∈Osupp(ψ)g. By Theorem 7.7, for all s T0, 5r BR the following set has BR measure at least 16 m (E): E s rs s := x ∈ E (ψ) ∩ E (χ ) ∩ E (χ ): ψ(xu )dt < (1 − r) ψ(xu )dt . r0 r0 Kψ r0 Kψ t t −rs −s  Therefore, for each s ≥ T0, there exists a compact subset G(s)ofEs ∩ E with BR G r BR m ( (s)) > 8 m (E). We may write G(s)asG(s)+ ∪G(s)− where rs 1 − r s G(s) = x ∈G(s): ψ(xu )dt ≤ ψ(xu )dt ; + t 2 t 0 0 0 1 − r 0 G(s)− = x ∈G(s): ψ(xut)dt ≤ ψ(xut)dt . −rs 2 −s → ∞ BR G ≥ r Therefore, there exists an infinite sequence pi + such that m ( (pi)+) 16 BR G ≥ r for all i or m ( (pi)−) 16 for all i. BR G ≥ r Inthefollowing,weassumetheformercasethatm ( (pi)+) 16 for all i. The argument is symmetric in the other case.

Proposition 8.5. Fix integers , m > 1. There exists an infinite sequence sk = sk(, m) and elements xk = xk(, m),yk = yk(, m) ∈G(sk)+ which satisfy the following:

−1 −2 −1 −2 −1 |(wk)| (1) yk = xknˇw where c s  ≤|wk|≤c1s  and |(wk)|≥ k 1 k k c1 where c1 > 1 is independent of , ε, k.

License or copyright restrictions may apply to redistribution; see https://www.ams.org/journal-terms-of-use 568 AMIR MOHAMMADI AND HEE OH

(2) each x satisfies k sk r sk ψ(xkut)dt ≥ ψ(xkut)dt. (1−ε)sk 4m 0 Proof. If x ∈G (s), then, as r<1, + s s ψ(xut) ≥ r ψ(xut)dt. rs 0 (j−1) j 1 By subdividing [r, 1] into m subintervals Ii =(r + m ,r+ m )’s of length m , there exists an integer 1 ≤ j = j(x, s) ≤ m such that (r+j/m)s r s ψ(xut)dt ≥ ψ(xut)dt. (r+(j−1)/m)s 4m 0

Let d0 = d0(r/16) > 0 be as in Proposition 4.4. Applying Proposition 4.4 to G 2 ∈G each (pi)+ and a sequence (pi) , we can find xi,yi (pi)+ satisfying yi = xinˇwi −1 −2 −1 −2 −1 |wi| with d p  ≤|wi|≤d0p  and |(wi)|≥ . Choose a subsequence 0 i i d0 { } j0 xik of xi such that j(xik ,pik )isaconstant,say,j0. Setting sk := (r + m )pik , x := x and y = y ,wehave k ik k ik sk r sk ψ(xkut)dt ≥ ψ(xkut)dt (1−ε)sk 4m 0 ≤ ≤ 2  and rpik sk (r +1)pik . Hence, the claim follows with c1 = d0(r +1) .

We now use the fact that the two orbits xkut and ykut stay “close” to each other, for all t ∈ [0,sk], to show that yk’s in Proposition 8.5 also satisfy the same type of window estimate. Let us fix some notation; writing y u = x u (u− nˇ u ), we set k t k t t wk t 1+twk wk − pk(t):=u tnˇwk ut = 2 , −t wk 1 − twk

and gk = pk(sk).

Proposition 8.6. There are positive constants c2 = c2(ψ) and ε0 = ε0(ψ) such 1 that for all ε = <ε0 and all k  1, m sk sk ψ(ykut)dt ≥ c2 · ε ψ(ykut)dt, (1−ε)sk 0

where yk = yk(, ε) is as in Proposition 8.5. | −1| Proof. There is a constant c>0 (independent of ε) such that pk(t)gk 0, independent of ε, we have for all k  1, 1 sk sk (8.2) ψ(xkut)dt ≥ b1 ψ(ykut)dt. 0 0  By the definition of K and K ,sincey u ∈ x u O,wehaveχ (y u ) ≤ ψ ψ k t k t Kψ k t χK (xkut) for all t ∈ [0,sk]. In particular, we have ψ sk sk χ (x u )dt ≥ χ (y u )dt. Kψ k t Kψ k t 0 0

License or copyright restrictions may apply to redistribution; see https://www.ams.org/journal-terms-of-use BURGER-ROBLIN 569

On the other hand, we have sk sk sk μy (ψ) ψ(y u )dt = k χ (y u )dt + a (ψ, s ) χ (x u )dt; k t Kψ k t xk k Kψ k t μy (χK ) 0 k ψ 0 0 sk sk sk μxk (ψ) ψ(xkut)dt = χKψ (xkut)dt + axk (ψ, sk) χKψ (xkut)dt 0 μxk (χKψ ) 0 0 {| | | |} ≤ → →∞ with max axk (ψ, sk) , ayk (ψ, sk) a(sk) 0ask . μyk (ψ) μxk (ψ) As and are uniformly bounded from below and above, by the μy (χK ) μx (χK ) k ψ k ψ choice of xk and yk,thereexistsb>0 such that for all large k  1, sk sk sk sk 2 ψ(x u )dt ≥ b χ (x u )dt ≥ b χ (y u )dt ≥ b ψ(y u )dt k t Kψ k t Kψ k t k t 0 0 0 0 finishing the proof of Claim (1). Claim (2):Forsomeconstantb2 > 0, independent of ε, we have for all k  1, sk sk ≤ (8.3) χKψ (xkut)dt b2 ψ(xkut)dt. (1−ε)sk (1−ε)sk By Lemma 7.8 and its proof, we have sk sk μxk (χKψ ) (8.4) χKψ (xkut)dt = ψ(xkut)dt − μ (ψ) − (1 ε)sk xk (1 ε)sk sk 4(a(sk)+a(εsk)) + · ψ(xkut)dt. rε (1−ε)sk

μ (ψ) Since xk is uniformly bounded from above and below by positive constants, μxk (χKψ ) it suffices to take k large enough so that (a(sk)+a(εsk)) ≤ ε to finish the proof of Claim (2). We have sk sk ψ(ykut)dt = ψ(xkutpk(t))dt − − (1 ε)sk (1 ε)sk sk sk ≥ ψ(xkutgk)dt − |ψ(xkutpk(t)) − ψ(xkutgk)|dt. (1−ε)sk (1−ε)sk By (8.3), for all large k, sk sk | − | ≤ ψ(xkutpk(t)) ψ(xkutgk) dt cψε χKψ (xkut)dt − − (1 ε)sk (1 ε)sk sk ≤ cψb2ε ψ(xkut) (1−ε)sk

μxk (gkψ) where cψ is the Lipschitz constant of ψ.Sincegk ∈Oand hence is μxk (ψ) uniformly bounded from above and below, we can deduce that for some c>1, sk sk sk −1 (8.5) c ψ(xkut)dt ≤ ψ(xkutgk)dt ≤ c ψ(xkut)dt. (1−ε)sk (1−ε)sk (1−ε)sk

License or copyright restrictions may apply to redistribution; see https://www.ams.org/journal-terms-of-use 570 AMIR MOHAMMADI AND HEE OH

Therefore, the above estimates together with (8.2) imply that for all k large, sk sk −1 ψ(ykut)dt ≥ (c − cψb2ε) ψ(xkut)dt − − (1 ε)sk (1 ε)sk (c−1 − c b ε)rε sk ≥ ψ 2 ψ(x u )dt 4 k t 0 −1 sk b1(c − cψb2ε)rε ≥ ψ(ykut)dt. 4 0

b1r 1 Now the proposition follows with c2 = and ε0 = .  8c 2b2cψ c

We will now flow xk and yk for the period of time [(1 − ε)sk,sk]. By the con- struction of these points, these two pieces of orbits are almost parallel and they essentially differ by gk which is of size O(1). More importantly these “short” pieces of the orbits already become equidistributed. This will show that some ergodic component is invariant by a nontrivial element in N − U and the proof can be concluded from there using standard arguments. ∈ N 1 ∈ N ∈ Fix  .Letεi = i > 0fori .Wechoosesk(ε1,)andxk(ε1,),yk(ε1,) G(sk(ε1,))+ as in Proposition 8.5. Together with Proposition 8.6, there exists α1 > 0 independent of ε1 and k such that sk(ε1,) sk(ε1,) ψ(xk(ε1,)ut)dt ≥ α1ε1 ψ(xk(ε1,)ut)dt and (1−ε )s (ε ,) 0 1 k 1 sk(ε1,) sk(ε1,) ψ(yk(ε1,)ut)dt ≥ α1ε1 ψ(yk(ε1,)ut)dt. (1−ε1)sk(ε1,) 0 → By passing to a subsequence, we assume that xk(ε1,) xε1,, and hence 10 y (ε ,) → y ,andp (s (ε ,)) converges to n := ∈ N where k 1 ε1, k k 1 vε1, vε1, 1 | | 1 c1 (vε1,) ≤|vε ,|≤ and |(vε ,)|≥ . c1 1  1 c1 We proceed by induction: by dividing the interval [(1 − εi)sk(εi,),sk(εi,)] into subintervals of length εi+1 as in the proof of Proposition 8.5, we can find a sequence sk(εi+1,) and subsequences xk(εi+1,)ofxk(εi,)andyk(εi+1,)of yk(εi,) satisfying sk(εi+1,) sk(εi+1,) ψ(xk(εi+1,)ut)dt ≥ α1εi+1 ψ(xk(εi+1,)ut)dt; (1−ε )s (ε ,) 0 i+1 k i+1 sk(εi+1,) sk(εi+1,) ψ(yk(εi+1,)ut)dt ≥ α1εi+1 ψ(yk(εi+1,)ut)dt (1−εi+1)sk(εi+1,) 0 10 and p (s (ε ,)) converges to some element n := ∈ N where k k i+1 vεi+1, vεi+1, 1 |(v )| 1 ≤| |≤ c1 | |≥ εi+1, c  vεi+1,  and (vεi+1,) c . 1 →∞ → 1 → Clearly, as i ,wehavexk(εi,) xε1, and yk(εi,) xε1,. By passing ∈ to a subsequence, we may assume that vεi, converges to an element v N.Note | | 1 ≤| |≤ c1 | |≥ (v) that c  v  and (v) c . 1 1 ∈O Let 0 > 1 be large enough so that nv for all >0.

License or copyright restrictions may apply to redistribution; see https://www.ams.org/journal-terms-of-use BURGER-ROBLIN 571

∈ Proposition 8.7. Let >0 and set x := xε1,. For any f Cc(X), we have

μ (f) μ (n .f) x = x v . μx (ψ) μx (nv .ψ) Proof. We claim that there exists a constant b>0 such that for each i ≥ 1, the following holds for all k i 1: μ (f) μx (ε ,)(nv .f) (8.6) yk(εi,) − k i εi,

We first deduce the proposition from this claim. Since both yk(εi,),xk(εi,)belong to the set Er/16(ψ) and converge to x,andf,ψ ∈ Cc(X) have compact supports, → → →∞ μyk(εi,)(f) μx (f)andμyk(εi,)(ψ) μx (ψ)ask . Since n .f converges to n .f pointwise as i →∞and the supports of vεi, v all functions involved are contained in one fixed compact subset of X,wehave μ (n .f) → μ (n .f)ask →∞. Similarly, μ (n .ψ) → xk(εi,) vεi, x vεi, xk(εi,) vεi, μ (n .ψ)ask →∞. Hence (8.6) implies, by taking k →∞,that x vεi, μ (f) μx (nv .f) x −  εi, ≤ bε . μ (ψ) μ (n .ψ) i x x vεi,

Now by taking i →∞, this proves the proposition as εi → 0.

To prove Claim (8.6), fixing ε := εi,wesetv = vεi , sk = sk(εi), xk = xk(εi,) and yk = yk(εi,) for simplicity. By Lemma 7.8, we have, as k →∞, sk − f(ykut)dt μ (f) (1 ε)sk ∼ yk (8.7) sk . ψ(ykut)dt μyk (ψ) (1−ε)sk

Since n+ ∈O, similar calculation implies that, as k →∞, v sk · − nv f(xkut)dt μ (n .f) (1 ε)sk ∼ xk v (8.8) sk . nv.ψ(xkut)dt μxk (nv.ψ) (1−ε)sk

Therefore the claim follows if we show for all large k i 1, sk sk − f(ykut)dt − nv.f(xkut)dt (8.9) (1 ε)sk − (1 ε)sk ≤ bε sk sk ψ(ykut)dt nv.ψ(xkut)dt (1−ε)sk (1−ε)sk

for some b>0 independent of ε.Letcf and cψ denote the Lipschitz constants of f and ψ respectively. Hence for all t ∈ [(1 − ε)sk,sk] and large k  1,

| − | −1 ≤ f(xkutpk(t)) f(xkutv)

and | − | −1 ≤ ψ(xkutpk(t)) ψ(xkutv)

License or copyright restrictions may apply to redistribution; see https://www.ams.org/journal-terms-of-use 572 AMIR MOHAMMADI AND HEE OH

We have sk sk − f(ykut)dt − f(xkutpk(t))dt (1 ε)sk (1 ε)sk sk = sk − ψ(ykut)dt − ψ(xkutpk(t))dt (1 ε)sk (1 ε)sk sk · − nv f(xkut)dt (1 ε)sk = sk − ψ(xkutpk(t))dt (1 ε)sk sk − · − f(xkutpk(t)) nv f(xkut)dt (1 ε)sk + sk . ψ(xkutpk(t))dt (1−ε)sk ∪ Let K0 = Kψ Kf . Then the above estimate, (8.5) and (8.8) with f = χK0 , imply that sk − sk − f(xkutpk(t)) nv.f(xkut)dt − χK0 (xkut)dt (1 ε)sk ≤ 2c cε (1 ε)sk sk f sk ψ(xkutpk(t))dt ψ(xkut)dt (1−ε)sk (1−ε)sk

μxk (K0) ≤ 4cf cε . μxk (ψ) On the other hand, we have sk − nv.f(xkut)dt (1 ε)sk sk ψ(xkutpk(t))dt (1−ε)sk − sk sk − 1 − nv.f(xkut)dt − ψ(xkutpk(t)) nv.ψ(xkut)dt (1 ε)sk · (1 ε)sk = sk 1+ sk . nv.ψ(xkut)dt nv · ψ(xkut)dt (1−ε)sk (1−ε)sk Similar estimate as above gives sk − − ψ(xkutpk(t)) nv.ψ(xkut)dt μ (K ) (1 ε)sk ≤ 4c cε xk 0 . sk ψ nv.ψ(xkut)dt μxk (ψ) (1−ε)sk All these together imply there exists a constant c > 0 (depending on f and ψ but independent of ε) such that sk sk − f(ykut)dt − nv.f(xkut)dt (1 ε)sk  (1 ε)sk  sk =(1+c ε) sk + c ε. ψ(ykut)dt nv.ψ(xkut)dt (1−ε)sk (1−ε)sk Now by (8.7), (8.8), this implies the claim (8.9). 

The following proposition finishes the proof of Theorem 8.1.

Proposition 8.8. μx0 is invariant under N. { ∈ } Proof. The set n N : n.μx0 = μx0 is a closed subgroup which contains U.Let | | 1 ≤| |≤ c1 | |≥ (v) x and v be as in Proposition 8.7. Since c  v  and (v) c ,it 1 ∈ 1 suffices to show that μx0 is invariant under nv for all >0.Notethatx Eρ(ψ). Set N0 := {n ∈ N : xn ∈ Eρ(ψ)}. We have for any n ∈ N0 and f ∈ Cc(X), T μ (n.f) f(x u n)dt μ (f) x 0  t xn = lim T = . μx (n.ψ) T μx n(ψ)  0 ψ(xutn)dt 

License or copyright restrictions may apply to redistribution; see https://www.ams.org/journal-terms-of-use BURGER-ROBLIN 573

On the other hand, by Proposition 8.7, we have

μ (n.f) μx n (n.f) μ (n .f) x =  v = xn v . μ (n.ψ) μ (n.ψ) μ (n .ψ) x x.nv x.n v

Therefore for any n ∈ N0, μ (f) μ (ψ) xn = xn (=0) . μxn(nv .f) μx.n(nv .ψ)   As x ∈ E , it follows from the definition of E that we can take a sequence nm −1 such that xnm ∈ Eρ(ψ) ∩ B(x0,m ) and hence xnm → x0 as m →∞. In particular, μ (f) μ (f) μ (ψ) μ (ψ) x0 = lim xnm = lim xnm = x0 . m→∞ m→∞ μx0 (nv .f) μxnm (nv .f) μxnm (nv .ψ) μx0 (nv .ψ)

It follows that μx0 and nv .μx0 are not mutually singular to each other. Hence  by Lemma 8.2, μx0 = nv .μx0 . Finally, we state the following: recall the notation mBR from the subsection 2.5. N0

Theorem 8.9. If U0 is a one-parameter unipotent subgroup of G and Γ is a convex cocompact subgroup with δ>1,thenmBR is U -ergodic for N = C (U ). N0 0 0 G 0 ∈ −1 ⊂ Proof. Let k0 K be such that U0 = k0 Uk0.IfB X is a Borel subset which BR is U0 invariant, then Bk0 is U-invariant. Hence by Corollary 8.3, m (Bk0)=0 or mBR(X − Bk ) = 0. By the definition of mBR, it follows that mBR(B)=0or 0 N0 N0 mBR(X − B)=0.  N0 9. BR is not ergodic if 0 <δ≤ 1 Let Γ be a non-elementary torsion-free discrete subgroup of G. In this final section, we show that mBR is never U-ergodic if Γ is convex cocompact and δ ≤ 1. By the Hopf decomposition theorem (cf. [11]), any ergodic measure preserving flow on a σ-finite measure space is either completely dissipative or completely con- servative. In the former case, the action is isomorphic to the translation action of R on R with respect to the Lebesgue measure [1], Since Γ is non-elementary, it follows that if mBR were U-ergodic, then it must be completely conservative. We first consider Fuchian groups: Γ is called Fuchsian if it is contained in a conjugate of PSL2(R). For a Fuchsian group Γ, Γ is geometrically finite if and only if it is finitely generated. Theorem 9.1. If Γ is a finitely generated Fuchsian group, then mBR is not U-ergodic. BR Proof. The support ofm ˜ consists of xnzmθ on the unstable horospheres H(x) − based on Λ(Γ), i.e., x ∈ G with x ∈ Λ(Γ), z ∈ C and mθ ∈ M. As Γ is Fuchsian, the convex hull of Λ(Γ) is contained in a geodesic plane, say, H, preserved by Γ. Let d denote a right K-invariant and left G-invariant metric on G.Then xnzmθut = xnz+teiθ mθ and hence d(xnzmθut,H)=d(xnz+teiθ ,H) →∞as t → ∞, except for two directions of θ parallel to H.SinceH is Γ-invariant, we have d(γxnz+teiθ ,H)=d(xnz+teiθ ,H) for any γ ∈ Γ. Therefore for any z ∈ C and θ not BR parallel to H,Γ\Γxnzmθut →∞.Thisimpliesthatalmostallm -points w ∈ X, BR wut goes to ∞ as t →∞. Therefore m cannot be completely conservative for the U action and hence is not U-ergodic. 

License or copyright restrictions may apply to redistribution; see https://www.ams.org/journal-terms-of-use 574 AMIR MOHAMMADI AND HEE OH

To prove the non-ergodicity in the remaining cases we begin by recalling some standard facts. For a 1-set Λ in the plane (see 6.1 for the definition), Λ is called purely unrectifiable if H1(Λ ∩C) = 0 for every rectifiable curve C. We will use the following: Theorem 9.2. (cf. [8, Theorem 6.4] or [21])IfΛ ⊂ R2 has Hausdorff dimension δ ≤ 1, then the orthogonal projection of Λ on a.e. direction has also Hausdorff dimension δ. Moreover if Λ ⊂ R2 is a 1-set which is purely unrectifiable, then the orthogonal projection of Λ on a.e. direction has zero Lebesgue length. Our proof in the non-Fuchsian case follows much the same philosophy that the BR measure and the BMS measure are very closely related. To be more precise, Theorem 9.2 and the definition of the BMS-measure imply that the BMS measure is not recurrent when 0 <δ<1. We will use this fact to show the somewhat weaker non-recurrence holds for the BR measure. The following makes this more precise.

Definition 9.3. A measure preserving flow ut on a σ-finite measure space (X, μ) is called strongly recurrent if for any two measurable subsets A1,A2 with μ(Ai) > 0 we have {t : xut ∈ A2} is unbounded for μ-a.e. x ∈ A1. Note that by the Hopf ratio ergodic theorem, any conservative, ergodic measure preserving flow is strongly recurrent. In particular, since Γ is non-elementary, it follows that if mBR were U-ergodic, then it must be strongly recurrent. Theorem 9.4. Let Γ be convex cocompact. Suppose either that 0 <δ<1 or that Λ(Γ) is a purely unrectifiable 1-set. Then the action of U = {ut} on X =Γ\G is not strongly recurrent for mBR.InparticularmBR is not U-ergodic. Proof. We will prove this by contradiction. So, let us assume that the action of U is strongly recurrent. We fix a BMS box E = x0Bρ in X with small 0 <ρ 1. We claim that there exists a Borel subset E of E ∩ supp(mBR) with mBR(E − E)=0  such that for any x ∈ E , {t ∈ R : xnut ∈ E} is unbounded for almost all n ∈ N (with respect to the Lebesgue measure of N). To show this, for any  ∈ N,letε()=ε(, E) > 0bechosensothatifwelet ˇ B = Nε()Aε()NM, then the map b → xgb is injective on B for all x ∈ E and all g in the ρ-neighborhood of e in G. This is possible as Γ does not contain any parabolic ˇ element. Recalling the notation Tε() = Nε()Aε()M,let{xj Tε()Nε() : j ∈ J} be a finite cover of E. Then, the assumption on the strong recurrence and the Fubini theorem imply that for any j ∈ J and almost every x ∈ xj Tε()Nε(),wehavethat for a. e. n ∈ N,theset{t : xnut ∈ E} is unbounded. Hence, the claim follows.    Fix x ∈ E and consider a sequence x b for b = alog(ρ/),  ∈ N.Since  − ∈ → ∞ {  } (x ) Λ(Γ) there exists a subsequence i + such that x bi converges   + PS ∈ ∈ | to some y Ω. As (y ) Λ(Γ), μy y Nρ is a positive δ-dimensional Hausdorff measure. By the assumption, Theorem 6.3 and Theorem 9.2 imply that for almost ∈  ∩ PS all mθ M we have pθ(y Nρ supp(μy )) has zero Lebesgue length (see Section 5.1  for the notation pθ). Since E is M-invariant and E is of full measure in E,wecan ∈  ∈   ∩ PS choose mθ M such that x mθ E and pθ(y Nρ supp(μy )) has zero Lebesgue length. We set     x = x mθ,xi := x bi mθ = x mθbi , and y = y mθ. → → ∞ { ∈ } ∈ Then xi y as i + , t : xnut E is unbounded for a.e. n N and the ∩ PS projection p0(yNρ supp(μy )) has zero Lebesgue length.

License or copyright restrictions may apply to redistribution; see https://www.ams.org/journal-terms-of-use BURGER-ROBLIN 575

By the Fubini theorem, there exists a Borel subset V  ⊂ V of co-measure zero  that for each v ∈ V ,theset{t : xvut ∈ E} is unbounded. Hence, for each j ∈ Z,  ∈  ∩{ ∈ − } we can find a sequence vj V nit : t [j 1,j+1] . With abuse of notation,  R  ∈ − we consider vj as an element of and write vj [j 1,j+1]. ± ∈ ∈ − ∈ Since E = x0Bρ with x0 Λ(Γ), the condition gnz E with g Λ(Γ) implies ∈ ∈ C | |≤  ∈  that gnz+w Ωforsomew with w ρ. Therefore, for each vj V ,  ∈ ∈ R ∈ C | |≤ xvj ut E for some t implies the existence of w with w ρ such  ∈ ∈ Z that xvj utnw Ω. This in particular implies that for each j ,thereexists ∈ ∩ PS | −  |≤ ≤ ∈ − xvj xV p0(supp(μx )) with vj vj ρ 1. Note that vj [j 2,j+2]for each j ∈ Z. −1  −1 ∈ Flowing xvj by b,wegetxvj b = x(b vj b). Setting vj := (b vj b) V ,we have − (j 2)ρ  (j +2)ρ  PS ≤ v ≤ and v ∈ p0(supp(μ )).  j  j x | ± | Therefore x V ∩ p (supp(μPS)) contains x v whenever j 2 < 1.  ρ 0 x  j  ∩ PS { | | + ∈ } Note that xi Nρ supp(μ )= xi nz : z <ρ,(xi nz) Λ(Γ) and that xi → → + any limit of xi n is of the form yn as xi y. Since the visual map, g g , ∩ is continuous and Λ(Γ) is closed, it follows that the sequence of subsets xi Nρ PS ∩ PS ∈ ∩ PS supp(μ )convergestoyNρ supp(μ ) in the sense that if wi xi Nρ supp(μ ) xi y xi ∈ ∩ PS →∞ ∩ converges to w,thenw yNρ supp(μy ). Therefore, as i , p0(xi Nρ PS PS supp(μ )) converges to p0(yNρ ∩ supp(μ )) as well in the similar sense as above. xi y ⊂ PS Thus we have obtained that yVρ p0(supp(μy )) and hence the Lebesgue length ∩ PS  of p0(Nρ supp(μy )) is positive, yielding a contradiction. A Fuchsian group Γ is called of the first kind if Λ(Γ) is a great circle and of the second kind otherwise. By Canary and Taylor [6], any torsion-free and convex cocompact group Γ with δ = 1 is either a Fuchsian subgroup of the first kind or a quasi-conformal conjugation of a Fuchsian group of the second kind, and in the latter case, the limit set Λ(Γ) is totally disconnected. We thank Chris Bishop for providing the proof of the following theorem: Theorem 9.5 (Bishop). If the limit set of a convex cocompact Kleinian group is totally disconnected, it is always purely unrectifiable. Proof. If the limit set of a convex co-compact group hits a rectifiable curve in positive length, then the limit set contains a circle (or line). This can be proved by taking a sequence of balls shrinking to a point of density of the set of positive length that is also a point of tangency for the rectifiable curve and rescaling the balls by group elements to approximately unit size (this can always be done in a convex co-compact group). The rescaled sets must have a subsequence that converges to a line segment or circular arc in the limit set. Rescaling of the arc leads to a full circle or line in the limit set. In particular, a convex co-compact limit set that is totally disconnected cannot hit any rectifiable curve in positive length. Thus, it is purely unrectifiable. 

Therefore by Theorem 9.1, Theorem 9.4, and Theorem 9.5, we have: Theorem 9.6. Let Γ be torsion-free and convex cocompact with 0 <δ≤ 1.Then mBR is not U-ergodic.

License or copyright restrictions may apply to redistribution; see https://www.ams.org/journal-terms-of-use 576 AMIR MOHAMMADI AND HEE OH

Acknowledgments We are very grateful to Tim Austin for numerous helpful discussions regarding various aspects of this project. We also thank Chris Bishop and Edward Taylor for helpful correspondences regarding totally disconnected limit sets of Kleinian groups.

References

[1] Jon Aaronson, An introduction to infinite ergodic theory, Mathematical Surveys and Monographs, vol. 50, American Mathematical Society, Providence, RI, 1997. MR1450400 (99d:28025) [2] Martine Babillot, On the mixing property for hyperbolic systems, Israel J. Math. 129 (2002), 61–76, DOI 10.1007/BF02773153. MR1910932 (2003g:37008) [3] M. Brin, Ergodic theory of frame flows, Ergodic theory and dynamical systems, II (College Park, Md., 1979/1980), Progr. Math., vol. 21, Birkh¨auser, Boston, Mass., 1982, pp. 163–183. MR670078 (83m:58059) [4] B. H. Bowditch, Geometrical finiteness for hyperbolic groups, J. Funct. Anal. 113 (1993), no. 2, 245–317, DOI 10.1006/jfan.1993.1052. MR1218098 (94e:57016) [5] Marc Burger, Horocycle flow on geometrically finite surfaces, Duke Math. J. 61 (1990), no. 3, 779–803, DOI 10.1215/S0012-7094-90-06129-0. MR1084459 (91k:58102) [6] Richard D. Canary and Edward Taylor, Kleinian groups with small limit sets, Duke Math. J. 73 (1994), no. 2, 371–381, DOI 10.1215/S0012-7094-94-07316-X. MR1262211 (94m:57028) [7] M. Einsiedler and E. Lindenstrauss, Diagonal actions on locally homogeneous spaces, Homo- geneous flows, moduli spaces and arithmetic, Clay Math. Proc., vol. 10, Amer. Math. Soc., Providence, RI, 2010, pp. 155–241. MR2648695 (2011f:22026) [8] Kenneth Falconer, , 2nd ed., John Wiley & Sons, Inc., Hoboken, NJ, 2003. Mathematical foundations and applications. MR2118797 (2006b:28001) [9] L. Flaminio and R. J. Spatzier, Geometrically finite groups, Patterson-Sullivan measures and Ratner’s rigidity theorem, Invent. Math. 99 (1990), no. 3, 601–626, DOI 10.1007/BF01234433. MR1032882 (91d:58201) [10] Michael Hochman, A ratio ergodic theorem for multiparameter non-singular actions,J. Eur. Math. Soc. (JEMS) 12 (2010), no. 2, 365–383, DOI 10.4171/JEMS/201. MR2608944 (2011g:37006) [11] E Hopf, Ergodentheorie, Ergebnisse der Mathematik Number 5 (1937). [12] Xiaoyu Hu and S. James Taylor, Fractal properties of products and projections of mea- sures in Rd, Math. Proc. Cambridge Philos. Soc. 115 (1994), no. 3, 527–544, DOI 10.1017/S0305004100072285. MR1269937 (95f:28013) [13] A. Katok and R. J. Spatzier, Invariant measures for higher-rank hyperbolic abelian actions, Ergodic Theory Dynam. Systems 16 (1996), no. 4, 751–778, DOI 10.1017/S0143385700009081. MR1406432 (97d:58116) [14] Ulrich Krengel, Ergodic theorems, de Gruyter Studies in Mathematics, vol. 6, Walter de Gruyter & Co., Berlin, 1985. With a supplement by Antoine Brunel. MR797411 (87i:28001) [15] Elon Lindenstrauss, Invariant measures and arithmetic quantum unique ergodicity, Ann. of Math. (2) 163 (2006), no. 1, 165–219, DOI 10.4007/annals.2006.163.165. MR2195133 (2007b:11072) [16] J.-L. Lions and E. Magenes, Non-homogeneous boundary value problems and applications. Vol. I, Springer-Verlag, New York-Heidelberg, 1972. Translated from the French by P. Kenneth; Die Grundlehren der mathematischen Wissenschaften, Band 181. MR0350177 (50 #2670) [17] G. A. Margulis, On the action of unipotent groups in the space of lattices, Lie groups and their representations (Proc. Summer School, Bolyai, J´anos Math. Soc., Budapest, 1971), Halsted, New York, 1975, pp. 365–370. MR0470140 (57 #9907) [18] G. A. Margulis, Indefinite quadratic forms and unipotent flows on homogeneous spaces,Dy- namical systems and ergodic theory (Warsaw, 1986), Banach Center Publ., vol. 23, PWN, Warsaw, 1989, pp. 399–409. MR1102736 (92g:11034) [19] J. M. Marstrand, Some fundamental geometrical properties of plane sets of fractional dimen- sions, Proc. London Math. Soc. (3) 4 (1954), 257–302. MR0063439 (16,121g)

License or copyright restrictions may apply to redistribution; see https://www.ams.org/journal-terms-of-use BURGER-ROBLIN 577

[20] Pertti Mattila, Geometry of sets and measures in Euclidean spaces, Cambridge Studies in Advanced Mathematics, vol. 44, Cambridge University Press, Cambridge, 1995. and rectifiability. MR1333890 (96h:28006) [21] Pertti Mattila, Hausdorff dimension, projections, and the Fourier transform, Publ. Mat. 48 (2004), no. 1, 3–48, DOI 10.5565/PUBLMAT 48104 01. MR2044636 (2004k:28018) [22] Calvin C. Moore, Ergodicity of flows on homogeneous spaces,Amer.J.Math.88 (1966), 154–178. MR0193188 (33 #1409) [23] Calvin C. Moore, Ergodicity of flows on homogeneous spaces,Amer.J.Math.88 (1966), 154–178. MR0193188 (33 #1409) [24] Hee Oh and Nimish A. Shah, Equidistribution and counting for orbits of geometrically finite hyperbolic groups,J.Amer.Math.Soc.26 (2013), no. 2, 511–562, DOI 10.1090/S0894-0347- 2012-00749-8. MR3011420 [25] Jean-Pierre Otal and Marc Peign´e, Principe variationnel et groupes kleiniens (French, with English and French summaries), Duke Math. J. 125 (2004), no. 1, 15–44, DOI 10.1215/S0012- 7094-04-12512-6. MR2097356 (2005h:37053) [26] S. J. Patterson, The limit set of a Fuchsian group,ActaMath.136 (1976), no. 3-4, 241–273. MR0450547 (56 #8841) [27] Marc Peign´e, On the Patterson-Sullivan measure of some discrete group of isometries,Israel J. Math. 133 (2003), 77–88, DOI 10.1007/BF02773062. MR1968423 (2004b:30079) [28] Yuval Peres and Wilhelm Schlag, Smoothness of projections, Bernoulli convolutions, and the dimension of exceptions, Duke Math. J. 102 (2000), no. 2, 193–251, DOI 10.1215/S0012- 7094-00-10222-0. MR1749437 (2001d:42013) [29] Marina Ratner, On measure rigidity of unipotent subgroups of semisimple groups,ActaMath. 165 (1990), no. 3-4, 229–309, DOI 10.1007/BF02391906. MR1075042 (91m:57031) [30] Marina Ratner, On Raghunathan’s measure conjecture, Ann. of Math. (2) 134 (1991), no. 3, 545–607, DOI 10.2307/2944357. MR1135878 (93a:22009) [31] Thomas Roblin, Ergodicit´eet´equidistribution en courbure n´egative (French, with English and French summaries), M´em.Soc.Math.Fr.(N.S.)95 (2003), vi+96. MR2057305 (2005d:37060) [32] Thomas Roblin, Sur l’ergodicit´e rationnelle et les propri´et´es ergodiques du flot g´eod´esique dans les vari´et´es hyperboliques (French, with French summary), Ergodic Theory Dynam. Systems 20 (2000), no. 6, 1785–1819, DOI 10.1017/S0143385700000997. MR1804958 (2001m:37061) [33] Daniel J. Rudolph, Ergodic behaviour of Sullivan’s geometric measure on a geometrically finite , Ergodic Theory Dynam. Systems 2 (1982), no. 3-4, 491–512 (1983), DOI 10.1017/S0143385700001735. MR721736 (85i:58101) [34] Dennis Sullivan, The density at infinity of a discrete group of hyperbolic motions, Inst. Hautes Etudes´ Sci. Publ. Math. 50 (1979), 171–202. MR556586 (81b:58031) [35] Dennis Sullivan, Entropy, Hausdorff measures old and new, and limit sets of geometrically finite Kleinian groups,ActaMath.153 (1984), no. 3-4, 259–277, DOI 10.1007/BF02392379. MR766265 (86c:58093) [36] D Winter, Mixing of frame flow for rank one locally symmetric spaces and measure classifi- cation, available at arXiv:1403.2425. [37] Shing Tung Yau, Harmonic functions on complete Riemannian manifolds, Comm. Pure Appl. Math. 28 (1975), 201–228. MR0431040 (55 #4042) [38] R. Zweim¨uller, Hopf ’s ratio ergodic theorem by inducing.Preprint.

Department of Mathematics, The University of Texas at Austin, Austin, Texas 78750 E-mail address: [email protected] Department of Mathematics, Yale University, New Haven, Connecticut 06520 and Korea Institute for Advanced Study, Seoul, Korea E-mail address: [email protected]

License or copyright restrictions may apply to redistribution; see https://www.ams.org/journal-terms-of-use