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

Uniform spectral gap and orthogeodesic counting for strong convergence of Kleinian groups

Beibei Liu Department of Mathematics, The Ohio State University, Columbus, OH 43210, USA [email protected]  and  Franco Vargas Pallete Department of Mathematics, Yale University, New Haven, CT 06511, USA [email protected]
Abstract.

We show convergence of small eigenvalues for geometrically finite hyperbolic nn-manifolds under strong limits. For a class of convergent convex sets in a strongly convergent sequence of Kleinian groups, we use the spectral gap of the limit manifold and the exponentially mixing property of the geodesic flow along the strongly convergent sequence to find asymptotically uniform counting formulas for the number of orthogeodesics between the convex sets. In particular, this provides asymptotically uniform counting formulas (with respect to length) for orthogeodesics between converging Margulis tubes, geodesic loops based at converging basepoints, and primitive closed geodesics.

1. Introduction

The critical exponent of a discrete isometry subgroup of the hyperbolic space n\mathbb{H}^{n} is an important numerical invariant which relates the dynamical properties of the group action to the measure theory and the spectrum of operators on the quotient manifold via the celebrated work of Patterson and Sullivan [Pat76, Sul79, Sul84]. More explicitly, this invariant was shown to be equal to the Hausdorff dimension of the limit set for any geometrically finite discrete isometry subgroup Γ\Gamma [Sul79, Sul84], and is related to the bottom spectrum λ0\lambda_{0} of the negative Laplace operator for any nonelementary complete hyperbolic manifold [Sul87]. A natural line of inquiry is to ask whether this quantitative invariant can be uniformly controlled for a sequence of hyperbolic manifolds (Mk=n/Γk)k(M_{k}=\mathbb{H}^{n}/\Gamma_{k})_{k\in\mathbb{N}}, for example, sequences of quasi-Fuchsian manifolds in Bers’ model for the Teichmüller space of a surface SS. It turns out that the critical exponent of Γk\Gamma_{k}, the Hausdorff dimension of the limit set, and the bottom of the spectrum λ0(n/Γk)\lambda_{0}(\mathbb{H}^{n}/\Gamma_{k}), converge to the ones of the limit group Γ<Isom(n)\Gamma<\operatorname{{\mathrm{I}som}}(\mathbb{H}^{n}) under the assumption that Γ\Gamma is geometrically finite and δ(Γ)>(n1)/2\delta(\Gamma)>(n-1)/2 for strongly convergent sequences of hyperbolic manifolds (Mk)k(M_{k})_{k\in\mathbb{N}} [CT99, McM99]. See Section 2.5 for the definition of strong convergence.

Besides the bottom spectrum of the quotient manifold, there are finitely many small eigenvalues of the negative Laplace operator in the interval [λ0,(n1)2/4][\lambda_{0},(n-1)^{2}/4], where (n1)2/4(n-1)^{2}/4 is the bottom spectrum of the hyperbolic space n\mathbb{H}^{n} [LP82]. It is natural to ask whether these small eigenvalues converge to the ones of Γ\Gamma, respectively. We prove the convergence of small eigenvalues for strongly convergent sequences of hyperbolic manifolds (Mk=n/Γk)k(M_{k}=\mathbb{H}^{n}/\Gamma_{k})_{k\in\mathbb{N}}. In particular, we give a uniform bound on the Lax-Phillips spectral gap s1s_{1} defined by s1:=min{λ1(M),(n1)2/4}λ0(M)s_{1}:=\min\{\lambda_{1}(M),(n-1)^{2}/4\}-\lambda_{0}(M), where λ1(M)\lambda_{1}(M) is the smallest eigenvalue of the negative Laplacian in (λ0(M),)(\lambda_{0}(M),\infty).

Theorem 1.1.

Suppose that (Mk=Isom(n)/Γk)k(M_{k}=\operatorname{{\mathrm{I}som}}(\mathbb{H}^{n})/\Gamma_{k})_{k\in\mathbb{N}} is a sequence of hyperbolic manifolds which converges strongly to a geometrically finite hyperbolic manifold M=n/ΓM=\mathbb{H}^{n}/\Gamma. The set of small eigenvalues in [λ0(Mk),(n1)2/4][\lambda_{0}(M_{k}),(n-1)^{2}/4] converges to the small eigenvalues of the limit manifold MM, counting multiplicities. In particular, the sequence of Lax-Phillips spectral gaps of (Mk)k(M_{k})_{k\in\mathbb{N}} converges to that of the limit manifold MM.

Remark 1.2.

We explain what the convergence of the set of small eigenvalues means in Section 3, and leave the precise statement in Theorem 3.3. The statement of Theorem 1.1 for small eigenvalues holds for negatively pinched manifolds, and the details are discussed in Section 3. The statement referring to Lax-Phillips spectral gap is done in Theorem 3.4 for Kleinian groups. It could be possible that the set of small eigenvalues is equal to the singleton {(n1)2/4}\{(n-1)^{2}/4\} (or the empty set by considering pinched negative manifolds), but it won’t affect the statement of the theorem.

Sequences of hyperbolic manifolds with uniform spectral gap are interesting to study, as the uniform spectral gap sometimes controls the dynamical properties of the geodesic flow of the manifold. For instance, following [EO21], uniform spectral gaps of hyperbolic manifolds imply uniform exponential mixings of geodesic flows. In the same paper, they provided another family of hyperbolic manifolds with uniform spectral gaps, coming from congruence subgroups of certain arithmetic lattice of Isom(n)\operatorname{{\mathrm{I}som}}(\mathbb{H}^{n}).

The exponentially mixing geodesic flow can be used to find good estimates for error in asymptotic approximations of counting functions, such as the estimates available for orthogeodesic counting (as done in [PP17]). Namely, given D,D+D^{-},D^{+} (locally) convex sets (or equivalently, π1(M)\pi_{1}(M) precisely invariant convex sets in the universal covering) in MM, one can estimate 𝒩D,D+(t)\mathcal{N}_{D^{-},D^{+}}(t), the number of orthogeodesics between DD^{-} and D+D^{+} of length less than t>0t>0, by

𝒩D,D+(t)Aeδt(1+O(eκt))\mathcal{N}_{D^{-},D^{+}}(t)\approx Ae^{\delta t}(1+O(e^{-\kappa t}))

where A,δ,κA,\delta,\kappa and O(.)O(.) depend on the geometric/dynamical features of M,D,D+M,D^{-},D^{+}, with exponential decay of correlations among these features. We consider the following two interesting cases in this paper:

  1. (1)

    D±D^{\pm} are connected components in the thin part of MM, i.e. Margulis tubes or cusps.

  2. (2)

    D+=DD^{+}=D^{-} is an embedded ball at a given point xMx\in M. That is, the lifts of D±D^{\pm} are sufficiently small balls of lifts of xx in n\mathbb{H}^{n}.

The uniform orthogeodesic counting formula for strongly convergent sequences in case (1) can be used in the study of the renormalized volume. Given a hyperbolic manifold MM, the renormalized volume is a function on the deformation space of MM whose gradient flow has been of interest (see [BBB19], [BBP21]). In [BBP21] it is shown that for MM acylindrical the gradient flow of the renormalized volume converges to the unique critical point. This involves discarding strong limits with pinched rank-11 cusps by the use of the Gardiner formula. For such a method to work one needs a uniform control of contributing terms in the Gardiner formula, which would be provided by uniform orthogeodesic counting. The uniform orthogeodesic counting formula for case (2) gives a uniform asymptotic counting result with uniform error term for geodesic loops based on a given point in MM.

Motivated by these applications, we show that the parameters A,δ,κA,\delta,\kappa and O(.)O(.) are uniform for strongly convergent sequences, and such parameters can be taken arbitrarily close to the corresponding parameters of the geometrically finite limit.

Theorem 1.3.

Let (Mk=n/Γk)k(M_{k}=\mathbb{H}^{n}/\Gamma_{k})_{k\in\mathbb{N}} be a sequence of hyperbolic manifolds which strongly converges to a geometrically finite hyperbolic manifold M=n/ΓM=\mathbb{H}^{n}/\Gamma with δ(Γ)>(n1)/2\delta(\Gamma)>(n-1)/2.

  1. (1)

    Suppose that Dk±D^{\pm}_{k} are connected components in the thin part of MkM_{k}, and (Dk±)k(D^{\pm}_{k})_{k\in\mathbb{N}} converge strongly to connected components D±D^{\pm} in the thin part of MM. Then there is a uniform counting formula for orthogeodesics between DkD^{-}_{k} to Dk+D^{+}_{k} for the sequence (Mk)k(M_{k})_{k\in\mathbb{N}}.

  2. (2)

    Suppose that (xkMk)k(x_{k}\in M_{k})_{k\in\mathbb{N}} is a sequence of points converging to the point xMx\in M. Then there is a uniform counting formula for geodesic loops based at xkx_{k} for the sequence (Mk)k(M_{k})_{k\in\mathbb{N}}.

Remark 1.4.

We in fact prove the result for strongly convergent sequences of well-positioned convex sets in a strongly convergent sequence of hyperbolic manifolds (Theorem 5.3). We refer readers to Section 2.5 for the definitions of well-positioned and strong convergence of convex sets in hyperbolic manifolds.

The counting of primitive closed geodesics follows from the counting of geodesic loops in manifolds with negatively pinched curvatures [Rob03, Chapter 5]. Hence we obtain the following asymptotic counting of primitive closed geodesics along sequences of strongly convergent hyperbolic manifolds.

Corollary 1.5.

Suppose that (Mk=n/Γk)k(M_{k}=\mathbb{H}^{n}/\Gamma_{k})_{k\in\mathbb{N}} is a sequence of hyperbolic manifolds which strongly converges to a geometrically finite hyperbolic manifold M=n/ΓM=\mathbb{H}^{n}/\Gamma with δ(Γ)>(n1)/2\delta(\Gamma)>(n-1)/2. Then we can count the number of primitive closed geodesics with length less than \ell in MkM_{k}, denoted by #𝒢Mk()\#\mathcal{G}_{M_{k}}(\ell), uniformly, in the sense that

#𝒢Mk()eδ(Γk)δ(Γk)\#\mathcal{G}_{M_{k}}(\ell)\approx\frac{e^{\delta(\Gamma_{k})\ell}}{\delta(\Gamma_{k})\ell}

up to a multiplicative error uniformly close to 1 along the sequence as \ell gets larger and limkδ(Γk)=δ(Γ)\lim_{k}\delta(\Gamma_{k})=\delta(\Gamma).

The proof of Theorem 1.3 involves the uniformity of the exponential mixing and the convergence of certain measures for strongly convergent sequences. These measures refer to the classical Patterson-Sullivan measures, the Bowen-Margulis measure and the skinning measures. The convergence of Patterson-Sullivan measures has been proved for strongly convergent sequences under the assumption that the limit manifold is geometrically finite and its critical exponent is greater than (n1)/2(n-1)/2, [McM99]. The Bowen-Margulis measure and the skinning measures are defined in terms of the Patterson-Sullivan measures. Answering an question of Oh, we prove the convergence of these two measures, which could have its own interest.

Proposition 1.6.

Suppose that (Mk=n/Γk)k(M_{k}=\mathbb{H}^{n}/\Gamma_{k})_{k\in\mathbb{N}} is a sequence of hyperbolic manifolds which are strongly convergent to a geometrically finite hyperbolic manifold M=n/ΓM=\mathbb{H}^{n}/\Gamma with δ(Γ)>(n1)/2\delta(\Gamma)>(n-1)/2. For r>0r>0 we denote by Mk<rMk,M<rMM_{k}^{<r}\subset M_{k},M^{<r}\subset M the sets of points with injectivity radius less than r. Then the Bowen-Margulis measures mBMkm_{\rm{BM}}^{k} on T1Mk<rT^{1}M_{k}^{<r} converge to the one on T1M<rT^{1}M^{<r} weakly. Moreover, we have the convergence of total masses.

Remark 1.7.

The convergence of the Bowen-Margulis measures on T1Mk<rT^{1}M_{k}^{<r} might be helpful for proving that the Benjamini-Schramm limit of (Mk)k(M_{k})_{k\in\mathbb{N}} is also MM (see for instance [ABB+17, Section 3.9] for a general definition of Benjamini-Schramm convergence).

We now discuss the convergence of skinning measures σ±\sigma^{\pm} for the special type of well-positioned convex sets in hyperbolic manifolds. Geodesic balls with sufficiently small radii and the thin part in a hyperbolic manifolds are well-positioned. We refer readers to Section 2.5 for the definition and detailed discussions.

Corollary 1.8.

Suppose that (Mk=n/Γk)k(M_{k}=\mathbb{H}^{n}/\Gamma_{k})_{k\in\mathbb{N}} is a sequence of hyperbolic manifolds that strongly converges to a geometrically finite hyperbolic manifold M=n/ΓM=\mathbb{H}^{n}/\Gamma with δ(M)>(n1)/2\delta(M)>(n-1)/2. Let DkMk,DMD_{k}\subset M_{k},D\subset M be well-positioned convex sets, so that (Dk)k(D_{k})_{k\in\mathbb{N}} strongly converges to DD. Then

σDk±σD±.\|\sigma^{\pm}_{\partial D_{k}}\|\rightarrow\|\sigma^{\pm}_{\partial D}\|.

The relative result also holds for subsets ΩkDk,ΩD\Omega_{k}\subseteq D_{k},\,\Omega\subseteq D so that (Ωk)k(\Omega_{k})_{k\in\mathbb{N}} strongly converges to Ω\Omega.

Organization of the paper. We review definitions of geometric finiteness, the Bowen-Margulis measure, and skinning measures in Section 2.1, 2.3, 2.4, respectively. Section 2.2 is about the relation between the critical exponent and the bottom spectrum. Section 2.5 defines strong convergence of hyperbolic manifolds and the convergence of well-positioned convex sets. Section 3 discusses small eigenvalues of the negative Laplacian on negatively pinched Hadamard manifolds and gives a proof of Theorem 1.1. In Section 4, we prove the convergence results of the Bowen-Margulis measure and the skinning measures, i.e. Proposition 1.6 and Corollary 1.8. The last section, Section 5, proves the uniform asymptotic counting results of geodesic loops and orthogeodesics along strongly convergent sequences, i.e., the proof of Theorem 1.3.

Acknowledgements

We would like to thank Martin Bridgeman for pointing out the uniform counting question to us, and Hee Oh for suggesting Propostion 1.6 and useful discussions. We are very grateful to Curtis T. McMullen, Frédéric Paulin for helpful comments on an earlier draft, and Ian Biringer for email correspondence. We also appreciate the anonymous referees for the helpful suggestions. The first author is partially supported by the NSF grant DMS-2203237. The second author is supported by NSF grant DMS-2001997.

2. Background

2.1. Geometric finiteness

In this subsection, we let XX denote an nn-dimensional negatively pinched Hadamard manifold whose sectional curvatures lie between κ2-\kappa^{2} and 1-1 for some κ1\kappa\geq 1. For any isometry γIsom(X)\gamma\in\operatorname{{\mathrm{I}som}}(X), we define its translation length τ(γ)\tau(\gamma) as follows:

τ(γ):=infpXdX(p,γ(p)),\tau(\gamma):=\inf_{p\in X}d_{X}(p,\gamma(p)),

where dXd_{X} is the Riemannian distance function in XX. Based on the translation length, we can classify isometries in XX into 3 types; we call γ\gamma loxodromic if τ(γ)>0\tau(\gamma)>0. In this case, the infimum is attained exactly when the points are on the axis of γ\gamma. The isometry γ\gamma is called parabolic if τ(γ)=0\tau(\gamma)=0 and the infimum is not attained. The isometry γ\gamma is elliptic if τ(γ)=0\tau(\gamma)=0 and the infimum is attained.

From now on, we consider torsion-free discrete isometry subgroups Γ<Isom(X)\Gamma<\operatorname{{\mathrm{I}som}}(X), i.e. Γ\Gamma contains no elliptic elements. If Γ<Isom(n)\Gamma<\operatorname{{\mathrm{I}som}}(\mathbb{H}^{n}) is a torsion-free discrete isometry subgroup, we call it a Kleinian group. Given 0<ϵ<ϵ(n,κ)0<\epsilon<\epsilon(n,\kappa), where ϵ(n,κ)\epsilon(n,\kappa) is the Margulis constant depending on the dimension nn and the constant κ\kappa, let 𝒯ϵ(Γ)\mathcal{T}_{\epsilon}(\Gamma) be the set consisting of all points pXp\in X such that there exists an isometry γΓ\gamma\in\Gamma with

d(p,γp)ϵ.d(p,\gamma p)\leq\epsilon.

It is an Γ\Gamma-invariant set, and the quotient 𝒯ϵ(Γ)/Γ\mathcal{T}_{\epsilon}(\Gamma)/\Gamma is the thin part of the quotient manifold M=X/ΓM=X/\Gamma, denoted by M<ϵM^{<\epsilon}.

A subgroup P<ΓP<\Gamma is called parabolic if the fixed point set of PP consists of a single point ξX\xi\in\partial_{\infty}X, where X\partial_{\infty}X is the visual boundary of XX. Note that 𝒯ϵ(P)X\mathcal{T}_{\epsilon}(P)\subset X is precisely invariant under PP, i.e. stabΓ(𝒯ϵ(P))=P\operatorname{{\mathrm{s}tab}}_{\Gamma}(\mathcal{T}_{\epsilon}(P))=P [Bow95, Corollary 3.5.6]. By abuse of notation, we can regard 𝒯ϵ(P)\mathcal{T}_{\epsilon}(P) as a subset of M=X/ΓM=X/\Gamma, which is called a Margulis cusp. The union of all Margulis cusps consists of the cuspidal part of MM, denoted by cuspϵ(M)\operatorname{{\mathrm{c}usp}}_{\epsilon}(M).

The limit set Λ(Γ)\Lambda(\Gamma) of a discrete, torsion-free isometry subgroup Γ<Isom(X)\Gamma<\operatorname{{\mathrm{I}som}}(X) is defined to be the set of accumulation points of a Γ\Gamma-orbit Γ(p)\Gamma(p) in X\partial_{\infty}X for any point pXp\in X. We call Γ\Gamma elementary if Λ(Γ)\Lambda(\Gamma) is finite; Otherwise, we say Γ\Gamma is nonelementary. For any two points ξ\xi and η\eta in X\partial_{\infty}X, we use ξη\xi\eta to denote the unique geodesic in XX connecting these two points. The convex hull of Λ(Γ)X\Lambda(\Gamma)\subset\partial_{\infty}X is the smallest closed convex subset in XX whose accumulation set is Λ(Γ)\Lambda(\Gamma), denoted by Hull(Γ)\operatorname{{\mathrm{H}ull}}(\Gamma). We let C(M)=Hull(Γ)/ΓC(M)=\operatorname{{\mathrm{H}ull}}(\Gamma)/\Gamma denote the convex core of quotient manifold M=X/ΓM=X/\Gamma. For any constant ϵ>0\epsilon>0, we define the truncated core by

C(M)>ϵ=C(M)M<ϵ.C(M)^{>\epsilon}=C(M)-M^{<\epsilon}.

Given a constant 0<ϵ<ϵ(n,κ)0<\epsilon<\epsilon(n,\kappa), a discrete isometry subgroup Γ\Gamma is geometrically finite if the truncated core C(M)>ϵC(M)^{>\epsilon} is compact in M=X/ΓM=X/\Gamma. If, in addition, C(M)C(M) is compact, i.e. Γ\Gamma contains no parabolic isometries, then Γ\Gamma is called convex co-compact. Furthermore, if Γ<Isom(X)\Gamma<\operatorname{{\mathrm{I}som}}(X) is geometrically finite, the parabolic fixed points in Λ(Γ)\Lambda(\Gamma) are bounded, defined as follows:

Definition 2.1.

[Bow93] A parabolic fixed point ξΛ(Γ)\xi\in\Lambda(\Gamma) is bounded if (Λ(Γ){p})/stabΓ(p)(\Lambda(\Gamma)\setminus\{p\})/\operatorname{{\mathrm{s}tab}}_{\Gamma}(p) is compact.

Given a point xXx\in X and a discrete isometry group ΓIsom(X)\Gamma\in\operatorname{{\mathrm{I}som}}(X), the Poincaré series is defined as

Ps(Γ,x)=γΓesdX(x,γx).P_{s}(\Gamma,x)=\sum_{\gamma\in\Gamma}e^{-sd_{X}(x,\gamma x)}.

The critical exponent of Γ\Gamma is defined as

δ(Γ):=inf{sPs(Γ,x)<}.\delta(\Gamma):=\inf\{s\mid P_{s}(\Gamma,x)<\infty\}.

It is not hard to see that the definition of δ(Γ)\delta(\Gamma) is independent of the choice of xx.

2.2. Eigenvalues and spectrum

As in Section 2.1, we let M=X/ΓM=X/\Gamma, where XX is a negatively pinched Hadamard manifold, and Γ\Gamma is a torsion-free discrete isometry subgroup. Define the Sobolev space H1(M)H^{1}(M) as the space obtained by the completion of C0(M)C^{\infty}_{0}(M) with respect to the norm f=M|f|2+M|f|2\|f\|=\sqrt{\int_{M}|f|^{2}+\int_{M}|\nabla f|^{2}}. This space can be also defined as functions in L2(M)L^{2}(M) whose weak derivative (in the sense of distributions) is also in L2(M)L^{2}(M).

Given fH1(M)f\in H^{1}(M), we define the Rayleigh quotient R(f)R(f) of ff by

R(f)=M|f|2M|f|2.R(f)=\frac{\int_{M}|\nabla f|^{2}}{\int_{M}|f|^{2}}.

The Rayleigh quotient is closely related to the spectrum Spec(M)Spec(M) of the negative Laplace operator. Namely, by posing the following minimization problem

λ=inf{R(f)|fH1(M)}\lambda=\inf\bigg{\{}R(f)\,\bigg{|}\,f\in H^{1}(M)\bigg{\}}

we obtain a L2L^{2} integrable smooth function ff satisfying Δf=λf-\Delta f=\lambda f.

We let λ0(M)\lambda_{0}(M) denote the bottom of the spectrum, and we say that λSpec(M)\lambda\in Spec(M) is a small eigenvalue of MM if λ<(n1)2/4\lambda<(n-1)^{2}/4. Moreover, given a constant μ<(n1)2/4\mu<(n-1)^{2}/4, we define Specμ(M)Spec_{\mu}(M) as the collection (counting multiplicities) of eigenvalues of the negative Laplacian on MM less than or equal to μ\mu. The set of small eigenvalues is a finite set (see [Ham04]).

In the rest of the subsection, we list several properties of the bottom of the spectrum λ0(M)\lambda_{0}(M). We will use these properties in Section 3 to prove the uniform spectral gap for strongly convergent sequences of geometrically finite groups (Γk<Isom(X))k(\Gamma_{k}<\operatorname{{\mathrm{I}som}}(X))_{k\in\mathbb{N}}.

Lemma 2.2.

[Ham04] Let Γ<Isom(X)\Gamma<\operatorname{{\mathrm{I}som}}(X) be a torsion-free discrete elementary isometry subgroup of a negatively pinched Hadamard manifold XX with dimension nn. Then λ0(X/Γ)(n1)2/4\lambda_{0}(X/\Gamma)\geq(n-1)^{2}/4.

Lemma 2.3.

[Ham04, Lemma 2.3] Suppose that Γ<Isom(X)\Gamma<\operatorname{{\mathrm{I}som}}(X) is a geometrically finite discrete isometry subgroup of a negatively pinched Hadamard manifold XX with dimension nn. Then for every r>0r>0 we have that μ1(MBr(C(M)))(tanhr)2(n1)2/4\mu_{1}(M\setminus B_{r}(C(M)))\geq(\tanh r)^{2}(n-1)^{2}/4, where M=X/ΓM=X/\Gamma and μ1(MBr(C(M)))\mu_{1}(M\setminus B_{r}(C(M))) denotes the smallest Rayleigh quotient for all smooth functions ff with compact support in MBr(C(M))M\setminus B_{r}(C(M)).

If X=nX=\mathbb{H}^{n}, we have the following result relating λ0(M)\lambda_{0}(M) to the critical exponent δ(Γ)\delta(\Gamma).

Theorem 2.4.

[Sul87] For any nonelementary complete hyperbolic manifold M=n/ΓM=\mathbb{H}^{n}/\Gamma, one has

λ0(M)={(n1)2/4if δ(Γ)(n1)/2,δ(Γ)(n1δ(Γ))if δ(Γ)(n1)/2.\lambda_{0}(M)=\begin{cases}(n-1)^{2}/4&\text{if $\delta(\Gamma)\leq(n-1)/2$,}\\ \delta(\Gamma)(n-1-\delta(\Gamma))&\text{if $\delta(\Gamma)\geq(n-1)/2$.}\end{cases}

2.3. Patterson-Sullivan measure

Given a point pnp\in\mathbb{H}^{n}, and ξn\xi\in\partial_{\infty}\mathbb{H}^{n}, the Busemann function B(x,ξ)B(x,\xi) on n\mathbb{H}^{n} with respect to pp is defined by

B(x,ξ)=limt(d(x,ρξ(t))t)B(x,\xi)=\lim_{t\rightarrow\infty}(d(x,\rho_{\xi}(t))-t)

where ρξ(t)\rho_{\xi}(t) is the unique geodesic ray from pp to ξ\xi. The Busemann cocycle βξ(x,y):n×n×n\beta_{\xi}(x,y):\mathbb{H}^{n}\times\mathbb{H}^{n}\times\partial_{\infty}\mathbb{H}^{n}\rightarrow\mathbb{R} is defined by

βξ(x,y)=limt(d(ρξ(t),x)d(ρξ(t),y)).\beta_{\xi}(x,y)=\lim_{t\rightarrow\infty}(d(\rho_{\xi}(t),x)-d(\rho_{\xi}(t),y)).

For a discrete isometry subgroup Γ<Isom(n)\Gamma<\operatorname{{\mathrm{I}som}}(\mathbb{H}^{n}), there exists a family of finite measures (μx)xn(\mu_{x})_{x\in\mathbb{H}^{n}} on n\partial_{\infty}\mathbb{H}^{n} whose support is the limit set Λ(Γ)\Lambda(\Gamma) and satisfies the following conditions:

  1. (1)

    It is Γ\Gamma-invariant, i.e. γ(μx)=μγx\gamma_{\ast}(\mu_{x})=\mu_{\gamma x}.

  2. (2)

    The Radon-Nikodym derivatives exist for all x,ynx,y\in\mathbb{H}^{n}, and for all ξn\xi\in\partial_{\infty}\mathbb{H}^{n} they satisfy

    dμxdμy(ξ)=eδ(Γ)βξ(x,y).\dfrac{d\mu_{x}}{d\mu_{y}}(\xi)=e^{-\delta(\Gamma)\beta_{\xi}(x,y)}.

Such family of measures is a family of Patterson-Sullivan density of dimension δ(Γ)\delta(\Gamma) for Γ\Gamma. The Patterson-Sullivan measures have very nice properties when the group Γ\Gamma is geometrically finite.

Theorem 2.5.

[McM99, Theorem 3.1] Let Γ<Isom(n)\Gamma<\operatorname{{\mathrm{I}som}}(\mathbb{H}^{n}) be a geometrically finite Kleinian group. Then n\partial_{\infty}\mathbb{H}^{n} carries a unique Γ\Gamma-invariant density μ\mu of dimension δ(Γ)\delta(\Gamma) with total mass one; Moreover, μ\mu is nonatomic and supported on Λ(Γ)\Lambda(\Gamma), and the Poincaré series diverges at δ(Γ)\delta(\Gamma).

Theorem 2.6.

[McM99, Theorem 1.2] Suppose that (Γk<Isom(n))k(\Gamma_{k}<\operatorname{{\mathrm{I}som}}(\mathbb{H}^{n}))_{k\in\mathbb{N}} is a sequence of Kleinian groups converging strongly to Γ<Isom(n)\Gamma<\operatorname{{\mathrm{I}som}}(\mathbb{H}^{n}). If Γ\Gamma is geometrically finite with δ(Γ)>(n1)/2\delta(\Gamma)>(n-1)/2, then the Patterson-Sullivan densities μk\mu_{k} of Γk\Gamma_{k} converge to the Patterson-Sullivan density μ\mu of Γ\Gamma in the weak topology on measures.

Remark 2.7.

Theorem 1.2 in [McM99] is stated for the 33-dimensional hyperbolic space. However, the proof works exactly the same for general hyperbolic spaces n\mathbb{H}^{n}.

The proof Theorem 2.6 relies heavily on the analysis of the Poincaré series of parabolic groups and its uniform convergence. This is also essential in the later proof of the convergence of Bowen-Margulis measures and the uniform counting formulas for orthogeodesics in the rest of the paper. For readers’ convenience, we list the analytic properties of the Poincaré series corresponding to parabolic groups in the section. The details can be found in [McM99, Section 6].

Let L<Isom(n)L<\operatorname{{\mathrm{I}som}}(\mathbb{H}^{n}) be a torsion-free elementary isometry subgroup, which is either a hyperbolic group, i.e. a cyclic group generated by a loxodromic isometry, or a parabolic group. Given xnx\in\partial_{\infty}\mathbb{H}^{n} and s0s\geq 0, the absolute Poincaré series for LL is defined to be

Ps(L,x)=γL|γ(x)|s,P_{s}(L,x)=\sum_{\gamma\in L}|\gamma^{\prime}(x)|^{s},

where the derivative is measured in the spherical measure. Given any open subset UnU\subset\partial_{\infty}\mathbb{H}^{n}, define

Ps(L,U,x)=γ(x)U|γ(x)|s.P_{s}(L,U,x)=\sum_{\gamma(x)\in U}|\gamma^{\prime}(x)|^{s}.

Suppose that (Lk<Isom(n))k(L_{k}<\operatorname{{\mathrm{I}som}}(\mathbb{H}^{n}))_{k\in\mathbb{N}} is a sequence of torsion-free elementary isometry subgroups which converges geometrically to a parabolic group L<Isom(n)L<\operatorname{{\mathrm{I}som}}(\mathbb{H}^{n}) with parabolic fixed point cc, i.e., LkL_{k} converges to LL in the Hausdorff topology on closed subsets of Isom(n)\operatorname{{\mathrm{I}som}}(\mathbb{H}^{n}). The Poincaré series for (Lk,sk)(L_{k},s_{k}) where sk0s_{k}\geq 0 converges uniformly if for any compact subset Kn{c}K\subset\partial_{\infty}\mathbb{H}^{n}\setminus\{c\} and ϵ>0\epsilon>0, there is a neighborhood UU of cc such that for all xKx\in K,

Psk(Lk,U,x)<ϵP_{s_{k}}(L_{k},U,x)<\epsilon

for k0k\gg 0 sufficiently large. By using the same argument of the proof of Theorem 6.1 in [McM99], we have the following:

Theorem 2.8.

Suppose that (Γk<Isom(n))k(\Gamma_{k}<\operatorname{{\mathrm{I}som}}(\mathbb{H}^{n}))_{k\in\mathbb{N}} is a sequence of torsion-free discrete isometry subgroups which strongly converges to a geometrically finite torsion-free group Γ<Isom(n)\Gamma<\operatorname{{\mathrm{I}som}}(\mathbb{H}^{n}). Let L<ΓL<\Gamma be a parabolic subgroup and (Lk<Γk)k(L_{k}<\Gamma_{k})_{k\in\mathbb{N}} be a sequence of elementary groups which converges to LL geometrically. If

δ(Γ)>{1if n=3, or(n2)/2if n>3,\delta(\Gamma)>\begin{cases}1&\text{if $n=3$, or}\\ (n-2)/2&\text{if $n>3$,}\end{cases}

then the Poincaré series for (Lk,δ(Γk))(L_{k},\delta(\Gamma_{k})) converges uniformly to the one of (L,δ(Γ))(L,\delta(\Gamma)).

2.4. Bowen-Margulis measure

The Bowen-Margulis measure is a measure defined on the unit tangent bundle T1nT^{1}\mathbb{H}^{n} of n\mathbb{H}^{n} in terms of the Patterson-Sullivan measures. One can identify the unit tangent bundle T1nT^{1}\mathbb{H}^{n} with the set of geodesic lines l:nl:\mathbb{R}\rightarrow\mathbb{H}^{n} such that the inverse map sends the geodesic line ll to its unit tangle vector l˙(0)\dot{l}(0) at t=0t=0. Given a point x0nx_{0}\in\mathbb{H}^{n}, we can also identify T1nT^{1}\mathbb{H}^{n} with n×n×\partial_{\infty}\mathbb{H}^{n}\times\partial_{\infty}\mathbb{H}^{n}\times\mathbb{R} via the Hopf’s parametrization:

v(v,v+,t)v\rightarrow(v_{-},v_{+},t)

where v,v+v_{-},v_{+} are the endpoints at -\infty and \infty of the geodesic line defined by vv and tt is the signed distance of the closest point to x0x_{0} on the geodesic line.

We let π:T1nn\pi:T^{1}\mathbb{H}^{n}\rightarrow\mathbb{H}^{n} denote the basepoint projection. The geodesic flow on T1nT^{1}\mathbb{H}^{n} is the smooth one-parameter group of diffeomorphisms (gt)t(g^{t})_{t\in\mathbb{R}} of T1nT^{1}\mathbb{H}^{n} such that gt(l(s))=l(s+t)g^{t}(l(s))=l(s+t), for all lT1nl\in T^{1}\mathbb{H}^{n}, and s,ts,t\in\mathbb{R}. Similarly one can define the geodesic flow on T1MT^{1}M by replacing the geodesic lines ll by locally geodesic lines. The Kleinian group Γ\Gamma acts on T1nT^{1}\mathbb{H}^{n} via postcomposition, i.e. γl\gamma\circ l, and it commutes with the geodesic flow. For simplicity, we sometimes write δ(Γ)\delta(\Gamma) as δ\delta if the context is clear in the rest of the paper.

Given the Patterson-Sullivan density (μx)xn(\mu_{x})_{x\in\mathbb{H}^{n}} and a point x0nx_{0}\in\mathbb{H}^{n}, one can define the Bowen-Margulis measure m~BM\tilde{m}_{\rm{BM}} on T1nT^{1}\mathbb{H}^{n} given by

dm~BM(v)=eδ(βv(π(v),x0)+βv+(π(v),x0))dμx0(v)dμx0(v+)dt=e2δ(v|v+)x0dμx0(v)dμx0(v+)dt.d\tilde{m}_{\rm{BM}}(v)=e^{-\delta(\beta_{v_{-}}(\pi(v),x_{0})+\beta_{v_{+}}(\pi(v),x_{0}))}d\mu_{x_{0}}(v_{-})d\mu_{x_{0}}(v_{+})dt=e^{-2\delta(v_{-}|v_{+})_{x_{0}}}d\mu_{x_{0}}(v_{-})d\mu_{x_{0}}(v_{+})dt.

Here we introduce the notation (v|v+)x0=12(βv(y,x0)+βv+(y,x0))(v_{-}|v_{+})_{x_{0}}=\frac{1}{2}(\beta_{v_{-}}(y,x_{0})+\beta_{v_{+}}(y,x_{0})), where yy is any point in the geodesic joining v,v+v_{-},v_{+}. It is not hard to verify that (v|v+)x0(v_{-}|v_{+})_{x_{0}} does not depend on yy.

The Bowen-Margulis measure m~BM\tilde{m}_{\rm{BM}} is independent of the choice of x0x_{0}, and it is invariant under both the action of the group Γ\Gamma and the geodesic flow. Hence, it descends to a measure mBMm_{\rm{BM}} on T1MT^{1}M invariant under the quotient geodesic flow, which is called the Bowen-Margulis measure on T1MT^{1}M.

Theorem 2.9.

[Sul84, Bab02] Let Γ<Isom(n)\Gamma<\operatorname{{\mathrm{I}som}}(\mathbb{H}^{n}) be a geometrically finite Kleinian group. The Bowen-Margulis measure mBMm_{\rm{BM}} has finite total mass, and the geodesic flow is mixing with respect to mBMm_{\rm{BM}}.

Another related measure we consider in the paper is the so called skinning measure. Let DD be a nonempty proper closed convex subset in n\mathbb{H}^{n}. We denote its boundary by D\partial D and the set of points at infinity by D\partial_{\infty}D. Let

(1) PD:n(nD)DP_{D}:\mathbb{H}^{n}\cup(\partial_{\infty}\mathbb{H}^{n}\setminus\partial_{\infty}D)\rightarrow D

be the closest point map. In particular, for points xnx\in\mathbb{H}^{n}, PD(x)P_{D}(x) is the point on DD which minimizes the distance function d(y,x)d(y,x) for yDy\in D, and for points ξnD\xi\in\partial_{\infty}\mathbb{H}^{n}\setminus\partial_{\infty}D, PD(ξ)P_{D}(\xi) is the point yDy\in D which minimizes the function yβξ(y,x0)y\rightarrow\beta_{\xi}(y,x_{0}) for a given x0x_{0}.

The outer unit normal bundle +1D\partial^{1}_{+}D of the boundary of DD is the topological submanifold of T1nT^{1}\mathbb{H}^{n} consisting of the geodesic lines v:nv:\mathbb{R}\rightarrow\mathbb{H}^{n} such that PD(v+)=v(0)P_{D}(v_{+})=v(0). Similarly, one can define the inner unit normal bundle 1D\partial^{1}_{-}D which consists of geodesic lines vv such that PD(v)=v(0)P_{D}(v_{-})=v(0). Note that when DD is totally geodesic, +1D=1D\partial^{1}_{+}D=\partial^{1}_{-}D. Given the Patterson-Sullivan density (μx)xn(\mu_{x})_{x\in\mathbb{H}^{n}}, the outer skinning measure on +1D\partial^{1}_{+}D is the measure σ~D+\tilde{\sigma}^{+}_{D} defined by

dσ~D+(v)=eδβv+(PD(v+),x0)dμx0(v+).d\tilde{\sigma}_{D}^{+}(v)=e^{-\delta\beta_{v_{+}}(P_{D}(v_{+}),x_{0})}d\mu_{x_{0}}(v_{+}).

Similarly, one can define the inner skinning measure σ~D\tilde{\sigma}_{D}^{-} on 1D\partial^{1}_{-}D as follows:

dσ~D(v)=eδβv(PD(v),x0)dμx0(v).d\tilde{\sigma}_{D}^{-}(v)=e^{-\delta\beta_{v_{-}}(P_{D}(v_{-}),x_{0})}d\mu_{x_{0}}(v_{-}).

For simplicity, we sometimes identify a precisely invariant subset CM=n/ΓC\subset M=\mathbb{H}^{n}/\Gamma with its fundamental domain C~\tilde{C} in the universal cover, and use the notation σC±\sigma^{\pm}_{\partial C} to denote the outer/inner skinning measure σ~C~±\tilde{\sigma}^{\pm}_{\tilde{C}} on ±1C~\partial^{1}_{\pm}\tilde{C}.

2.5. Convergence of convex sets

In this subsection, we first define strong convergence that admits disconnected limits. Suppose that ((Mk,gk))k((M_{k},g_{k}))_{k\in\mathbb{N}} is a sequence of nn-manifolds of pinched sectional curvature κ2K1-\kappa^{2}\leq K\leq-1. We say that the sequence converges strongly to a (possibly disconnected) geometrically finite nn-manifold (N=imNi,g)(N=\cup_{i}^{m}N_{i},g) if the following holds:

  1. (1)

    There exist points pk,iMkp_{k,i}\in M_{k}, piNip_{i}\in N_{i} so that d(pk,i,pk,j)+d(p_{k,i},p_{k,j})\rightarrow+\infty for iji\neq j and (Mk,pk,i)(Ni,pi)(M_{k},p_{k,i})\rightarrow(N_{i},p_{i}) geometrically, i.e., there exists an exhaustion U1,iU2,iU_{1,i}\subset U_{2,i}\subset\ldots of relatively compact open sets of (Ni,pi)(N_{i},p_{i}) and smooth maps φk,i:Uk,iMk\varphi_{k,i}:U_{k,i}\rightarrow M_{k} so that φk,i(pi)=pk,i\varphi_{k,i}(p_{i})=p_{k,i} and φk,igk\varphi_{k,i}^{*}g_{k} converges smoothly in compact sets to gg.

  2. (2)

    For any ϵ\epsilon, the truncated cores C(Mk)>ϵC(M_{k})^{>\epsilon} converge to the disjoint union iC(Ni)>ϵ\cup_{i}C(N_{i})^{>\epsilon}. This means that for large kk we have C(Mk)>ϵ=i=1mC(Mk,i)>ϵC(M_{k})^{>\epsilon}=\cup_{i=1}^{m}C(M_{k,i})^{>\epsilon} where C(Mk,i)>ϵIm(φk,i)C(M_{k,i})^{>\epsilon}\subset Im(\varphi_{k,i}) and φk,i1(C(Mk,i)>ϵ)\varphi^{-1}_{k,i}(C(M_{k,i})^{>\epsilon}) converges to C(Ni)>ϵC(N_{i})^{>\epsilon} in the Hausdorff topology of compact sets in NiN_{i}.

The definition accommodates situations like Dehn drilling and pinching closed geodesics in hyperbolic 3-manifolds. The pinching case can result in disconnected limit manifolds. If MkM_{k} and NN are hyperbolic manifolds, and NN is connected, this definition is equivalent to the one described in [McM99] for strong convergence. Because of this, in the cases when NN is connected we will simply omit the mention of possibly disconnected, as well as the sub-index ii from our notation.

Moreover, given a sequence (Mk)k(M_{k})_{k\in\mathbb{N}} converging strongly to a possibly disconnected manifold NN, we say that the sequence of functions (fk:Mk)k(f_{k}:M_{k}\rightarrow\mathbb{R})_{k\in\mathbb{N}} converges strongly to a function f:Nf:N\rightarrow\mathbb{R} if, with the notation above, we have that for any basepoint pip_{i} the sequence (fkφk,i)k(f_{k}\circ\varphi_{k,i})_{k\in\mathbb{N}} converges smoothly in compact sets to ff. Similarly, if Σk,Σ\Sigma_{k},\Sigma are smooth properly embedded submanifolds in Mk,NM_{k},N, we say that (Σk)k(\Sigma_{k})_{k\in\mathbb{N}} converges strongly to Σ\Sigma if φk,i1(Σk)\varphi_{k,i}^{-1}(\Sigma_{k}) converges smoothly in compact sets to Σ\Sigma. Since for any fixed compact set in NiN_{i} the maps φk,i\varphi_{k,i} are embeddings for kk sufficiently large, we can define strong convergence of functions and submanifolds of T1MkT^{1}M_{k} to T1MT^{1}M by composing the derivatives of φk,i\varphi_{k,i} with the projections from TMkT^{*}M_{k} to T1MkT^{1}M_{k}.

Using the definition of strong convergence, we obtain a straightforward corollary:

Corollary 2.10.

Suppose that (Mk)k(M_{k})_{k\in\mathbb{N}} is a sequence of manifolds with negatively pinched curvature which converges strongly to a (possibly disconnected) geometrically finite manifold NN with negatively pinched curvature. Then the manifolds MkM_{k} are also geometrically finite for sufficiently large kk.

Proof.

Suppose that N=imNiN=\cup_{i}^{m}N_{i}. The truncated core C(N)>ϵC(N)^{>\epsilon} is compact for any 0<ϵ<ϵ(n,κ)0<\epsilon<\epsilon(n,\kappa), since NN is geometrically finite. By item (2) in the definition of strong convergence, C(Mk)>ϵC(M_{k})^{>\epsilon} is also compact for large kk, since C(Mk,i)>ϵC(M_{k,i})^{>\epsilon} is compact for large kk, and all 1im1\leq i\leq m. ∎

In Section 3, we work on sequences of manifolds of negatively pinched curvature that converge strongly to (possibly disconnected) limit manifolds. Given an nn-dimensional manifold MM with negatively pinched curvature (possibly disconnected) and a constant μ<(n1)2/4\mu<(n-1)^{2}/4, Specμ(M)Spec_{\mu}(M) is defined as the collection of eigenvalues of the negative Laplacian on MM less than μ\mu. If MM is disconnected, Specμ(M)Spec_{\mu}(M) agrees with the union of SpecμSpec_{\mu} of each component of MM (counting multiplicity). Specifically, a function f:Mf:M\rightarrow\mathbb{R} satisfies the equation Δf=λf-\Delta f=\lambda f if and only if its restriction to each component of MM is either an eigenfunction with eigenvalue λ\lambda, or 0. Moreover, while taking orthonormal eigenfunctions for MM we can consider that each eigenfunction has support in a unique component of MM.

In Section 4 and Section 5, we focus on sequences of hyperbolic manifolds strongly converging to connected limit manifolds. Suppose now M=n/ΓM=\mathbb{H}^{n}/\Gamma is an nn-dimensional hyperbolic manifold. As we stated in the Introduction, locally convex sets in MM are in 1-to-1 correspondence with Γ\Gamma-precisely invariant convex sets in n\mathbb{H}^{n} by the projection map Proj:nM\textup{Proj}:\mathbb{H}^{n}\rightarrow M. In particular, we sometimes identify local convex sets with one of their lifts which are Γ\Gamma-precisely invariant, and we don’t consider immersed locally convex sets, e.g. nonprimitive closed geodesics. For simplicity, we will omit the word locally and plainly denote the sets as convex.

We say that a convex set DD in MM is well-positioned if D~\partial\tilde{D} is smooth, where D~\tilde{D} denotes the lift of DD to 3\mathbb{H}^{3}, and σD±\sigma^{\pm}_{\partial D} has compact support.

Example 2.11.

Suppose that MM is a geometrically finite hyperbolic manifold. Embedded geodesic balls and the thin part of MM are well positioned convex sets.

Proof.

Geodesic balls with radii smaller than the injectivity radius of the center and Margulis tubes are compact convex subsets, so they are well-positioned. The lifts of a cusp neighbourhood DD in MM are horoballs whose boundaries are smooth. Since MM is geometrically finite, all parabolic fixed points are bounded. Hence, the intersection of D\partial D with the convex core is compact. Thus, σD±\sigma^{\pm}_{\partial D} has compact support and DD is well-positioned.

Suppose that (Mk=n/Γk)k(M_{k}=\mathbb{H}^{n}/\Gamma_{k})_{k\in\mathbb{N}} is a sequence of hyperbolic manifolds that converges strongly to a geometrically finite hyperbolic manifold M=n/ΓM=\mathbb{H}^{n}/\Gamma. We say that well-positioned convex sets DkMkD_{k}\subset M_{k} strongly converge to a well-positioned convex set DMD\subset M if

  1. (1)

    the boundary Dk\partial D_{k} converges strongly to D\partial D, or equivalently, the lifts of φk1(Dk)\varphi_{k}^{-1}(\partial D_{k}) converge smoothly in compact sets to lifts of D\partial D, where φk:UkMk\varphi_{k}:U_{k}\rightarrow M_{k} are the smooth maps in the definition of strong convergence of (Mk)k(M_{k})_{k\in\mathbb{N}},

  2. (2)

    π¯(supp(σDk±))\bar{\pi}(supp(\sigma^{\pm}_{\partial D_{k}})) is contained in φk(N1(π¯(supp(σD±))))\varphi_{k}\left(N_{1}(\bar{\pi}(supp(\sigma^{\pm}_{\partial D})))\right) for large kk, where π¯:T1MM\bar{\pi}:T^{1}M\rightarrow M and N1N_{1} denotes the 1-neighborhood.

Example 2.12.

Let (Mk=n/Γk)k(M_{k}=\mathbb{H}^{n}/\Gamma_{k})_{k\in\mathbb{N}} be a sequence of hyperbolic manifolds that converges strongly to a geometrically finite hyperbolic manifold M=n/ΓM=\mathbb{H}^{n}/\Gamma.

  1. (1)

    Suppose that (xk)k(x_{k})_{k\in\mathbb{N}} is a sequence of points in MkM_{k} that converges to xMx\in M. The geodesic balls around the xkx_{k} with radius rr converge strongly to the geodesic ball of xx with the same radius, where rr is smaller than the injectivity radius of xx.

  2. (2)

    Given 0<ϵ<ϵ(n,κ)0<\epsilon<\epsilon(n,\kappa), the thin parts Mk<ϵM_{k}^{<\epsilon} converge strongly to the thin part M<ϵM^{<\epsilon}.

3. Convergence of small eigenvalues

In this section, we study the convergence of small eigenvalues and prove the uniform spectral gap for strongly convergent sequences of geometrically finite nn-manifolds of negatively pinched curvature κ2K1-\kappa^{2}\leq K\leq-1.

Proposition 3.1.

Let MM be a geometrically finite Riemannian nn-manifold of pinched sectional curvature κ2K1-\kappa^{2}\leq K\leq-1 and let ϵ=ϵ(n,κ)>0\epsilon=\epsilon(n,\kappa)>0 be the Margulis constant. Given μ<(n1)2/4\mu<(n-1)^{2}/4, there exists a sufficiently large constant r(μ)=r>0r(\mu)=r>0 and some constant η(μ,r)=η>0\eta(\mu,r)=\eta>0, so that if fH1(M)f\in H^{1}(M) with R(f)μR(f)\leq\mu, then B2r(C(M)>ϵ)|f|2ηM|f|2\int_{B_{2r}(C(M)^{>\epsilon})}|f|^{2}\geq\eta\int_{M}|f|^{2}. Moreover, one can take η(14μ(n1)2)\eta\rightarrow\bigg{(}1-\frac{4\mu}{(n-1)^{2}}\bigg{)} as r+r\rightarrow+\infty.

Proof.

Since C0(M)C^{\infty}_{0}(M) is dense in H1(M)H^{1}(M), we can assume without loss of generality that ff is compactly supported with Mf2=1\int_{M}f^{2}=1.

Observe that we can find C1C^{1} functions g,h:[0,1]g,h:\mathbb{R}\rightarrow[0,1] so that

  • g2(x)+h2(x)=1g^{2}(x)+h^{2}(x)=1 for any xx\in\mathbb{R},

  • supp(g)(,1],supp(h)(0,+]supp(g)\subseteq(-\infty,1],\,supp(h)\subseteq(0,+\infty].

Given a positive constant r>0r>0, we define u:=ur(x),α:=αr(x)C01(M)u:=u_{r}(x),\alpha:=\alpha_{r}(x)\in C^{1}_{0}(M) satisfying the following properties, by using scalings of g,hg,h along equidistant sets to C(M)\partial C(M):

  1. (1)

    0u,α10\leq u,\alpha\leq 1

  2. (2)

    supp(u)B2r(C(M))supp(u)\subseteq B_{2r}(C(M))

  3. (3)

    supp(α)int(Brc(C(M)))supp(\alpha)\subseteq int(B^{c}_{r}(C(M)))

  4. (4)

    u2+α21u^{2}+\alpha^{2}\equiv 1

  5. (5)

    |u|,|α|Cr|\nabla u|,|\nabla\alpha|\leq\frac{C}{r} everywhere in MM, for some constant CC independent of rr and MM.

Similarly, for the thick-thin decomposition of MM we define functions v:=vr(x),β:=βr(x)C01(M)v:=v_{r}(x),\beta:=\beta_{r}(x)\in C^{1}_{0}(M) along equidistant sets to M>ϵ\partial M^{>\epsilon} satisfying the following properties

  1. (1)

    0v,β10\leq v,\beta\leq 1

  2. (2)

    supp(v)B2r(M>ϵ)supp(v)\subseteq B_{2r}(M^{>\epsilon})

  3. (3)

    supp(β)int(Brc(M>ϵ))supp(\beta)\subseteq int(B^{c}_{r}(M^{>\epsilon}))

  4. (4)

    v2+β21v^{2}+\beta^{2}\equiv 1

  5. (5)

    |v|,|β|Cr|\nabla v|,|\nabla\beta|\leq\frac{C}{r} everywhere in MM, for some constant CC independent of rr and MM.

Define then f1:=uvf,f2:=uβf,f3:=αff_{1}:=uvf,\,f_{2}:=u\beta f,\,f_{3}:=\alpha f which are in C01(M)C^{1}_{0}(M). By the definitions of u,α,v,βu,\alpha,v,\beta we have

  1. (1)

    supp(f1)B2r(M>ϵC(M))supp(f_{1})\subseteq B_{2r}(M^{>\epsilon}\cap C(M))

  2. (2)

    supp(f2)Brc(M>ϵ)supp(f_{2})\subseteq B^{c}_{r}(M^{>\epsilon})

  3. (3)

    supp(f3)Brc(C(M))supp(f_{3})\subseteq B^{c}_{r}(C(M))

  4. (4)

    f12+f22+f32=f2f^{2}_{1}+f^{2}_{2}+f^{2}_{3}=f^{2}.

We can expand R(f1)R(f_{1}) as

R(f1)=(Mf2|(uv)|2+2uvf(uv),f+u2v2|f|2)/(Mu2v2f2).R(f_{1})=\bigg{(}\int_{M}f^{2}|\nabla(uv)|^{2}+2uvf\langle\nabla(uv),\nabla f\rangle+u^{2}v^{2}|\nabla f|^{2}\bigg{)}\bigg{/}\bigg{(}\int_{M}u^{2}v^{2}f^{2}\bigg{)}.

Since (uv)=uv+vu\nabla(uv)=u\nabla v+v\nabla u then it follows that |(uv)|2Cr|\nabla(uv)|\leq\frac{2C}{r}, and subsequently

Mf2|(uv)|24C2r2.\int_{M}f^{2}|\nabla(uv)|^{2}\leq\frac{4C^{2}}{r^{2}}.

By Cauchy-Schwarz, we also have that

|M2uvf(uv),f|2M|f(uv),f|4CrR(f).\bigg{|}\int_{M}2uvf\langle\nabla(uv),\nabla f\rangle\bigg{|}\leq 2\int_{M}|\langle f\nabla(uv),\nabla f\rangle|\leq\frac{4C}{r}\sqrt{R(f)}.

Collecting these inequalities and defining a:=Mf12a:=\int_{M}f^{2}_{1} for convenience, we arrive to

(2) R(f1)(4C2r2+4CrR(f)+Mu2v2|f|2)/a.R(f_{1})\leq\bigg{(}\frac{4C^{2}}{r^{2}}+\frac{4C}{r}\sqrt{R(f)}+\int_{M}u^{2}v^{2}|\nabla f|^{2}\bigg{)}\bigg{/}a.

Similarly, define b:=Mf22,c:=Mf32b:=\int_{M}f^{2}_{2},\,c:=\int_{M}f^{2}_{3}. Then

(3) R(f2)(4C2r2+4CrR(f)+Mu2β2|f|2)/b,R(f_{2})\leq\bigg{(}\frac{4C^{2}}{r^{2}}+\frac{4C}{r}\sqrt{R(f)}+\int_{M}u^{2}\beta^{2}|\nabla f|^{2}\bigg{)}\bigg{/}b,
(4) R(f3)(4C2r2+4CrR(f)+Mα2|f|2)/c.R(f_{3})\leq\bigg{(}\frac{4C^{2}}{r^{2}}+\frac{4C}{r}\sqrt{R(f)}+\int_{M}\alpha^{2}|\nabla f|^{2}\bigg{)}\bigg{/}c.

By doing a(2)+b(3)+c(4)a(\ref{eq:Rf1bound})+b(\ref{eq:Rf2bound})+c(\ref{eq:Rf3bound}) we obtain

(5) aR(f1)+bR(f2)+cR(f3)\displaystyle aR(f_{1})+bR(f_{2})+cR(f_{3}) (12C2r2+12CrR(f)+M(u2v2+u2β2+α2)|f|2)\displaystyle\leq\bigg{(}\frac{12C^{2}}{r^{2}}+\frac{12C}{r}\sqrt{R(f)}+\int_{M}(u^{2}v^{2}+u^{2}\beta^{2}+\alpha^{2})|\nabla f|^{2}\bigg{)}
=(12C2r2+12CrR(f)+M|f|2)\displaystyle=\bigg{(}\frac{12C^{2}}{r^{2}}+\frac{12C}{r}\sqrt{R(f)}+\int_{M}|\nabla f|^{2}\bigg{)}
(12C2r2+12Crμ+μ).\displaystyle\leq\bigg{(}\frac{12C^{2}}{r^{2}}+\frac{12C}{r}\sqrt{\mu}+\mu\bigg{)}.

Since supp(f2)int(Brc(M>ϵ)),supp(f3)int(Brc(C(M)))supp(f_{2})\subseteq int(B^{c}_{r}(M^{>\epsilon})),\,supp(f_{3})\subseteq int(B^{c}_{r}(C(M))) we have by Lemmas 2.2 and 2.3 (or more precisely, by applying a combination of the Lemmas on each component of MM) that R(f2)(tanhr)2(n1)2/4,R(f3)(tanhr)2(n1)2/4R(f_{2})\geq(\tanh r)^{2}(n-1)^{2}/4,\,R(f_{3})\geq(\tanh r)^{2}(n-1)^{2}/4. Using these bounds together with the obvious bound aR(f1)0aR(f_{1})\geq 0 we arrive to

(b+c)(tanhr)2(n1)2412C2r2+12Crμ+μ,(b+c)\frac{(\tanh r)^{2}(n-1)^{2}}{4}\leq\frac{12C^{2}}{r^{2}}+\frac{12C}{r}\sqrt{\mu}+\mu,
(b+c)4(tanhr)2(n1)2(12C2r2+12Crμ+μ).(b+c)\leq\frac{4}{(\tanh r)^{2}(n-1)^{2}}\bigg{(}\frac{12C^{2}}{r^{2}}+\frac{12C}{r}\sqrt{\mu}+\mu\bigg{)}.

By the fact that a+b+c=1a+b+c=1 we obtain

a4(tanhr)2(n1)2((tanhr)2(n1)2412C2r212Crμμ).a\geq\frac{4}{(\tanh r)^{2}(n-1)^{2}}\bigg{(}\frac{(\tanh r)^{2}(n-1)^{2}}{4}-\frac{12C^{2}}{r^{2}}-\frac{12C}{r}\sqrt{\mu}-\mu\bigg{)}.

The result follows from observing that for the left-hand side we have B2r(M>ϵC(M))f2a\int_{B_{2r}(M^{>\epsilon}\cap C(M))}f^{2}\geq a, whereas the right hand side depends only on μ,r\mu,r and converges to 14μ(n1)2>01-\frac{4\mu}{(n-1)^{2}}>0 as r+r\rightarrow+\infty.

Now we use Proposition 3.1 to take limits of eigenfunctions with small eigenvalues along a strongly convergent sequence of manifolds with negatively pinched curvature.

Lemma 3.2.

Suppose that (Mk)k(M_{k})_{k\in\mathbb{N}} is a sequence of nn-manifolds of pinched curvature κ2K1-\kappa^{2}\leq K\leq-1 that converges strongly to a (possibly disconnected) geometrically finite nn-manifold MM. Let μ<(n1)2/4\mu<(n-1)^{2}/4 and for each MkM_{k}, let fkf_{k} be an eigenfunction of the negative Laplacian so that R(fk)μR(f_{k})\leq\mu and Mk|fk|2=1\int_{M_{k}}|f_{k}|^{2}=1. Then, after possibly taking a subsequence, we have that fkf_{k} converges strongly to ff, a non-zero eigenfunction of the negative Laplacian in MM with R(f)μR(f)\leq\mu.

Proof.

By Proposition 3.1 there exist r>0r>0 and η>0\eta>0 independent of kk so that B2r(C(Mk)>ϵ)|fk|2η\int_{B_{2r}(C(M_{k})^{>\epsilon})}|f_{k}|^{2}\geq\eta. By elliptic regularity and strong convergence, we have that the Sobolev norms

fkW2,(B2r(C(Mk)>ϵ))\|f_{k}\|_{W^{2,\ell}(B_{2r}(C(M_{k})^{>\epsilon}))}

are uniformly bounded for any given \ell. By the Rellich-Kondrachov compactness theorem, we can take a convergent subsequence with limit ff in B2r(C(M)>ϵ)B_{2r}(C(M)^{>\epsilon}) in any W2,W^{2,\ell} norm. Taking r+r\rightarrow+\infty and doing a Cantor diagonal argument, we have that R(f)μ,ΔMf=R(f)f,M|f|2ηR(f)\leq\mu,\,-\Delta_{M}f=R(f)f,\,\int_{M}|f|^{2}\geq\eta, which concludes the Lemma. ∎

Recall that Specμ(M)Spec_{\mu}(M) denotes the collection of eigenvalues of the negative Laplacian on the negatively pinched manifold MM which are smaller than μ\mu, where for convenience we assume that μ<(n1)2/4\mu<(n-1)^{2}/4 is not an eigenvalue of MM (this is possible for all μ<(n1)2\mu<(n-1)^{2}/4 with the exception of finitely many values). Suppose that (Mk)k(M_{k})_{k\in\mathbb{N}} is a sequence of negatively pinched manifolds which converges strongly to a geometrically finite nn-manifold MM. Given any small eigenvalue λSpecμ(M)\lambda\in Spec_{\mu}(M), we can use the discreteness of small eigenvalues to take ϵ>0\epsilon>0 small enough so that (λϵ,λ+ϵ)Specμ(M)={λ}(\lambda-\epsilon,\lambda+\epsilon)\cap Spec_{\mu}(M)=\{\lambda\}. We have then that (λϵ,λ+ϵ)Specμ(Mk)(\lambda-\epsilon,\lambda+\epsilon)\cap Spec_{\mu}(M_{k}) is either empty or accumulates to λ\lambda as kk\rightarrow\infty, where we desire to prove the later case. Let then mλm_{\lambda} be the multiplicity of λ\lambda and mλ,km_{\lambda,k} be the cardinality of (λϵ,λ+ϵ)Specμ(Mk)(\lambda-\epsilon,\lambda+\epsilon)\cap Spec_{\mu}(M_{k}) (counting multiplicities). We say Specμ(Mk)Spec_{\mu}(M_{k}) converges to Specμ(M)Spec_{\mu}(M), if limkmλ,k=mλ\lim_{k\rightarrow\infty}m_{\lambda,k}=m_{\lambda} for any small eigenvalue λSpecμ(M)\lambda\in Spec_{\mu}(M).

Theorem 3.3.

Suppose that (Mk)k(M_{k})_{k\in\mathbb{N}} is a sequence of nn-manifolds of pinched curvature κ2K1-\kappa^{2}\leq K\leq-1 that converges strongly to a (possibly disconnected) geometrically finite nn-manifold MM. Then for any given μ<(n1)2/4\mu<(n-1)^{2}/4 not in Spec(M)Spec(M) we have that Specμ(Mk)Spec_{\mu}(M_{k}) converges (counting multiplicities) to Specμ(M)Spec_{\mu}(M).

Proof.

To prove the theorem, we will show the convergence of eigenspaces. Namely, let Vk,VV_{k},V denote the linear spaces of functions generated by the eigenfunctions with eigenvalues in Specμ(Mk)Spec_{\mu}(M_{k}) and Specμ(M)Spec_{\mu}(M), which have a natural orthogonal decomposition by the eigenspaces of Specμ(Mk)Spec_{\mu}(M_{k}) and Specμ(M)Spec_{\mu}(M). We show that VkVV_{k}\rightarrow V, in the following sense:

  1. (1)

    Any function fVf\in V can be obtained as the limit of a strongly convergent sequence (fkVk)k(f_{k}\in V_{k})_{k\in\mathbb{N}}.

  2. (2)

    Any sequence of families (fl,kVk)k(f_{l,k}\subset V_{k})_{k\in\mathbb{N}} of orthonormal functions in MkM_{k} converges strongly (after possibly taking a subsequence) to a linearly independent family of functions in MM.

Item (1) implies that lim infkmλ,kmλ\liminf_{k\rightarrow\infty}m_{\lambda,k}\geq m_{\lambda}, and Item (2) implies that lim supkmλ,kmλ\limsup_{k\rightarrow\infty}m_{\lambda,k}\leq m_{\lambda}. Thus, it suffices to prove the convergence of eigenspaces. We first show item (2). Suppose that f1,k,fl,kf_{1,k},\ldots f_{l,k} are orthonormal eigenfunctions of MkM_{k}. By Lemma 3.2 we can assume they converge in compact sets to f1,,flf_{1},\ldots,f_{l}. If the functions f1,,flf_{1},\ldots,f_{l} are not linearly independent in L2(M)L^{2}(M), there exist real numbers α1,,αl\alpha_{1},\ldots,\alpha_{l} not all vanishing so that α1f1++αlfl0\alpha_{1}f_{1}+\ldots+\alpha_{l}f_{l}\equiv 0. Hence, gk=α1f1,k++αlfl,kg_{k}=\alpha_{1}f_{1,k}+\ldots+\alpha_{l}f_{l,k} are functions in H1(Mk)H^{1}(M_{k}) with norm α12++αk20\sqrt{\alpha_{1}^{2}+\ldots+\alpha_{k}^{2}}\neq 0. We can normalize gkL2(M)=1\|g_{k}\|_{L^{2}(M)}=1 so that R(gk)μR(g_{k})\leq\mu, and since the limit of gkg_{k} in compact sets is not identically zero from Proposition 3.1, we have a contradiction.

Now we prove Item (1). Assume that not all functions in VV are obtained as limits of functions in VkV_{k}. Let VV^{\prime} be the proper maximal space in VV, consisting of functions that can be obtained as limits. Assume that there exists an eigenfunction ff of MM with eigenvalue λ\lambda, such that ff is orthogonal to VV^{\prime}. Approximate ff in H1(M)H^{1}(M) by a compactly supported function f0f_{0}, which is normalized so that M|f0|2=1\int_{M}|f_{0}|^{2}=1 and R(f0)R(f_{0}) is close to λ\lambda. It follows that Mf0g=M(f0f)g\int_{M}f_{0}g=\int_{M}(f_{0}-f)g can be taken uniformly small for all gVg\in V^{\prime} with M|g|2=1\int_{M}|g|^{2}=1. Let f0kf^{k}_{0} be the pullback of f0f_{0} in MkM_{k} by the maps φk,i\varphi_{k,i} from the definition of strong convergence. Then for sufficiently large kk we have that (after identifying the compact cores) Mkf0kgk\int_{M_{k}}f_{0}^{k}g_{k} can be also taken uniformly small for any gkVkg_{k}\in V_{k} with Mk|gk|2=1\int_{M_{k}}|g_{k}|^{2}=1 by Proposition 3.1. For large kk we also have that in MkM_{k} the Rayleigh quotient R(f0k)R(f^{k}_{0}) is close to λ\lambda. Denote then by f0,kf_{0,k} the projection of f0kf_{0}^{k} perpendicular to VkV_{k}. Then R(f0,k)R(f_{0,k}) is also very close to λ\lambda for sufficiently large kk. Hence, this contributes to an eigenfunction in MkM_{k} which does not belong to VkV_{k}. However, by construction, VkV_{k} is the linear space of functions generated by eigenfunctions with eigenvalues in Specμ(Mk)Spec_{\mu}(M_{k}), which gives a contradiction. Therefore, any function fVf\in V can be obtained as the limit of a strongly convergent sequence (fkVk)(f_{k}\in V_{k}).

Recall that the Lax-Phillips spectral gap s1=min{λ1(M),(n1)2/4}λ0(M)s_{1}=\min\{\lambda_{1}(M),(n-1)^{2}/4\}-\lambda_{0}(M) for a hyperbolic manifold M=n/ΓM=\mathbb{H}^{n}/\Gamma. We obtain the following convergence result of spectral gap for strongly convergent sequences of hyperbolic manifolds.

Theorem 3.4.

Suppose that (Mk=Isom(n)/Γk)k(M_{k}=\operatorname{{\mathrm{I}som}}(\mathbb{H}^{n})/\Gamma_{k})_{k\in\mathbb{N}} is a sequence of hyperbolic manifolds which converges strongly to a geometrically finite hyperbolic manifold M=n/ΓM=\mathbb{H}^{n}/\Gamma. Then the sequence of Lax-Phillips spectral gaps s1(Mk)s_{1}(M_{k}) converges to s1(M)s_{1}(M).

Proof.

By [McM99, Theorem 1.5] and Theorem 2.4 we have that limkλ0(Mk)=λ0(M)\lim_{k\rightarrow\infty}\lambda_{0}(M_{k})=\lambda_{0}(M). By Theorem 3.3, if λ1(M)(n1)2/4\lambda_{1}(M)\geq(n-1)^{2}/4, then lim infλ1(Mk)(n1)2/4\liminf\lambda_{1}(M_{k})\geq(n-1)^{2}/4 for sufficiently large kk, or if λ1(M)<(n1)2/4\lambda_{1}(M)<(n-1)^{2}/4, we have that limkλ1(Mk)=λ1(M)>λ0(M)\lim_{k\rightarrow\infty}\lambda_{1}(M_{k})=\lambda_{1}(M)>\lambda_{0}(M) for sufficiently large kk. In either case, the convergence of s1(Mk)s_{1}(M_{k}) to s1(M)s_{1}(M) follows. ∎

Proof of Theorem 1.1: The proof follows from Theorem 3.3 and Theorem 3.4.

4. Uniform convergence of measures

In this section, we prove convergence for skinning measures and the Bowen-Margulis measure under strong convergence. We assume that MM is a hyperbolic nn-manifold, and by M<ϵM^{<\epsilon} we denote the ϵ\epsilon-thin part of MM for a constant ϵ\epsilon smaller than the nn-dimensional Margulis constant. We first prove that the Bowen-Margulis measure of the thin part is (uniformly) relatively small.

Proposition 4.1.

Suppose (Mk=n/Γk)k(M_{k}=\mathbb{H}^{n}/\Gamma_{k})_{k\in\mathbb{N}} is a sequence of hyperbolic manifolds that strongly converges to a geometrically finite hyperbolic manifold M=n/ΓM=\mathbb{H}^{n}/\Gamma with δ(Γ)>(n1)/2\delta(\Gamma)>(n-1)/2. Let mBMk,mBMm^{k}_{\rm BM},m_{\rm BM} be the Bowen-Margulis measures on T1MkT^{1}M_{k} and T1MT^{1}M, respectively. Then for any α>0\alpha>0 there exist ϵ>0\epsilon>0 and N>0N>0 so that for ϵ<ϵ\epsilon^{\prime}<\epsilon and k>Nk>N we have that

T1Mk<ϵ𝑑mBMk<α.\int_{T^{1}M^{<\epsilon^{\prime}}_{k}}dm^{k}_{\rm BM}<\alpha.
Proof.

This follows Dalbo-Otal-Peigne’s proof [DOP00] on the finiteness of mBMm_{\rm BM}. We first let ϵ>0\epsilon>0 be a constant which is smaller than the shortest geodesic in MM. Take a fundamental domain FF for the convex core of MM in the universal cover n\mathbb{H}^{n}, and divide FF as the thin part F<ϵF^{<\epsilon} (i.e., the intersection of FF with the thin part of MM) and the thick part F>ϵF^{>\epsilon}. Consider a component DD of F<ϵF^{<\epsilon}, which must be a cuspidal component. Suppose that \mathcal{H} is the corresponding horoball based at the parabolic fixed point ξ\xi, so that DD is a fundamental domain for the parabolic subgroup 𝒫<π1(M)\mathcal{P}<\pi_{1}(M) that preserves \mathcal{H}.

As detailed in [DOP00, page 118] we can bound m~BM\tilde{m}_{\rm BM} in DD by

m~BM(T1D)p𝒫𝒟×p𝒟cμ(dηdη+)(ηη+)𝑑t,\tilde{m}_{\rm BM}(T^{1}D)\leq\sum_{p\in\mathcal{P}}\int_{\mathcal{D}\times p\mathcal{D}}c^{\mu}(d\eta^{-}d\eta^{+})\int_{(\eta^{-}\eta^{+})\cap\mathcal{H}}dt,

where cμ(dηdη+)=e2δ(Γ)(η|η+)xdμx(η)dμx(η+)c^{\mu}(d\eta^{-}d\eta^{+})=e^{-2\delta(\Gamma)(\eta^{-}|\eta^{+})_{x}}d\mu_{x}(\eta^{-})d\mu_{x}(\eta^{+}) for a given point xnx\in\mathbb{H}^{n} and 𝒟n{ξ}\mathcal{D}\subseteq\partial_{\infty}\mathbb{H}^{n}\setminus\{\xi\} is a compact set such that {p𝒟}p𝒫\{p\mathcal{D}\}_{p\in\mathcal{P}} covers Λ(M){ξ}\Lambda(M)\setminus\{\xi\}. The existence of the compact set 𝒟\mathcal{D} is ensured by the assumption that MM is geometrically finite, hence the parabolic fixed point ξ\xi is bounded [Bow93]. Now, let 𝒫k\mathcal{P}_{k} be the elementary group in MkM_{k} that converges to 𝒫\mathcal{P}, which is either parabolic or loxodromic. We discuss the proof that when the groups 𝒫k\mathcal{P}_{k} are loxodromic. The argument for parabolic subgroups is similar.

Let k\mathcal{H}_{k} be a neighborhood of the geodesic ξkξk+\xi^{-}_{k}\xi^{+}_{k} preserved by 𝒫k\mathcal{P}_{k} so that k\mathcal{H}_{k}\rightarrow\mathcal{H}, ξk±ξ\xi^{\pm}_{k}\rightarrow\xi. Since MkM_{k} converges to MM strongly, by [McM99] we can take 𝒟\mathcal{D} large enough so that {p𝒟}p𝒫k\{p\mathcal{D}\}_{p\in\mathcal{P}_{k}} covers Λ(Mk){ξk,ξk+}\Lambda(M_{k})\setminus\{\xi^{-}_{k},\xi^{+}_{k}\}. Hence it follows that

mBMk(T1(k/𝒫k))p𝒫k𝒟×p𝒟ckμ(dηdη+)(ηη+)k𝑑t.m^{k}_{\rm BM}(T^{1}(\mathcal{H}_{k}/\mathcal{P}_{k}))\leq\sum_{p\in\mathcal{P}_{k}}\int_{\mathcal{D}\times p\mathcal{D}}c^{\mu}_{k}(d\eta^{-}d\eta^{+})\int_{(\eta^{-}\eta^{+})\cap\mathcal{H}_{k}}dt.

Assume without loss of generality that we can take a common point xk,x\in\mathcal{H}_{k},\mathcal{H}. There exist compact set KnK\subset\mathbb{H}^{n} and open neighbourhood VnV\subseteq\mathbb{H}^{n} of ξ\xi so that for kk large, if the (oriented) geodesic ηη+\eta^{-}\eta^{+} with η𝒟Λ(Γk)\eta^{-}\in\mathcal{D}\cap\Lambda(\Gamma_{k}) and η+p𝒟Λ(Γk)\eta^{+}\in p\mathcal{D}\cap\Lambda(\Gamma_{k}) intersects k\mathcal{H}_{k}, then the point of entry belongs to KkK\cap\partial\mathcal{H}_{k} and p1xp^{-1}x belongs to VV. In particular such geodesic ηη+\eta^{-}\eta^{+} verifies 0(η|η+)xdiam(K)0\leq(\eta^{-}|\eta^{+})_{x}\leq diam(K). Moreover, we have that |(ηη+)k𝑑td(x,px)|<2diam(K)|\int_{(\eta^{-}\eta^{+})\cap\mathcal{H}_{k}}dt-d(x,px)|<2diam(K). Hence there exists a constant C>0C>0 depending only on diam(K)diam(K) so that

mBMk(T1(k/𝒫k))C(p𝒫kμxk(𝒟)μxk(p𝒟)(d(x,px)+C)),m^{k}_{\rm BM}(T^{1}(\mathcal{H}_{k}/\mathcal{P}_{k}))\leq C\left(\sum_{p\in\mathcal{P}^{\prime}_{k}}\mu^{k}_{x}(\mathcal{D})\mu^{k}_{x}(p\mathcal{D})(d(x,px)+C)\right),

where μxk\mu^{k}_{x} denotes the Patterson-Sullivan measure on MkM_{k} and 𝒫k\mathcal{P}^{\prime}_{k} is the subset of {p𝒫k|p1xV}\{p\in\mathcal{P}_{k}\,|\,p^{-1}x\in V\} so that the summand 𝒟×p𝒟ckμ(dηdη+)(ηη+)k𝑑t\int_{\mathcal{D}\times p\mathcal{D}}c^{\mu}_{k}(d\eta^{-}d\eta^{+})\int_{(\eta^{-}\eta^{+})\cap\mathcal{H}_{k}}dt is non-zero.

Recall that

μxk(p𝒟)=𝒟eδ(Γk)Bη(p1x,x)μxk(dη),\mu^{k}_{x}(p\mathcal{D})=\int_{\mathcal{D}}e^{-\delta(\Gamma_{k})B_{\eta}(p^{-1}x,x)}\mu^{k}_{x}(d\eta),

so we would like to estimate Bη(p1x,x)B_{\eta}(p^{-1}x,x). Observe that as k\mathcal{H}_{k} is preserve by 𝒫k\mathcal{P}_{k}, we have that if ηη+\eta^{-}\eta^{+} is a geodesic with η𝒟Λ(Γk)\eta^{-}\in\mathcal{D}\cap\Lambda(\Gamma_{k}) and η+p𝒟Λ(Γk)\eta^{+}\in p\mathcal{D}\cap\Lambda(\Gamma_{k}) that intersects k\mathcal{H}_{k}, then the exit point of ηη+\eta^{-}\eta^{+} from k\mathcal{H}_{k} belongs to pKkpK\cap\partial\mathcal{H}_{k}. By triangular inequality we have that under such conditions |(ηη+)k𝑑tBη(x,px)|<2diam(K)|\int_{(\eta^{-}\eta^{+})\cap\mathcal{H}_{k}}dt-B_{\eta}(x,px)|<2diam(K). Hence for p𝒫kp\in\mathcal{P}^{\prime}_{k} have |Bη(p1x,x)d(p1x,x)|4diam(K)|B_{\eta}(p^{-1}x,x)-d(p^{-1}x,x)|\leq 4diam(K). Combining this with our previous inequality (and making the domain of the sum bigger if necessary) we get

mBMk(T1(k/𝒫k))C(p𝒫k,p1xV(μxk(𝒟))2eδ(Γk)d(p1x,x)(d(x,px)+C)).m^{k}_{\rm BM}(T^{1}(\mathcal{H}_{k}/\mathcal{P}_{k}))\leq C\left(\sum_{p\in\mathcal{P}_{k},p^{-1}x\in V}(\mu^{k}_{x}(\mathcal{D}))^{2}e^{-\delta(\Gamma_{k})d(p^{-1}x,x)}(d(x,px)+C)\right).

for some C>0C>0 independent of ϵ\epsilon and kk.

We claim that the above discussion holds for smaller ϵ\epsilon corresponding to a smaller neighborhood V(ϵ)V(\epsilon) for the same basepoint xx. Consider a smaller thin part corresponding to ϵ<ϵ\epsilon^{\prime}<\epsilon. The sets k,K\mathcal{H}_{k},K vary with ϵ\epsilon^{\prime}, although it is clear that k(ϵ)k(ϵ)\mathcal{H}_{k}(\epsilon^{\prime})\subset\mathcal{H}_{k}(\epsilon) and diam(K(ϵ))<diam(K(ϵ))diam(K(\epsilon^{\prime}))<diam(K(\epsilon)). Hence after taking a basepoint yK(ϵ)y\in K(\epsilon^{\prime}) we have

mBMk(T1(k(ϵ)/𝒫k))C(p𝒫k,p1yV(ϵ)(μyk(𝒟))2eδ(Γk)d(p1y,y)(d(y,py)+C))m^{k}_{\rm BM}(T^{1}(\mathcal{H}_{k}(\epsilon^{\prime})/\mathcal{P}_{k}))\leq C\left(\sum_{p\in\mathcal{P}_{k},p^{-1}y\in V(\epsilon^{\prime})}(\mu^{k}_{y}(\mathcal{D}))^{2}e^{-\delta(\Gamma_{k})d(p^{-1}y,y)}(d(y,py)+C)\right)

for a constant C>0C>0 independent of ϵ\epsilon^{\prime} and kk.

The neighborhood V(ϵ)V(\epsilon^{\prime}) is smaller and smaller as ϵ0\epsilon^{\prime}\rightarrow 0, as if ηη+\eta^{-}\eta^{+} intersects k(ϵ)\mathcal{H}_{k}(\epsilon^{\prime}) then it has to intersect k(ϵ)\mathcal{H}_{k}(\epsilon). Then for the pp summands considered for ϵ\epsilon^{\prime} we have

d(y,py)d(x,px)+C,μyk(𝒟)Ceδ(Γk)d(x,y)μxk(𝒟)d(y,py)\leq d(x,px)+C^{\prime},\quad\mu^{k}_{y}(\mathcal{D})\leq C^{\prime}e^{-\delta(\Gamma_{k})d(x,y)}\mu^{k}_{x}(\mathcal{D})

for CC^{\prime} constant independent of ϵ\epsilon^{\prime} and kk. We always have the bound d(y,py)d(x,px)2d(x,y)d(y,py)\geq d(x,px)-2d(x,y) by triangular inequality and the fact that pp is an isometry.

Putting altogether, we have that

(6) mBMk(T1(k(ϵ)/𝒫k))C′′(μxk(𝒟))2(p𝒫k,p1xV(ϵ)eδ(Γk)d(p1x,x)(d(x,px)+C′′))m^{k}_{\rm BM}(T^{1}(\mathcal{H}_{k}(\epsilon^{\prime})/\mathcal{P}_{k}))\leq C^{\prime\prime}(\mu^{k}_{x}(\mathcal{D}))^{2}\left(\sum_{p\in\mathcal{P}_{k},p^{-1}x\in V(\epsilon^{\prime})}e^{-\delta(\Gamma_{k})d(p^{-1}x,x)}(d(x,px)+C^{\prime\prime})\right)

for a constant C′′>0C^{\prime\prime}>0 independent of ϵ\epsilon^{\prime} and kk. Recall that when δ(Γk)\delta(\Gamma_{k}) is strictly bigger than (n1)/2(n-1)/2, by [McM99, Theorem 6.1], the tails of the series p𝒫keδ(Γk)d(p1x,x)\sum_{p\in\mathcal{P}_{k}}e^{-\delta(\Gamma_{k})d(p^{-1}x,x)} are uniformly small. Specifically, for any η>0\eta>0 there exists a neighborhood UnU\subset\mathbb{H}^{n} of ξ\xi so that

p𝒫k,pxUeδ(Γk)d(p1x,x)<η,\sum_{p\in\mathcal{P}_{k},\,px\subset U}e^{-\delta(\Gamma_{k})d(p^{-1}x,x)}<\eta,

for kk sufficiently large. We also have that the tails of the series p𝒫keδ(Γk)d(p1x,x)(d(x,px)+C′′)\sum_{p\in\mathcal{P}_{k}}e^{-\delta(\Gamma_{k})d(p^{-1}x,x)}(d(x,px)+C^{\prime\prime}) are uniformly small, as d(x,px)d(x,px) is uniformly dominated by ecd(p1x,x)e^{cd(p^{-1}x,x)} for any c>0c>0. Hence by taking ϵ\epsilon^{\prime} sufficiently small, the right hand side of (6) corresponds to a smaller tail of the series p𝒫keδ(Γk)d(p1x,x)(d(x,px)+C′′)\sum_{p\in\mathcal{P}_{k}}e^{-\delta(\Gamma_{k})d(p^{-1}x,x)}(d(x,px)+C^{\prime\prime}). Thus, by applying [McM99, Theorem 6.2] for the sequence of exponents δ(Γk)c\delta(\Gamma_{k})-c, the right hand side of will be arbitrarily small for ϵ\epsilon^{\prime} sufficiently small and kk sufficiently large.

Next we use Proposition 4.1 to prove the convergence of the Bowen-Margulis measures. The following proposition is a restatement of Proposition 1.6.

Proposition 4.2.

Suppose that (Mk=n/Γk)k(M_{k}=\mathbb{H}^{n}/\Gamma_{k})_{k\in\mathbb{N}} is a sequence of hyperbolic manifolds which is strongly convergent to a geometrically finite hyperbolic manifold M=n/ΓM=\mathbb{H}^{n}/\Gamma with δ(Γ)>(n1)/2\delta(\Gamma)>(n-1)/2. Let mBMk,mBMm^{k}_{\rm BM},m_{\rm BM} be the Bowen-Margulis measures on T1MkT^{1}M_{k} and T1MT^{1}M, respectively. For r>0r>0 we denote by Mk<rMk,M<rMM_{k}^{<r}\subset M_{k},M^{<r}\subset M the set of points with injectivity radius less than r. Then for any r>0r>0 we have

limkT1Mk<r𝑑mBMkT1M<r𝑑mBM.\lim_{k\rightarrow\infty}\int_{T^{1}M_{k}^{<r}}dm^{k}_{\rm BM}\rightarrow\int_{T^{1}M^{<r}}dm_{\rm BM}.

Moreover, by taking rr sufficiently large we have that

mBMkmBM.\|m^{k}_{\rm BM}\|\rightarrow\|m_{\rm BM}\|.
Proof.

Denote Ma,b=M>aM<bM^{a,b}=M^{>a}\cap M^{<b}. Take U1,UmMU_{1},\ldots U_{m}\subset M balls with compact closure, whose union covers C(M)ϵ,r=C(M)Mϵ,rC(M)^{\epsilon,r}=C(M)\cap M^{\epsilon,r}. Take φ¯1,,φ¯m\bar{\varphi}_{1},\ldots,\bar{\varphi}_{m} partition of unity subordinated to U1,UmU_{1},\ldots U_{m}, in the sense that φ¯=i=1mφ¯i\bar{\varphi}=\sum_{i=1}^{m}\bar{\varphi}_{i} has support contained in Mϵη,r+ηM^{\epsilon-\eta,r+\eta} and is identically equal to 11 in C(M)ϵ,rC(M)^{\epsilon,r}, for some arbitrarily small η>0\eta>0. Let Ui~\tilde{U_{i}} be a lift of UiU_{i} in n\mathbb{H}^{n} such that the union covers a fundamental domain of MM. We denote φi\varphi_{i} a compactly supported function subordinated to U~i\tilde{U}_{i} such that φi=φ¯iProj\varphi_{i}=\bar{\varphi}_{i}\circ\textup{Proj}.

Then since the Patterson-Sullivan measures μx0k\mu^{k}_{x_{0}} converge weakly to μx0\mu_{x_{0}}, the critical exponents δk=δ(Γk)\delta_{k}=\delta(\Gamma_{k}) converge to δ=δ(Γ)\delta=\delta(\Gamma), and we can express the Bowen-Margulis measures as dm~BMk(v)=eδk(βv(π(v),x0)+βv+(π(v),x0))dμx0k(v)dμx0k(v+)dtd\tilde{m}^{k}_{\rm{BM}}(v)=e^{-\delta_{k}(\beta_{v_{-}}(\pi(v),x_{0})+\beta_{v^{+}}(\pi(v),x_{0}))}d\mu^{k}_{x_{0}}(v_{-})d\mu^{k}_{x_{0}}(v_{+})dt, then for kk sufficiently large we have

(7) |T1Ui~φi𝑑m~BMkT1Ui~φi𝑑m~BM|<α,\bigg{|}\int_{T^{1}\tilde{U_{i}}}\varphi_{i}d\tilde{m}^{k}_{\rm BM}-\int_{T^{1}\tilde{U_{i}}}\varphi_{i}d\tilde{m}_{\rm BM}\bigg{|}<\alpha,

for some small α>0\alpha>0.

By Proposition 4.1 we have that T1Mk<ϵ𝑑mBMk,T1M<ϵ𝑑mBM<α\int_{T^{1}M^{<\epsilon}_{k}}dm^{k}_{\rm BM},\int_{T^{1}M^{<\epsilon}}dm_{\rm BM}<\alpha, and by construction we have that

(8) |T1M<r𝑑mBMi=1mT1Ui~φi𝑑m~BM|<T1M<ϵ𝑑mBM+T1Mr,r+η𝑑mBM.\bigg{|}\int_{T^{1}M^{<r}}dm_{\rm BM}-\sum_{i=1}^{m}\int_{T^{1}\tilde{U_{i}}}\varphi_{i}d\tilde{m}_{\rm BM}\bigg{|}<\int_{T^{1}M^{<\epsilon}}dm_{\rm BM}+\int_{T^{1}M^{r,r+\eta}}dm_{\rm BM}.

Now, since MkM_{k} converges strongly to MM, for kk large we have that i=1mT1Ui~φi𝑑m~BMk\sum_{i=1}^{m}\int_{T^{1}\tilde{U_{i}}}\varphi_{i}d\tilde{m}^{k}_{\rm BM} is bounded between (1α)T1Mkϵ+η,rη𝑑mBMk(1-\alpha)\int_{T^{1}M_{k}^{\epsilon+\eta,r-\eta}}dm^{k}_{\rm BM} and (1+α)T1Mkϵη,r+η𝑑mBMk(1+\alpha)\int_{T^{1}M_{k}^{\epsilon-\eta,r+\eta}}dm^{k}_{\rm BM}. Hence

(9) |T1Mk<r𝑑mBMki=1mT1Ui~φi𝑑m~BMk|<T1Mk<ϵ𝑑mBMk+αT1Mkϵ,r𝑑mBMk+T1Mkr,r+η𝑑mBMk.\bigg{|}\int_{T^{1}M_{k}^{<r}}dm^{k}_{\rm BM}-\sum_{i=1}^{m}\int_{T^{1}\tilde{U_{i}}}\varphi_{i}d\tilde{m}^{k}_{\rm BM}\bigg{|}<\int_{T^{1}M_{k}^{<\epsilon}}dm^{k}_{\rm BM}+\alpha\int_{T^{1}M_{k}^{\epsilon,r}}dm^{k}_{\rm BM}+\int_{T^{1}M_{k}^{r,r+\eta}}dm^{k}_{\rm BM}.

By a similar partition of unity argument, we can show that for any 0<a<b0<a<b there exists η0>0\eta_{0}>0 sufficiently small so that for any k1k\gg 1 sufficiently large we have that

(10) T1Mka,b𝑑mBMk<2T1Maη0,b+η0𝑑mBM.\int_{T^{1}M_{k}^{a,b}}dm^{k}_{\rm BM}<2\int_{T^{1}M^{a-\eta_{0},b+\eta_{0}}}dm_{\rm BM}.

Finally, we have to see that the function (a,b)T1Ma,b𝑑mBMk(a,b)\mapsto\int_{T^{1}M^{a,b}}dm^{k}_{\rm BM} is continuous. Because of monotonicity this reduces to prove that for any r>0r>0, T1M<r𝑑mBMk=0\int_{T^{1}\partial M^{<r}}dm^{k}_{\rm BM}=0. Indeed, the lift M~<rn\partial\tilde{M}^{<r}\subseteq\mathbb{H}^{n} is contained the union of tubes around closed geodesics of length r\leq r (considering parabolic cusps corresponding to 0 length geodesics). For core geodesics of length strictly less than rr, these tubes are strictly convex and hence the boundaries intersect any geodesic in a discrete set. If we happen to have a geodesic of length rr, then the intersection of M~<r\partial\tilde{M}^{<r} with any geodesic is a discrete set, unless the geodesic is equal to the geodesic axis. In either case, the set M~<rn\partial\tilde{M}^{<r}\subseteq\mathbb{H}^{n} has zero measure for the Bowen-Margulis measure dm~BM(v)=e2δ(v|v+)x0dμx0(v)dμx0(v+)dtd\tilde{m}_{\rm{BM}}(v)=e^{-2\delta(v_{-}|v_{+})_{x_{0}}}d\mu_{x_{0}}(v_{-})d\mu_{x_{0}}(v_{+})dt, as for almost every geodesic line \ell the intersection M~<r\partial\tilde{M}^{<r}\cap\ell has length 0.

Applying the triangular inequality, replacing equations (7), (8), (9), and then using Proposition 4.1, (10) (for sufficiently large kk and η\eta sufficiently small) we have that

(11) |T1Mk<r𝑑mBMkT1M<r𝑑mBM|\displaystyle\bigg{|}\int_{T^{1}M_{k}^{<r}}dm^{k}_{\rm BM}-\int_{T^{1}M^{<r}}dm_{\rm BM}\bigg{|} <|T1Mk<r𝑑mBMki=1mT1Ui~φi𝑑m~BMk|\displaystyle<\bigg{|}\int_{T^{1}M_{k}^{<r}}dm^{k}_{\rm BM}-\sum_{i=1}^{m}\int_{T^{1}\tilde{U_{i}}}\varphi_{i}d\tilde{m}^{k}_{\rm BM}\bigg{|}
+i=1m|T1Ui~φi𝑑m~BMkT1Ui~φi𝑑m~BM|\displaystyle+\sum_{i=1}^{m}\bigg{|}\int_{T^{1}\tilde{U_{i}}}\varphi_{i}d\tilde{m}^{k}_{\rm BM}-\int_{T^{1}\tilde{U_{i}}}\varphi_{i}d\tilde{m}_{\rm BM}\bigg{|}
+|T1M<r𝑑mBMi=1mT1Ui~φi𝑑m~BM|\displaystyle+\bigg{|}\int_{T^{1}M^{<r}}dm_{\rm BM}-\sum_{i=1}^{m}\int_{T^{1}\tilde{U_{i}}}\varphi_{i}d\tilde{m}_{\rm BM}\bigg{|}
<T1Mk<ϵ𝑑mBMk+αT1Mkϵ,r𝑑mBMk+T1Mkr,r+η𝑑mBMk\displaystyle<\int_{T^{1}M_{k}^{<\epsilon}}dm^{k}_{\rm BM}+\alpha\int_{T^{1}M_{k}^{\epsilon,r}}dm^{k}_{\rm BM}+\int_{T^{1}M_{k}^{r,r+\eta}}dm^{k}_{\rm BM}
+mα+T1M<ϵ𝑑mBM+T1Mr,r+η𝑑mBM\displaystyle+m\alpha+\int_{T^{1}M^{<\epsilon}}dm_{\rm BM}+\int_{T^{1}M^{r,r+\eta}}dm_{\rm BM}
<(m+2)α+2αT1M<r+η𝑑mBM+3T1Mrη,r+2η𝑑mBM\displaystyle<(m+2)\alpha+2\alpha\int_{T^{1}M^{<r+\eta}}dm_{\rm BM}+3\int_{T^{1}M^{r-\eta,r+2\eta}}dm_{\rm BM}

which goes to 0 as k+k\rightarrow+\infty, and α,η0\alpha,\eta\rightarrow 0.

The last part of the section is to prove the convergence of skinning measures.

Proof of Corollary 1.8: Observe that since we have strong convergence for well-positioned convex sets DkDD_{k}\rightarrow D, we can take lifts Dk~,D~n\widetilde{D_{k}},\tilde{D}\subset\mathbb{H}^{n} and compact sets EkDk~,ED~E_{k}\subset\partial\widetilde{D_{k}},E\subset\partial\tilde{D} so that EkE_{k}, EE are fundamental domains for the support of σDk±,σD±\sigma^{\pm}_{\partial D_{k}},\sigma^{\pm}_{\partial D} (respectively) and EkE_{k} converges strongly to EE. We can further assume there exists a set F𝕊n1F\subset\mathbb{S}^{n-1} so that PDk(F)P_{D_{k}}(F), PD(F)P_{D}(F) cover EkE_{k} and EE (respectively) on its interior, see (1) for the definition of the maps PDkP_{D_{k}} and PDP_{D}. Under this assumptions we have

σDk±σ~Dk~±(PDk(F))σ~D~±(PD(F)).\|\sigma^{\pm}_{\partial D_{k}}\|\leq\tilde{\sigma}^{\pm}_{\widetilde{D_{k}}}(P_{D_{k}}(F))\rightarrow\tilde{\sigma}^{\pm}_{\tilde{D}}(P_{D}(F)).

Reducing the set FF so that σ~Dk~±(PDk(F)Ek),σ~D~±(PD(F)E)\tilde{\sigma}^{\pm}_{\widetilde{D_{k}}}(P_{D_{k}}(F)\setminus E_{k}),\tilde{\sigma}^{\pm}_{\tilde{D}}(P_{D}(F)\setminus E) are arbitrarily small, we then have

σDk±=σ~Dk~±(Ek)σ~D~±(E)=σD±,\|\sigma^{\pm}_{\partial D_{k}}\|=\tilde{\sigma}^{\pm}_{\widetilde{D_{k}}}(E_{k})\rightarrow\tilde{\sigma}^{\pm}_{\tilde{D}}(E)=\|\sigma^{\pm}_{\partial D}\|,

which proves the first statement

The relative result for subsets Ωk,Ω\Omega_{k},\Omega is proved by taking the fundamental domains EkDk~E^{\prime}_{k}\subset\partial\widetilde{D_{k}}, ED~E^{\prime}\subset\partial\tilde{D} for the support of σΩk±,σΩ±\sigma^{\pm}_{\partial\Omega_{k}},\sigma^{\pm}_{\partial\Omega} (respectively) and arguing as above.

5. Application: uniform orthogeodesic counting

In this section, we use the results of uniform spectral gap and convergence of the Bowen-Margulis and skinning measures in Section 3 and Section 4 to prove Theorem 1.3. Suppose that D+,DD^{+},D^{-} are well-positioned convex subsets of a hyperbolic manifold M=n/ΓM=\mathbb{H}^{n}/\Gamma. A common perpendicular from DD^{-} to D+D^{+} is a locally geodesic path in MM which starts perpendicularly from DD^{-} and arrives perpendicularly to D+D^{+}. For any t0t\geq 0, let 𝒩D,D+(t)\mathcal{N}_{D^{-},D^{+}}(t) be the cardinality of the set of common perpendiculars from DD^{-} to D+D^{+} with length at most tt.

As before, (Mk=n/Γk)k(M_{k}=\mathbb{H}^{n}/\Gamma_{k})_{k\in\mathbb{N}} is a sequence of hyperbolic manifolds which converges strongly to a geometrically finite manifold M=n/ΓM=\mathbb{H}^{n}/\Gamma, so that we have well-positioned convex subsets Dk±MkD^{\pm}_{k}\subset M_{k} that strongly converge to D±MD^{\pm}\subset M. Before the proof, we need to introduce the following notations.

Given vT1nv\in T^{1}\mathbb{H}^{n}, the strong stable/unstable manifold is defined as

W±(v)={vT1n:d(v(t),v(t))0 as t±},W^{\pm}(v)=\{v^{\prime}\in T^{1}\mathbb{H}^{n}:d(v(t),v^{\prime}(t))\rightarrow 0\text{ as }t\rightarrow\pm\infty\},

which is equipped with Hamenstädt’s distance function dW±(v)d_{W^{\pm}(v)}, see [Ham89, PP17]. Then given any constant r>0r>0, for all vT1nv\in T^{1}\mathbb{H}^{n}, we can define the open ball of radius rr centered at vv in the strong stable/unstable manifold in the following

B±(v,r)={vW±(v):dW±(v)(v,v)<r}.B^{\pm}(v,r)=\{v^{\prime}\in W^{\pm}(v):d_{W^{\pm}(v)}(v,v^{\prime})<r\}.

Given any vT1nv\in T^{1}\mathbb{H}^{n}, and η,η>0\eta,\eta^{\prime}>0, let

Vv,η,η±=s[η,η]gsB±(v,η).V^{\pm}_{v,\eta,\eta^{\prime}}=\bigcup_{s\in\left[-\eta,\eta\right]}g^{s}B^{\pm}(v,\eta^{\prime}).

Given a proper closed convex subset DD of n\mathbb{H}^{n}, for all subsets Ω\Omega^{-} of +1D\partial^{1}_{+}D and Ω+\Omega^{+} of 1D\partial^{1}_{-}D, let

𝒱η,η(Ω±)=vΩ±Vv,η,η.\mathcal{V}_{\eta,\eta^{\prime}}(\Omega^{\pm})=\bigcup_{v\in\Omega^{\pm}}V^{\mp}_{v,\eta,\eta^{\prime}}.

By using the projection map π:T1nn\pi:T^{1}\mathbb{H}^{n}\rightarrow\mathbb{H}^{n}, the strong stable/unstable manifold W±(v)W^{\pm}(v) projects to the stable/unstable horosphere of vv centered at v+v_{+} and vv_{-}, denoted by H±(v)=π(W±(v))H_{\pm}(v)=\pi(W^{\pm}(v)). The corresponding horoball bounded by H±(v)H_{\pm}(v) is denoted by HB±(v)HB_{\pm}(v). Following the notation in [PP17], we let

μW+(v)=σ~HB+(v) and μW(v)=σ~HB(v)+\mu_{W^{+}(v)}=\tilde{\sigma}^{-}_{HB_{+}(v)}\quad\text{ and }\quad\mu_{W^{-}(v)}=\tilde{\sigma}^{+}_{HB_{-}(v)}

denote the skinning measures on the strong stable/unstable manifolds W±(v)W^{\pm}(v).

Definition 5.1.

Given a discrete isometry subgroup Γ<Isom(n)\Gamma<\operatorname{{\mathrm{I}som}}(\mathbb{H}^{n}), we say (n,Γ)(\mathbb{H}^{n},\Gamma) has radius-continuous strong stable/unstable ball masses if, for every ϵ>0\epsilon>0, and r1r\geq 1 close enough to 11,

μW±(v)(B±(v,r))eϵμW±(v)(B±(v,1)),\mu_{W^{\pm}(v)}(B^{\pm}(v,r))\leq e^{\epsilon}\mu_{W^{\pm}(v)}(B^{\pm}(v,1)),

for all vT1nv\in T^{1}\mathbb{H}^{n} where B±(v,1)B^{\pm}(v,1) meets the support of μW±(v)\mu_{W^{\pm}(v)}.

The following proposition proves that the radius-continuous property of the strong stable/unstable ball masses can be taken uniformly along a strongly convergent sequence of geometrically finite hyperbolic manifolds.

Proposition 5.2.

Suppose (Mk=n/Γk)k(M_{k}=\mathbb{H}^{n}/\Gamma_{k})_{k\in\mathbb{N}} is a sequence of hyperbolic manifolds which strongly converges to a geometrically finite hyperbolic manifold M=n/ΓM=\mathbb{H}^{n}/\Gamma with δ(Γ)>(n1)/2\delta(\Gamma)>(n-1)/2. Let (DkMk)k(D_{k}\subseteq M_{k})_{k\in\mathbb{N}} be a sequence of well-positioned convex subsets in MkM_{k} which strongly converges to a well-positioned convex subset DD, with lifts to n\mathbb{H}^{n} denoted by Dk~,D~\widetilde{D_{k}},\widetilde{D}, respectively. Let Ωk±1Dk~\Omega_{k}^{\mp}\subseteq\partial^{1}_{\pm}\widetilde{D_{k}}, Ω±1D~\Omega^{\mp}\subseteq\partial^{1}_{\pm}\widetilde{D} be compact sets so that Ωk\Omega_{k}^{\mp} converges strongly to Ω\Omega^{\mp}. Then there exists sufficiently large R>0R>0 so that for any ϵ\epsilon we have η=η(ϵ,R)>0\eta=\eta(\epsilon,R)>0 satisfying that

μW±(v)k(B±(v,(1+r)R))eϵμW±(v)k(B±(v,R))\mu^{k}_{W^{\pm}(v)}(B^{\pm}(v,(1+r)R))\leq e^{\epsilon}\mu^{k}_{W^{\pm}(v)}(B^{\pm}(v,R))

for any vΩkv\in\Omega_{k}^{\mp}, 0<r<η0<r<\eta.

Proof.

Let’s prove that case for Ω+\Omega^{+}, and the proof for Ω\Omega^{-} is similar. As done in the proof of [Rob03, Proposition 6.2] (using [Rob00, Section 3.1]), the function (v,R)μW±(v)(B±(v,R))(v,R)\mapsto\mu_{W^{\pm}(v)}(B^{\pm}(v,R)) is continuous for vT1n,R>0v\in T^{1}\mathbb{H}^{n},R>0, as well as Γ\Gamma-invariant. Moreover, since Ω+\Omega^{+} is compact, there exists R>0R>0 sufficiently large so that the function vμW(v)(B(v,R))v\mapsto\mu_{W^{-}(v)}(B^{-}(v,R)) is a uniformly continuous positive function in some neighborhood of Ω\Omega^{-}. It suffices then to prove the statement for sufficiently large kk.

Denote by A(v,R,r)=B(v,(1+r)R)B(v,R)W(v)A(v,R,r)=B^{-}(v,(1+r)R)\setminus B^{-}(v,R)\subset W^{-}(v) the annulus in W(v)W^{-}(v) with center vv between radius R,(1+r)RR,(1+r)R. We will show that there exists m>0m>0 and function η(ϵ)>0\eta(\epsilon)>0 so that for kk large, v𝒱η,η(Ω+)v\in\mathcal{V}_{\eta,\eta}(\Omega^{+}) and 0<r<η0<r<\eta the following two statements hold

  1. (1)

    μW(v)k(B(v,R))m\mu^{k}_{W^{-}(v)}(B^{-}(v,R))\geq m,

  2. (2)

    μW(v)k(A(v,R,r))<ϵ\mu^{k}_{W^{-}(v)}(A(v,R,r))<\epsilon.

Then it is clear that the statement follows from (1) and (2) by making ϵ\epsilon arbitrarily small. Now we prove items (1) and (2) respectively.

  1. (1)

    For a vector uT1nu\in T^{1}\mathbb{H}^{n}, we define a function Pu:W(u)nP_{u}:W^{-}(u)\rightarrow\partial_{\infty}\mathbb{H}^{n} where Pu(v)P_{u}(v) is the endpoint of the bi-infinite geodesic uπ(v)u_{-}\pi(v) different from uu_{-} as shown in Figure 5.1. Since 𝒱η,η(Ω+)\mathcal{V}_{\eta,\eta}(\Omega^{+}) has compact closure, we can take finitely many vi𝒱η,η(Ω+)v_{i}\in\mathcal{V}_{\eta,\eta}(\Omega^{+}) so that for any u𝒱η,η(Ω+)u\in\mathcal{V}_{\eta,\eta}(\Omega^{+}), there exists viv_{i} such that

    Pu1Pvi(B(vi,R/2))B(u,R).P_{u}^{-1}P_{v_{i}}(B^{-}(v_{i},R/2))\subseteq B^{-}(u,R).

    Moreover, we can assume that the conformal factor between μW(vi)k\mu^{k}_{W^{-}(v_{i})} and μW(u)k\mu^{k}_{W^{-}(u)} at the sets B(vi,R/2)B^{-}(v_{i},R/2), Pu1Pvi(B(vi,R/2))P_{u}^{-1}P_{v_{i}}(B^{-}(v_{i},R/2)) is between 1/21/2 and 22. This can be done uniformly for all kk by following [Rob03, Subsection 1.H]. By taking η\eta small we can assume that μW(vi)(B(vi,R/2))>2m\mu_{W^{-}(v_{i})}(B^{-}(v_{i},R/2))>2m for some fixed m>0m>0 and for any vi𝒱η,η(Ω+)v_{i}\in\mathcal{V}_{\eta,\eta}(\Omega^{+}). Then by weak-convergence of measures, we have that for any viv_{i} (and large kk) μW(vi)k(B(vi,R/2))>2m\mu^{k}_{W^{-}(v_{i})}(B^{-}(v_{i},R/2))>2m. Then it follows that

    μW(u)k(B(u,R))μW(u)k(Pu1Pvi(B(vi,R/2)))12μW(vi)k(B(vi,R/2))>m.\mu^{k}_{W^{-}(u)}(B^{-}(u,R))\geq\mu^{k}_{W^{-}(u)}(P_{u}^{-1}P_{v_{i}}(B^{-}(v_{i},R/2)))\geq\frac{1}{2}\mu^{k}_{W^{-}(v_{i})}(B^{-}(v_{i},R/2))>m.
  2. (2)

    Since 𝒱η,η(Ω+)\mathcal{V}_{\eta,\eta}(\Omega^{+}) has compact closure and δ(Γ)>(n1)/2\delta(\Gamma)>(n-1)/2, given ϵ>0\epsilon>0 we can take η\eta small enough so that for v𝒱η,η(Ω+)v\in\mathcal{V}_{\eta,\eta}(\Omega^{+}) we have that μW(v)(A(v,R,5η))<ϵ\mu_{W^{-}(v)}(A(v,R,5\eta))<\epsilon. We will take again a finite collection of vectors viv_{i}, although now they need to satisfy the following list of properties.

    • The finite collection of viv_{i} is taken so that B(vi,4η)A(v,R,5η)B^{-}(v_{i},4\eta)\subset A(v,R,5\eta) for some v𝒱η,η(Ω+)v\in\mathcal{V}_{\eta,\eta}(\Omega^{+}). Denote their total number by C2C_{2},

    • For any v𝒱η,η(Ω+)v\in\mathcal{V}_{\eta,\eta}(\Omega^{+}) and any B(u,2η)A(v,R,5η)B^{-}(u,2\eta)\subset A(v,R,5\eta) we have that

      Pu1Pvi(B(vi,4η))B(u,2η)P_{u}^{-1}P_{v_{i}}(B^{-}(v_{i},4\eta))\supseteq B^{-}(u,2\eta)

      with conformal factor bounded between 12\frac{1}{2} and 22.

    Take sufficiently large kk so that μW(vi)k(B(vi,4η))μW(vi)(B(vi,4η))+ζ\mu^{k}_{W^{-}(v_{i})}(B^{-}(v_{i},4\eta))\leq\mu_{W^{-}(v_{i})}(B^{-}(v_{i},4\eta))+\zeta for ζ\zeta small still to be determined.

    Let v𝒱η,η(Ω+)v\in\mathcal{V}_{\eta,\eta}(\Omega^{+}). Cover A(v,R,η)A(v,R,\eta) by finitely many disjoint measurable sets BjB_{j}, so that each BjB_{j} is contained in a ball B(uj,2η)B^{-}(u_{j},2\eta) inside of A(v,R,5η)A(v,R,5\eta). Then by the second bullet point, for each uju_{j} we choose viv_{i} so that

    Puj1Pvi(B(vi,4η))B(uj,2η).P_{u_{j}}^{-1}P_{v_{i}}(B^{-}(v_{i},4\eta))\supseteq B^{-}({u_{j}},2\eta).

    Observe that each viv_{i} can only be repeatedly selected less than C3C_{3} times, for some constant C3C_{3} depending only on the dimension nn. Then we have the following chain of inequalities, which follow from the covering {Bj}\{B_{j}\} of A(v,R,r)A(v,R,r), the inclusion Puj1Pvi(B(vi,4η))B(uj,2η)BjP_{u_{j}}^{-1}P_{v_{i}}(B^{-}(v_{i},4\eta))\supseteq B^{-}({u_{j}},2\eta)\supseteq B_{j}, the bound on the conformal factor of Pu1PviP_{u}^{-1}P_{v_{i}}, the convergence μW(v)kμW(v)\mu^{k}_{W^{-}(v)}\rightarrow\mu_{W^{-}(v)}, the inclusion B(vi,4η)A(v,R,5η)B^{-}(v_{i},4\eta)\subset A(v,R,5\eta), and the bound on the cardinality of the finite set of viv_{i}’s.

    (12) μW(v)k(A(v,R,r))\displaystyle\mu^{k}_{W^{-}(v)}(A(v,R,r)) jμW(v)k(Bj)C3iμW(v)k(Puj1Pvi(B(vi,4η)))\displaystyle\leq\sum_{j}\mu^{k}_{W^{-}(v)}(B_{j})\leq C_{3}\sum_{i}\mu^{k}_{W^{-}(v)}(P_{u_{j}}^{-1}P_{v_{i}}(B^{-}(v_{i},4\eta)))
    2C3iμW(v)k(B(vi,4η))2C3i(μW(v)(B(vi,4η))+ζ)\displaystyle\leq 2C_{3}\sum_{i}\mu^{k}_{W^{-}(v)}(B^{-}(v_{i},4\eta))\leq 2C_{3}\sum_{i}\left(\mu_{W^{-}(v)}(B^{-}(v_{i},4\eta))+\zeta\right)
    4C3i(μW(v)(A(v,R,5η))+ζ)4C2C3(ϵ+ζ)\displaystyle\leq 4C_{3}\sum_{i}\left(\mu_{W^{-}(v)}(A(v,R,5\eta))+\zeta\right)\leq 4C_{2}C_{3}(\epsilon+\zeta)

    which is arbitrarily small for η\eta small and kk large.

Refer to caption
Figure 5.1.

Now we state and sketch the general uniform orthogeodesic counting for convergent sequences of convex sets in strongly convergent hyperbolic nn-manifolds. For a thorough presentation, we refer the reader to Theorem 5.5 in the Appendix.

Theorem 5.3.

Suppose that (Mk=n/Γk)k(M_{k}=\mathbb{H}^{n}/\Gamma_{k})_{k\in\mathbb{N}} is a sequence of hyperbolic manifolds which strongly converges to a geometrically finite hyperbolic manifold M=n/ΓM=\mathbb{H}^{n}/\Gamma with δ(Γ)>(n1)/2\delta(\Gamma)>(n-1)/2. Let (Dk±)k(D^{\pm}_{k})_{k\in\mathbb{N}} be a sequence of well-positioned convex subsets in MkM_{k} which converges strongly to a well-positioned convex subset D±D^{\pm} in MM, respectively. Then we can count 𝒩Dk,Dk+(t)\mathcal{N}_{D^{-}_{k},D^{+}_{k}}(t) uniformly, in the sense that

𝒩Dk,Dk+(t)σDk+σDk+δ(Γk)mBMkeδ(Γk)t\mathcal{N}_{D^{-}_{k},D^{+}_{k}}(t)\approx\dfrac{||\sigma^{+}_{D^{-}_{k}}||\cdot||\sigma^{-}_{D^{+}_{k}}||}{\delta(\Gamma_{k})||m^{k}_{\rm{BM}}||}e^{\delta(\Gamma_{k})t}

up to a multiplicative error uniformly close to 1 along the sequence as tt gets larger, and with σDk±,mBMk,δ(Γk)||\sigma^{\mp}_{D^{\pm}_{k}}||,||m^{k}_{\rm{BM}}||,\delta(\Gamma_{k}) converging to σD±,mBM,δ(Γ)||\sigma^{\pm}_{D^{\mp}}||,||m_{\rm{BM}}||,\delta(\Gamma), respectively. In particular, for n=3n=3, there exist constants A>0,0<b<2A>0,0<b<2 so that

𝒩Dk,Dk+(t)Aebt\mathcal{N}_{D^{-}_{k},D^{+}_{k}}(t)\leq Ae^{bt}
Proof.

There is an explicit counting formula of 𝒩D,D+(t)\mathcal{N}_{D^{-},D^{+}}(t) for orthogeodesic arcs between two convex sets D±D^{\pm} given in [PP17, Theorem 3]:

𝒩D,D+(t)=σD+σD+δmBMeδ(Γ)t(1+O(eκt)).\mathcal{N}_{D^{-},D^{+}}(t)=\dfrac{||\sigma^{+}_{D^{-}}||\cdot||\sigma^{-}_{D^{+}}||}{\delta||m_{\rm{BM}}||}e^{\delta(\Gamma)t}(1+O(e^{-\kappa t})).

This formula holds under the assumption that (n,Γ)(\mathbb{H}^{n},\Gamma) has radius-continuous strong stable/unstable masses. The constant O()O(\cdot) and the parameter κ\kappa depends on Γ\Gamma, the convex sets D±D^{\pm}, the speed of mixing, and the property of radius-continuous strong stable/unstable masses.

By Proposition 4.2 and Corollary 1.8, the Bowen-Margulis measure and the skinning measures converge to the ones of the limit manifold MM weakly. The critical exponent δ(Γk)\delta(\Gamma_{k}) converges to δ(Γ)\delta(\Gamma) [McM99, Theorem 1.5]. The convergence of the speed of mixing is controlled by the spectral gap [EO21]. Hence this quantity also converges to the one of the limit manifold by Theorem 1.1. Therefore, it suffices to prove the sequence Γk\Gamma_{k} and the limit Γ\Gamma have uniform radius-continuous strong stable/unstable ball masses property, which follows from Proposition 5.2.

Remark 5.4.

Careful readers might notice that [PP17, Theorem 3] has the assumption that the manifold has radius-Hölder-continuous strong stable/unstable ball masses, which is not satisfied by the strongly convergent sequence of hyperbolic manifolds MkM_{k} and the limit manifold MM. However, this assumption can be replaced by the property of radius-continuous strong stable/unstable masses [PP17, Lemma 11], and the uniform radius-continuity suffices to control the error term in our setting. We write down the details about the replacement in the Appendix for readers’ convenience, and most of the arguments follow from [PP17].

Proof of Theorem 1.3: By Example 2.12, connected components Dk±D_{k}^{\pm} in the thin part of MkM_{k} are well-positioned convex sets that are strongly convergent to the well-positioned convex sets D±D^{\pm} (respectively). By Theorem 5.3, there is a uniform counting formula for orthogeodesics between DkD^{-}_{k} to Dk+D^{+}_{k} along the sequence. This proves item (1). Similarly, for small r>0r>0, the radius rr embedded balls centered at xkx_{k} are well-positioned convex subsets which are strongly convergent to the embedded rr-ball centered at xx. In that case, let Dk+=DkD^{+}_{k}=D^{-}_{k} be the radius rr ball at xkx_{k}, and D+=DD^{+}=D^{-} be the radius rr ball at xx. Observe that if we change the radius r>0r>0 to a radius s>0,s<rs>0,s<r we have a one-to-one correspondence between the set of orthogeodesics by extending/shortening the geodesic arcs. Such correspondence takes an orthogeodesic of length \ell to its extension of length +2(rs)\ell+2(r-s). Hence applying Theorem 5.3 again and making ss arbitrarily small (or equivalently, translating by 2r2r the counting function for the balls of radius rr), we obtain the uniform counting for geodesic loops based at xkx_{k} along the sequence.

Proof of Corollary 1.5: As explained for instance by Roblin in [Rob03, Chapter 5], one can deduce an asymptotic counting of closed primitive geodesics in manifolds with negative pinched curvature from the asymptotic counting of orbit distance (i.e., geodesic loops), which only depends on the geometry of the universal cover. Namely, if 𝒢M()\mathcal{G}_{M}(\ell) is the set of closed primitive geodesics in MM of length less than >0\ell>0, then [Rob03, Corollary 5.3]

#𝒢M()eδδ as +.\#\mathcal{G}_{M}(\ell)\approx\frac{e^{\delta\ell}}{\delta\ell}\text{ as }\ell\rightarrow+\infty.

Combining with the uniform counting of geodesic loops (Theorem 1.3), we obtain the uniform counting of closed primitive geodesics along a strongly convergent sequence of hyperbolic manifolds.

Appendix

Let’s start with notations needed in the Appendix. Recall that PD:n(nD)DP_{D}:\mathbb{H}^{n}\cup(\partial_{\infty}\mathbb{H}^{n}\setminus\partial_{\infty}D)\rightarrow D is the closest point map defined in Section 2.4 for any nonempty proper closed convex subset DD in n\mathbb{H}^{n}. Let PD+P^{+}_{D} denote the inverse of the restriction to +1D\partial^{1}_{+}D of the positive endpoint map vv+v\mapsto v_{+}, which is a homeomorphism from nD\partial_{\infty}\mathbb{H}^{n}\setminus\partial_{\infty}D to +1D\partial^{1}_{+}D. It is a natural lift of PDP_{D} such that πPD+=PD\pi\circ P^{+}_{D}=P_{D} on nD\partial_{\infty}\mathbb{H}^{n}\setminus\partial_{\infty}D where π:T1nn\pi:T^{1}\mathbb{H}^{n}\rightarrow\mathbb{H}^{n}. Similarly, one can define PD=ιPD+P^{-}_{D}=\iota\circ P^{+}_{D}, where ι:T1nT1n\iota:T^{1}\mathbb{H}^{n}\rightarrow T^{1}\mathbb{H}^{n} is the antipodal flip map given by ιv=v\iota v=-v.

Define

𝒰D±={vT1n:v±D}.\mathcal{U}_{D}^{\pm}=\{v\in T^{1}\mathbb{H}^{n}:v_{\pm}\notin\partial_{\infty}D\}.

This is an open set in T1nT^{1}\mathbb{H}^{n} which is invariant under the geodesic flow and satisfies the 𝒰γD±=γ𝒰D±\mathcal{U}^{\pm}_{\gamma D}=\gamma\mathcal{U}^{\pm}_{D} for any γIsom(n)\gamma\in\operatorname{{\mathrm{I}som}}(\mathbb{H}^{n}). Define a fibration fD+:𝒰D++1Df_{D}^{+}:\mathcal{U}^{+}_{D}\rightarrow\partial^{1}_{+}D as the composition of the positive endpoint map and PD+P^{+}_{D}. Given w+1Dw\in\partial^{1}_{+}D, the fiber of ww for fD+f^{+}_{D} is the set

W0+(w)={vT1n:v+=w+}.W^{0+}(w)=\{v\in T^{1}\mathbb{H}^{n}:v_{+}=w_{+}\}.

Similarly, one can define a fibration fD=ιfD+ι:𝒰D1Df^{-}_{D}=\iota\circ f^{+}_{D}\circ\iota:\mathcal{U}^{-}_{D}\rightarrow\partial^{1}_{-}D and the fiber W0(w)={vT1n:v=w}W^{0-}(w)=\{v\in T^{1}\mathbb{H}^{n}:v_{-}=w_{-}\}.

Suppose that D±D^{\pm} are two well-positioned convex subsets in M=n/ΓM=\mathbb{H}^{n}/\Gamma, and ψ±C0(T1M)\psi^{\pm}\in C^{\infty}_{0}(T^{1}M) are compactly supported functions. Let

𝒩ψ,ψ+(t)=λ,0<λtψ(vλ)ψ+(vλ+)\mathcal{N}_{\psi^{-},\psi^{+}}(t)=\sum_{\lambda,0<\ell_{\lambda}\leq t}\psi^{-}(v^{-}_{\lambda})\psi^{+}(v^{+}_{\lambda})

where the sum is taken over all common perpendiculars λ\lambda between DD^{-} and D+D^{+} whose initial vector vλv^{-}_{\lambda} belongs to +1D\partial^{1}_{+}D^{-} and the terminal vector vλ+v^{+}_{\lambda} belongs to 1D+\partial^{1}_{-}D^{+}, and the length λt\ell_{\lambda}\leq t.

In order to count orthogeodesics between DD^{-} and D+D^{+}, we can parametrize the set of orthogeodesics by a quotient of Γ\Gamma up to a choice of basepoint. Denote by D±~\widetilde{D^{\pm}} the lifts of D±D^{\pm} in n\mathbb{H}^{n}, and distinguish two components D0±D±~D^{\pm}_{0}\subset\widetilde{D^{\pm}}. Then for each γΓ\gamma\in\Gamma we can consider the projection to MM of the unique orthogeodesic between D0D^{-}_{0} and γD0+\gamma D^{+}_{0} such that the closures of D0D^{-}_{0} and γD0+\gamma D^{+}_{0} in nn\mathbb{H}^{n}\cup\partial_{\infty}\mathbb{H}^{n} have empty intersection. It is a simple exercise to see that γ1,γ2Γ\gamma_{1},\gamma_{2}\in\Gamma map to the same orthogeodesic if and only if there exists g±Stab(D0±)g^{\pm}\in Stab(D^{\pm}_{0}) so that γ1=gγ2g+\gamma_{1}=g^{-}\gamma_{2}g^{+}. Hence we can parametrize orthogeodesic by taking the quotient Γ/:=Γ/{γ1=gγ2g+,g±Stab(D0±)}\Gamma/\sim:=\Gamma/\{\gamma_{1}=g^{-}\gamma_{2}g^{+},g^{\pm}\in Stab(D^{\pm}_{0})\}. Although this labeling depends on the choice of D0±D^{\pm}_{0}, we will always work once this decision has been made. We use vγ±1Dv^{\mp}_{\gamma}\in\partial^{1}_{\pm}D^{\mp} to denote the unit tangent vector of γ\gamma at the start/end.

Theorem 5.5.

Suppose that (Mk=n/Γk)k(M_{k}=\mathbb{H}^{n}/\Gamma_{k})_{k\in\mathbb{N}} is a sequence of hyperbolic manifolds which strongly converges to a geometrically finite hyperbolic manifold M=n/ΓM=\mathbb{H}^{n}/\Gamma with δ(Γ)>(n1)/2\delta(\Gamma)>(n-1)/2. Let (Dk±)k(D^{\pm}_{k})_{k\in\mathbb{N}} be a sequence of well-positioned convex subsets in MkM_{k} which strongly converges to D±D^{\pm} in MM, respectively. Let as well (ψk±C0(T1Mk))k(\psi^{\pm}_{k}\in C^{\infty}_{0}(T^{1}M_{k}))_{k\in\mathbb{N}}, ψ±C0(T1M)\psi^{\pm}\in C^{\infty}_{0}(T^{1}M) be compactly supported functions so that ψk±\psi^{\pm}_{k} converges strongly to ψ±\psi^{\pm}, respectively. Then for any ϵ>0\epsilon>0 there exists t0=t0(ϵ),k0=k0(ϵ)>0t_{0}=t_{0}(\epsilon),k_{0}=k_{0}(\epsilon)>0 so that for any t>t0t>t_{0}, k>k0k>k_{0} we have that

(13) σDk+(ψk)σDk+(ψk+)δ(Γk)mBMkϵNψk,ψk+(t)eδ(Γk)tσDk+(ψk)σDk+(ψk+)δ(Γk)mBMk+ϵ.\frac{\sigma^{+}_{D^{-}_{k}}(\psi^{-}_{k})\cdot\sigma^{-}_{D^{+}_{k}}(\psi^{+}_{k})}{\delta(\Gamma_{k})||m^{k}_{\rm{BM}}||}-\epsilon\leq\frac{N_{\psi^{-}_{k},\psi^{+}_{k}}(t)}{e^{\delta(\Gamma_{k})t}}\leq\frac{\sigma^{+}_{D^{-}_{k}}(\psi^{-}_{k})\cdot\sigma^{-}_{D^{+}_{k}}(\psi^{+}_{k})}{\delta(\Gamma_{k})||m^{k}_{\rm{BM}}||}+\epsilon.

Here σDk+(ψk)=+1Dkψk𝑑σk+\sigma^{+}_{D^{-}_{k}}(\psi^{-}_{k})=\int_{\partial^{1}_{+}D_{k}^{-}}\psi^{-}_{k}d\sigma^{+}_{k}, and σDk+(ψk+)\sigma^{-}_{D^{+}_{k}}(\psi^{+}_{k}) is similarly defined.

Proof.

Since both terms in (13) are bilinear in ψk±\psi^{\pm}_{k}, we can assume without lose of generality that, by using a partition of unity, the support of ψk±\psi^{\pm}_{k} is contained in a small relatively compact open set Uk±U^{\pm}_{k} in T1MkT^{1}M_{k}, and there is a small relatively compact open set Uk±~\widetilde{U^{\pm}_{k}} in T1nT^{1}\mathbb{H}^{n} such that the restriction of the quotient map qk:T1nT1Mkq_{k}:T^{1}\mathbb{H}^{n}\rightarrow T^{1}M_{k} to Uk±~\widetilde{U^{\pm}_{k}} is a diffeomorphism to Uk±U^{\pm}_{k}. Define ψk±~C0(T1)\widetilde{\psi^{\pm}_{k}}\in C^{\infty}_{0}(T^{1}\mathbb{H}) with support in Uk±~\widetilde{U^{\pm}_{k}} and coinciding with ψ±qk\psi^{\pm}\circ q_{k} on Uk±~\widetilde{U^{\pm}_{k}}. Similarly, we can define a compactly supported function ψ~C0(T1)\tilde{\psi}\in C^{\infty}_{0}(T^{1}\mathbb{H}) corresponding to ψ\psi. Observe that we can choose the lifts Uk±~\widetilde{U^{\pm}_{k}} and ψk±~\widetilde{\psi^{\pm}_{k}} appropriately such that ψk±~\widetilde{\psi^{\pm}_{k}} converges strongly to ψ±~\widetilde{\psi^{\pm}} and

±1Dk~ψk~𝑑σDk~±=±1Dkψk𝑑σDk±,\int_{\partial^{1}_{\pm}\widetilde{D^{\mp}_{k}}}\widetilde{\psi^{\mp}_{k}}d\sigma^{\pm}_{\widetilde{D^{\mp}_{k}}}=\int_{\partial^{1}_{\pm}D^{\mp}_{k}}\psi^{\mp}_{k}d\sigma^{\pm}_{\partial D^{\mp}_{k}},

where Dk±~\widetilde{D^{\pm}_{k}} are lifts of Dk±D^{\pm}_{k}. From now on, we will distinguish components of Dk±~\widetilde{D^{\pm}_{k}}. By abuse of notation we still denote by Dk±D^{\pm}_{k} a connected component of Dk±~\widetilde{D^{\pm}_{k}}, which we assume is the only connected component of Dk±~\widetilde{D^{\pm}_{k}} so that the intersection of 1Dk±\partial^{1}_{\mp}D^{\pm}_{k} with Uk±~\widetilde{U^{\pm}_{k}} is non-empty by using partition of unity. Observe that then we can label other components by Γk\Gamma_{k} left action γγDk±\gamma\mapsto\gamma D^{\pm}_{k}. These labels are redundant (i.e. label the same set) if and only if γ11γ2\gamma_{1}^{-1}\gamma_{2} belongs to the stabilizer of Dk±D^{\pm}_{k}.

Take η,R>0\eta,R>0 and kk sufficiently large so that the statement of Proposition 5.2 applies for the sequence of convergent precompact subsets (Ωk±:=1Dk±~supp(ψk±~))k(\Omega^{\pm}_{k}:=\partial^{1}_{\mp}\widetilde{D^{\pm}_{k}}\cap supp(\widetilde{\psi^{\pm}_{k}}))_{k\in\mathbb{N}}. We will fix R>0R>0 from now on, but will keep taking smaller (independent of kk) η\eta. Observe that for small, fixed τ>0\tau>0 we have the inclusion

𝒱ηeτ,Reτ(Ωk±)𝒱η,R(Ωk±)\mathcal{V}_{\eta e^{-\tau},Re^{-\tau}}(\Omega^{\pm}_{k})\subset\mathcal{V}_{\eta,R}(\Omega^{\pm}_{k})

is precompact in each slice Vw,η,R±V^{\pm}_{w,\eta,R}, and converges as a whole to 𝒱ηeτ,Reτ(Ω±)𝒱η,R(Ω±)\mathcal{V}_{\eta e^{-\tau},Re^{-\tau}}(\Omega^{\pm})\subset\mathcal{V}_{\eta,R}(\Omega^{\pm}) in the usual sense. If by 𝟙A\mathbbm{1}_{A} we denote the characteristic function of a set AA, then we can construct smooth functions χk±C(T1n)\chi^{\pm}_{k}\in C^{\infty}(T^{1}\mathbb{H}^{n}) so that the following items hold

  1. (1)

    For wΩk±,vW0(w)w\in\Omega^{\pm}_{k},v\in W^{0\mp}(w)

    𝟙𝒱ηeτ,Reτ(Ωk±)(v)χk±(v)𝟙𝒱η,R(Ωk±)(v).\mathbbm{1}_{\mathcal{V}_{\eta e^{-\tau},Re^{-\tau}}(\Omega^{\pm}_{k})}(v)\leq\chi^{\pm}_{k}(v)\leq\mathbbm{1}_{\mathcal{V}_{\eta,R}(\Omega^{\pm}_{k})}(v).
  2. (2)

    The Sobolev norms χk±β\|\chi^{\pm}_{k}\|_{\beta} are uniformly bounded (i.e. independent of kk), where β\beta is the Sobolev norm appearing in the statement of [EO21, Theorem 1.1].

  3. (3)

    For any wΩk±w\in\Omega^{\pm}_{k} we have that

    eϵνw±(Vw,η,R)Vw,η,Rχk±𝑑νw±νw±(Vw,η,R)e^{-\epsilon}\nu^{\pm}_{w}(V^{\mp}_{w,\eta,R})\leq\int_{V^{\mp}_{w,\eta,R}}\chi^{\pm}_{k}d\nu^{\pm}_{w}\leq\nu^{\pm}_{w}(V^{\mp}_{w,\eta,R})

    for ϵ>0\epsilon>0 independent of kk, where dνw±:=dsdμW(w)d\nu^{\pm}_{w}:=dsd\mu_{W^{\mp}(w)}.

In order to define the test functions to apply exponential mixing, we start with the functions Hk±:1Dk±~H^{\pm}_{k}:\partial^{1}_{\mp}\widetilde{D^{\pm}_{k}}\rightarrow\mathbb{R} defined by

Hk±(w)=1Vw,η,Rχk±𝑑νw±.H^{\pm}_{k}(w)=\frac{1}{\int_{V^{\mp}_{w,\eta,R}}\chi^{\pm}_{k}d\nu^{\pm}_{w}}.

Let Φk±:T1n\Phi^{\pm}_{k}:T^{1}\mathbb{H}^{n}\rightarrow\mathbb{R} defined by

Φk±=(Hk±ψk±~)fDk±χk±.\Phi^{\pm}_{k}=(H^{\pm}_{k}\widetilde{\psi^{\pm}_{k}})\circ f^{\mp}_{D^{\pm}_{k}}\chi^{\pm}_{k}.

By construction, we have that Φk±β\|\Phi^{\pm}_{k}\|_{\beta} are uniformly bounded and have support in 𝒱η,R(Ωk±)\mathcal{V}_{\eta,R}(\Omega^{\pm}_{k}). Moreover, Φk±\Phi^{\pm}_{k} are non-negative, measurable functions satisfying

(14) T1nΦk±𝑑m~BMk=1Dk±ψk±𝑑σk\int_{T^{1}\mathbb{H}^{n}}\Phi^{\pm}_{k}d\tilde{m}^{k}_{\rm{BM}}=\int_{\partial^{1}_{\mp}D^{\pm}_{k}}\psi^{\pm}_{k}d\sigma_{k}^{\mp}

Following [PP17], we will estimate in two ways the quantity

(15) Ik(T):=0Teδ(Γk)tγΓkT1n(Φkgt/2)(Φk+gt/2γ1)𝑑m~BMk𝑑tI_{k}(T):=\int^{T}_{0}e^{\delta(\Gamma_{k})t}\sum_{\gamma\in\Gamma_{k}}\int_{T^{1}\mathbb{H}^{n}}(\Phi^{-}_{k}\circ g^{-t/2})(\Phi^{+}_{k}\circ g^{t/2}\circ\gamma^{-1})d\tilde{m}^{k}_{\rm{BM}}dt

By [EO21, Theorem 1.1] and Theorem 1.1 there exist uniform κ>0,O(.)\kappa>0,O(.) such that

(16) Ik(T)=0Teδ(Γk)t(1mBMkT1nΦk𝑑m~BMkT1nΦk+𝑑m~BMk+O(eκtΦkβΦk+β))𝑑t=eδ(Γk)Tδ(Γk)mBMk1Dkψ𝑑σk+1Dk+ψ+𝑑σk++0Teδ(Γk)tO(eκtΦkβΦk+β)𝑑t=eδ(Γk)T(σDk+(ψk)σDk+(ψk+)δ(Γk)mBMk+eδ(Γk)T0Teδ(Γk)tO(eκtΦkβΦk+β)𝑑t)\begin{split}I_{k}(T)&=\int^{T}_{0}e^{\delta(\Gamma_{k})t}\left(\frac{1}{\|m^{k}_{\rm{BM}}\|}\int_{T^{1}\mathbb{H}^{n}}\Phi^{-}_{k}d\tilde{m}^{k}_{\rm{BM}}\int_{T^{1}\mathbb{H}^{n}}\Phi^{+}_{k}d\tilde{m}^{k}_{\rm{BM}}+O(e^{-\kappa t}\|\Phi^{-}_{k}\|_{\beta}\|\Phi^{+}_{k}\|_{\beta})\right)dt\\ &=\frac{e^{\delta(\Gamma_{k})T}}{\delta(\Gamma_{k})\|m^{k}_{\rm{BM}}\|}\int_{\partial^{1}_{-}D^{-}_{k}}\psi^{-}d\sigma_{k}^{-}\int_{\partial^{1}_{+}D^{+}_{k}}\psi^{+}d\sigma_{k}^{+}+\int^{T}_{0}e^{\delta(\Gamma_{k})t}O(e^{-\kappa t}\|\Phi^{-}_{k}\|_{\beta}\|\Phi^{+}_{k}\|_{\beta})dt\\ &=e^{\delta(\Gamma_{k})T}\left(\frac{\sigma^{+}_{D^{-}_{k}}(\psi^{-}_{k})\cdot\sigma^{-}_{D^{+}_{k}}(\psi^{+}_{k})}{\delta(\Gamma_{k})\|m^{k}_{\rm{BM}}\|}+e^{-\delta(\Gamma_{k})T}\int^{T}_{0}e^{\delta(\Gamma_{k})t}O(e^{-\kappa t}\|\Phi^{-}_{k}\|_{\beta}\|\Phi^{+}_{k}\|_{\beta})dt\right)\end{split}

where we used (14) for the second equality. Observe in the final line that we can make the error term eδ(Γk)T0Teδ(Γk)tO(eκtΦkβΦk+β)𝑑te^{-\delta(\Gamma_{k})T}\int^{T}_{0}e^{\delta(\Gamma_{k})t}O(e^{-\kappa t}\|\Phi^{-}_{k}\|_{\beta}\|\Phi^{+}_{k}\|_{\beta})dt arbitrarily small for any T>T0T>T_{0}, where T0T_{0} sufficiently large and independent of kk.

Now we use a second way to compute this integral Ik(T)I_{k}(T). Let δk=δ(Γk)\delta_{k}=\delta(\Gamma_{k}). We interchange the integral over tt and the summation over γ\gamma. Then

Ik(T)=γΓk0TeδktT1n(Φkgt/2)(Φk+gt/2γ1)𝑑m~BMk𝑑t.I_{k}(T)=\sum_{\gamma\in\Gamma_{k}}\int_{0}^{T}e^{\delta_{k}t}\int_{T^{1}\mathbb{H}^{n}}(\Phi^{-}_{k}\circ g^{-t/2})(\Phi^{+}_{k}\circ g^{t/2}\circ\gamma^{-1})d\tilde{m}^{k}_{\rm{BM}}dt.

Suppose that if vT1nv\in T^{1}\mathbb{H}^{n} belongs to the support of (Φkgt/2)(Φk+gt/2γ1)(\Phi^{-}_{k}\circ g^{-t/2})(\Phi^{+}_{k}\circ g^{t/2}\circ\gamma^{-1}), then

vgt/2𝒱η,R(+1Dk)gt/2𝒱η,R(γ1Dk+).v\in g^{t/2}\mathcal{V}_{\eta,R}(\partial^{1}_{+}D_{k}^{-})\cap g^{-t/2}\mathcal{V}_{\eta,R}(\gamma\partial^{1}_{-}D_{k}^{+}).

Then by [PP17, Lemma 7], which is proved by using hyperbolic geometry in n\mathbb{H}^{n}, we have the following

(17) d(wk±,vγ±)=O(η+eγ/2)d(w_{k}^{\pm},v_{\gamma}^{\pm})=O(\eta+e^{-\ell_{\gamma}/2})

where wk=fDk+(v),wk+=fγDk+(v)w_{k}^{-}=f^{+}_{D_{k}}(v),w_{k}^{+}=f^{-}_{\gamma D_{k}^{+}}(v), vγ±v_{\gamma}^{\pm} are endpoints of the common perpendicular between DkD_{k}^{-} and γDk+\gamma D_{k}^{+}, and γ\ell_{\gamma} is the length of the common perpendicular. Since the Lipschitz norm of ψk±~\widetilde{\psi^{\pm}_{k}} are uniformly bounded, and in particular bounded by the β\beta Sobolev norm of ψk±\psi^{\pm}_{k}, we have

|ψk±~(wk±)ψk±~(vγ±)|=O((η+eγ/2)ψk±β).|\widetilde{\psi^{\pm}_{k}}(w^{\pm}_{k})-\widetilde{\psi^{\pm}_{k}}(v^{\pm}_{\gamma})|=O((\eta+e^{-\ell_{\gamma}/2})||\psi_{k}^{\pm}||_{\beta}).

If we define Φ^k±=Hk±fDk±χk±\hat{\Phi}^{\pm}_{k}=H^{\pm}_{k}\circ f^{\mp}_{D^{\pm}_{k}}\chi^{\pm}_{k} so that Φk±=(ψk±~fDk±)Φ^k±\Phi^{\pm}_{k}=(\widetilde{\psi^{\pm}_{k}}\circ f^{\mp}_{D_{k}^{\pm}})\hat{\Phi}^{\pm}_{k}, by applying the previous equation we obtain

(18) Ik(T)=γΓk[ψk±(vγ)ψk±(vγ+)\displaystyle I_{k}(T)=\sum_{\gamma\in\Gamma_{k}}[\psi_{k}^{\pm}(v_{\gamma}^{-})\psi_{k}^{\pm}(v^{+}_{\gamma}) +O((η+eγ/2)||ψk||β||ψk+||β)]×\displaystyle+O((\eta+e^{-\ell_{\gamma}/2})||\psi^{-}_{k}||_{\beta}||\psi^{+}_{k}||_{\beta})]\times
0TeδktvT1nΦ^k(gt/2v)Φ^k+(γ1gt/2v)𝑑m~BMk(v)𝑑t,\displaystyle\int_{0}^{T}e^{\delta_{k}t}\int_{v\in T^{1}\mathbb{H}^{n}}\hat{\Phi}^{-}_{k}(g^{-t/2}v)\hat{\Phi}^{+}_{k}(\gamma^{-1}g^{t/2}v)d\tilde{m}_{\rm{BM}}^{k}(v)dt,

for O(.)O(.) independent of kk.

We now related another test function to Φ^k±\hat{\Phi}^{\pm}_{k} following [PP17, Lemma 8]. Let hk±:T1n[0,]h^{\pm}_{k}:T^{1}\mathbb{H}^{n}\rightarrow\left[0,\infty\right] be the Γk\Gamma_{k}-invariant measurable map defined by

(19) hk(w)=12ημWw±(B±(w,R))h^{\mp}_{k}(w)=\dfrac{1}{2\eta\mu_{W^{\pm}_{w}}(B^{\pm}(w,R))}

if μW±(w)(B±(w,R))>0\mu_{W^{\pm}(w)}(B^{\pm}(w,R))>0, and hk±(w)=0h^{\pm}_{k}(w)=0 otherwise. We define the test function ϕk=ϕη,R,Ωk±:T1n[0,]\phi^{\mp}_{k}=\phi^{\mp}_{\eta,R,\Omega^{\pm}_{k}}:T^{1}\mathbb{H}^{n}\rightarrow\left[0,\infty\right] by

ϕk=hkfDk±𝟙𝒱η,R(Ωk).\phi^{\mp}_{k}=h^{\mp}_{k}\circ f^{\pm}_{D^{\mp}_{k}}\mathbbm{1}_{\mathcal{V}_{\eta,R}(\Omega^{\mp}_{k})}.

By the properties of χk±\chi_{k}^{\pm}, we have

ϕηeτ,Reτ,1Ωk±±eϵΦ^k±ϕk±.\phi^{\pm}_{\eta e^{-\tau},Re^{-\tau},\partial^{1}_{\mp}\Omega^{\pm}_{k}}e^{-\epsilon}\leq\hat{\Phi}^{\pm}_{k}\leq\phi^{\pm}_{k}.

Hence, it suffices to consider the integral

(20) ik(T)=γΓk[ψk±(vγ)ψk±(vγ+)+O((η+eγ/2)||ψk||β||ψk+||β)]×0TeδktT1n(ϕkgt/2)(ϕk+gt/2γ1)𝑑m~BMk𝑑t.\begin{split}i_{k}(T)=\sum_{\gamma\in\Gamma_{k}}[\psi_{k}^{\pm}(v_{\gamma}^{-})\psi_{k}^{\pm}(v^{+}_{\gamma})&+O((\eta+e^{-\ell_{\gamma}/2})||\psi^{-}_{k}||_{\beta}||\psi^{+}_{k}||_{\beta})]\times\\ &\int^{T}_{0}e^{\delta_{k}t}\int_{T^{1}\mathbb{H}^{n}}(\phi^{-}_{k}\circ g^{-t/2})(\phi^{+}_{k}\circ g^{t/2}\circ\gamma^{-1})d\tilde{m}^{k}_{\rm{BM}}dt.\end{split}

By the definition of ϕk±\phi_{k}^{\pm}, the right hand side of (20) is equal to

(21) γΓk[ψk±(vγ)ψk±(vγ+)+O((η+eγ/2)||ψk||β||ψk+||β)]×0TeδktT1nhkfDk+(gt/2v)hk+fDk+(γ1gt/2v)×𝟙𝒱η,R(Ωk)(gt/2v)𝟙𝒱η,R(Ωk+)(γ1gt/2v)𝑑m~BMk𝑑t.\begin{split}\sum_{\gamma\in\Gamma_{k}}&[\psi_{k}^{\pm}(v_{\gamma}^{-})\psi_{k}^{\pm}(v^{+}_{\gamma})+O((\eta+e^{-\ell_{\gamma}/2})||\psi^{-}_{k}||_{\beta}||\psi^{+}_{k}||_{\beta})]\times\\ &\int_{0}^{T}e^{\delta_{k}t}\int_{T^{1}\mathbb{H}^{n}}h^{-}_{k}\circ f^{+}_{D_{k}^{-}}(g^{-t/2}v)h^{+}_{k}\circ f^{-}_{D_{k}^{+}}(\gamma^{-1}g^{t/2}v)\times\mathbbm{1}_{\mathcal{V}_{\eta,R}(\Omega_{k}^{-})}(g^{-t/2}v)\mathbbm{1}_{\mathcal{V}_{\eta,R}(\Omega^{+}_{k})}(\gamma^{-1}g^{t/2}v)d\tilde{m}_{\rm{BM}}^{k}dt.\end{split}

By the Γ\Gamma-invariance of hk±h^{\pm}_{k}, one has

hkfDk+(gt/2v)=eδk(t/2)hk,et/2R(gt/2wk),h^{-}_{k}\circ f^{+}_{D^{-}_{k}}(g^{-t/2}v)=e^{-\delta_{k}(t/2)}h^{-}_{k,e^{-t/2}R}(g^{t/2}w^{-}_{k}),
hk+fDk+(γ1gt/2v)=eδk(t/2)hk,et/2R+(gt/2wk+)h^{+}_{k}\circ f^{-}_{D^{+}_{k}}(\gamma^{-1}g^{t/2}v)=e^{-\delta_{k}(t/2)}h^{+}_{k,e^{-t/2}R}(g^{-t/2}w^{+}_{k})

where wk=fDk+(v)w^{-}_{k}=f^{+}_{D^{-}_{k}}(v), wk+=fγDk+(v)=γfDk+(γ1v)w^{+}_{k}=f^{-}_{\gamma D^{+}_{k}}(v)=\gamma f^{-}_{D^{+}_{k}}(\gamma^{-1}v), and hk,et/2Rh^{-}_{k,e^{-t/2}R} is defined the same as in (19) except we replace RR by et/2Re^{-t/2}R . Therefore,

hkfDk+(gt/2v)hk+fDk+(γ1gt/2v)=eδkthk,et/2R(gt/2wk)hk,et/2R+(gt/2wk+).h^{-}_{k}\circ f^{+}_{D_{k}^{-}}(g^{-t/2}v)h^{+}_{k}\circ f^{-}_{D_{k}^{+}}(\gamma^{-1}g^{t/2}v)=e^{-\delta_{k}t}h^{-}_{k,e^{-t/2}R}(g^{t/2}w^{-}_{k})h^{+}_{k,e^{-t/2}R}(g^{-t/2}w^{+}_{k}).

The remaining part 𝟙𝒱η,R(Ωk)(gt/2v)𝟙𝒱η,R(Ωk+)(γ1gt/2v)\mathbbm{1}_{\mathcal{V}_{\eta,R}(\Omega_{k}^{-})}(g^{-t/2}v)\mathbbm{1}_{\mathcal{V}_{\eta,R}(\Omega^{+}_{k})}(\gamma^{-1}g^{t/2}v) if nonzero if and only if

vgt/2𝒱η,R(Ωk)γgt/2𝒱η,R(Ωk+)=𝒱η,et/2R(gt/2Ωk)𝒱η,et/2R(γgt/2Ωk+).v\in g^{t/2}\mathcal{V}_{\eta,R}(\Omega_{k}^{-})\cap\gamma g^{-t/2}\mathcal{V}_{\eta,R}(\Omega_{k}^{+})=\mathcal{V}_{\eta,e^{-t/2}R}(g^{t/2}\Omega^{-}_{k})\cap\mathcal{V}_{\eta,e^{-t/2}R}(\gamma g^{-t/2}\Omega^{+}_{k}).

By [PP17, Lemma 7], there exist constants t0>0t_{0}>0 and c0c_{0} (independent of kk), such that if tt0t\geq t_{0}, the followings holds: there exists a common perpendicular αγ\alpha_{\gamma} from DkD^{-}_{k} to γ(Dk+)\gamma(D^{+}_{k}) with

  1. (1)

    |γt|2η+c0et/2|\ell_{\gamma}-t|\leq 2\eta+c_{0}e^{-t/2},

  2. (2)

    d(π(vγ±),π(wk±))c0et/2,d(\pi(v^{\pm}_{\gamma}),\pi(w^{\pm}_{k}))\leq c_{0}e^{-t/2},

  3. (3)

    d(π(g±t/2wk),π(v))η+c0et/2d(\pi(g^{\pm t/2}w^{\mp}_{k}),\pi(v))\leq\eta+c_{0}e^{-t/2}.

For all γΓk\gamma\in\Gamma_{k} and Tt0T\geq t_{0}, we define

𝒜k,γ(T)={(t,v)[t0,T]×T1n:v𝒱η,et/2R(gt/2Ωk)𝒱η,et/2R(γgt/2Ωk+)},\mathcal{A}_{k,\gamma}(T)=\{(t,v)\in[t_{0},T]\times T^{1}\mathbb{H}^{n}:v\in\mathcal{V}_{\eta,e^{-t/2}R}(g^{t/2}\Omega^{-}_{k})\cap\mathcal{V}_{\eta,e^{-t/2}R}(\gamma g^{-t/2}\Omega^{+}_{k})\},

and the integral

(22) jk,γ(T)\displaystyle j_{k,\gamma}(T) =(t,v)𝒜k,γ(T)hk,et/2R(gt/2wk)hk,et/2R+(gt/2wk+)𝑑t𝑑m~BMk(v)\displaystyle=\int\int_{(t,v)\in\mathcal{A}_{k,\gamma}(T)}h^{-}_{k,e^{-t/2}R}(g^{t/2}w^{-}_{k})h^{+}_{k,e^{-t/2}R}(g^{-t/2}w^{+}_{k})dtd\tilde{m}_{BM}^{k}(v)
=1(2η)2(t,v)𝒜k,γ(T)dtdm~BMk(v)μW+(wt)(B+(wt,rt))μW(wt+)(B(wt+,rt))\displaystyle=\dfrac{1}{(2\eta)^{2}}\int\int_{(t,v)\in\mathcal{A}_{k,\gamma}(T)}\dfrac{dtd\tilde{m}^{k}_{BM}(v)}{\mu_{W^{+}(w^{-}_{t})}(B^{+}(w^{-}_{t},r_{t}))\mu_{W^{-}(w^{+}_{t})}(B^{-}(w^{+}_{t},r_{t}))}

where

rt=et/2R,wt=gt/2wk,wt+=gt/2wk+.r_{t}=e^{-t/2}R,\quad w^{-}_{t}=g^{t/2}w^{-}_{k},\quad w^{+}_{t}=g^{-t/2}w^{+}_{k}.

There exists then a constant c0′′>0c^{\prime\prime}_{0}>0 (independent of kk) such that for Tt0T\geq t_{0}, one has

(23) c0′′+γΓTO(η+eγ/2),O(η+eγ/2),k[ψk±(vγ)ψk±(vγ+)+O((η+eγ/2)||ψk||β||ψk+||β)]jk,γ(T)ik(T)c0′′+γΓT+O(η+eγ/2),O(η+eγ/2),k[ψk±(vγ)ψk±(vγ+)+O((η+eγ/2)||ψk||β||ψk+||β)]jk,γ(T+O(η+eγ/2)),\begin{split}-c^{\prime\prime}_{0}+\sum_{\gamma\in\Gamma_{T-O(\eta+e^{-\ell_{\gamma}/2}),-O(\eta+e^{-\ell_{\gamma}/2}),k}}[\psi_{k}^{\pm}(v_{\gamma}^{-})\psi_{k}^{\pm}(v^{+}_{\gamma})&+O((\eta+e^{-\ell_{\gamma}/2})||\psi^{-}_{k}||_{\beta}||\psi^{+}_{k}||_{\beta})]j_{k,\gamma}(T)\\ &\leq i_{k}(T)\leq\\ c^{\prime\prime}_{0}+\sum_{\gamma\in\Gamma_{T+O(\eta+e^{-\ell_{\gamma}/2}),O(\eta+e^{-\ell_{\gamma}/2}),k}}[\psi_{k}^{\pm}(v_{\gamma}^{-})\psi_{k}^{\pm}(v^{+}_{\gamma})&+O((\eta+e^{-\ell_{\gamma}/2})||\psi^{-}_{k}||_{\beta}||\psi^{+}_{k}||_{\beta})]j_{k,\gamma}(T+O(\eta+e^{-\ell_{\gamma}/2})),\end{split}

where Γs,r,k={γΓk|t0+2+c0γs,vγ±NrΩ±}\Gamma_{s,r,k}=\{\gamma\in\Gamma_{k}|t_{0}+2+c_{0}\leq\ell_{\gamma}\leq s,v^{\pm}_{\gamma}\in N_{r}\Omega^{\pm}\} for all s,rs,r\in\mathbb{R}.

Claim 5.6.

For any ϵ>0\epsilon>0, if η\eta is small enough and γ\ell_{\gamma} is large enough, then

jk,γ(T)=eO(η+eγ/2)eO(ϵc)(2η+O(eγ/2))2(2η)2,j_{k,\gamma}(T)=e^{O(\eta+e^{-\ell_{\gamma}/2})}e^{O(\epsilon^{c^{\prime}})}\dfrac{(2\eta+O(e^{-\ell_{\gamma}/2}))^{2}}{(2\eta)^{2}},

for c>0c^{\prime}>0 independent of kk.

Proof.

Since (n,Γk)(\mathbb{H}^{n},\Gamma_{k}) has radius-continous strong stable/unstable ball masses. By [PP17, Lemma 11], for every ϵ>0\epsilon>0 and every (t,v)𝒜k,γ(T)(t,v)\in\mathcal{A}_{k,\gamma}(T), one has

μW±(wt)(B±(wt,rt))=eO(ϵ)μW±(vγ)(B±(vγ,rγ))\mu_{W^{\pm}(w_{t}^{\mp})}(B^{\pm}(w^{\mp}_{t},r_{t}))=e^{O(\epsilon)}\mu_{W^{\pm}(v_{\gamma})}(B^{\pm}(v_{\gamma},r_{\ell_{\gamma}}))

if η\eta is small enough and γ\ell_{\gamma} is large enough, independent of kk. Here vγv_{\gamma} denote the midpoint of the common perpendicular from DkD^{-}_{k} to γ(Dk+)\gamma(D^{+}_{k}). Hence,

jk,γ(T)=eO(ϵ)(t,v)𝒜k,γ(T)𝑑t𝑑m~BMk(v)μW+(vγ)(B+(vγ,rγ))μW(vγ)(B(vγ,rγ))j_{k,\gamma}(T)=\dfrac{e^{O(\epsilon)}\int\int_{(t,v)\in\mathcal{A}_{k,\gamma}(T)}dtd\tilde{m}_{\rm{BM}}^{k}(v)}{\mu_{W^{+}(v_{\gamma})}(B^{+}(v_{\gamma},r_{\ell_{\gamma}}))\mu_{W^{-}(v_{\gamma})}(B^{-}(v_{\gamma},r_{\ell_{\gamma}}))}

By [PP17, Lemma 10], for every (t,v)𝒜k,γ(T)(t,v)\in\mathcal{A}_{k,\gamma}(T), one has

dtdm~BMk(v)=eO(η+eγ/2)dtdsdμW(vγ)(v)dμW+(vγ)(v′′)dtdtd\tilde{m}_{\rm{BM}}^{k}(v)=e^{O(\eta+e^{-\ell_{\gamma}/2})}dtdsd\mu_{W^{-}(v_{\gamma})}(v^{\prime})d\mu_{W^{+}(v_{\gamma})}(v^{\prime\prime})dt

where v=fHB(vγ)+(v)v^{\prime}=f^{+}_{HB_{-}(v_{\gamma})}(v) and v′′=fHB+(vγ)(v)v^{\prime\prime}=f^{-}_{HB_{+}(v_{\gamma})}(v). By [PP17, Lemma 9], the distances d(v,vγ),d(v,vγ)d(v,v_{\gamma}),d(v^{\prime},v_{\gamma}) and d(v′′,vγ)d(v^{\prime\prime},v_{\gamma}) are O(η+et/2)O(\eta+e^{-t/2}). Combining these equations together, the claim follows. ∎

Applying then Claim 5.6 in equation (23) we get

(24) Ik(t)=γΓk[ψk±(vγ)ψk±(vγ+)]+O(ηeδkt),\begin{split}I_{k}(t)=\sum_{\gamma\in\Gamma_{k}}&[\psi_{k}^{\pm}(v_{\gamma}^{-})\psi_{k}^{\pm}(v^{+}_{\gamma})]+O(\eta e^{\delta_{k}t}),\end{split}

for tt sufficiently large independent of kk, and O(.)O(.) independent of k,ηk,\eta. Then by multiplying eδkte^{-\delta_{k}t} to equations (16), (24) we get that for fixed η>0\eta>0

(25) σDk+(ψk)σDk+(ψk+)δ(Γk)mBMkO(η)Nψk,ψk+(t)eδ(Γk)tσDk+(ψk)σDk+(ψk+)δ(Γk)mBMk+O(η)\frac{\sigma^{+}_{D^{-}_{k}}(\psi^{-}_{k})\cdot\sigma^{-}_{D^{+}_{k}}(\psi^{+}_{k})}{\delta(\Gamma_{k})||m^{k}_{\rm{BM}}||}-O(\eta)\leq\frac{N_{\psi^{-}_{k},\psi^{+}_{k}}(t)}{e^{\delta(\Gamma_{k})t}}\leq\frac{\sigma^{+}_{D^{-}_{k}}(\psi^{-}_{k})\cdot\sigma^{-}_{D^{+}_{k}}(\psi^{+}_{k})}{\delta(\Gamma_{k})||m^{k}_{\rm{BM}}||}+O(\eta)

for tt sufficiently large independent of kk and O(.)O(.) independent of k,ηk,\eta, from where the result follows.

More about the proof of Theorem 5.3: As (Dk±)k(D^{\pm}_{k})_{k\in\mathbb{N}} is a sequence of well-positioned convex sets in MkM_{k} which converges strongly to a well-positioned set D±D^{\pm} in MM, we can select ψk±C0(T1Mk)\psi^{\pm}_{k}\in C^{\infty}_{0}(T^{1}M_{k}), ψ±C0(T1M)\psi^{\pm}\in C^{\infty}_{0}(T^{1}M) be compactly supported functions so that (ψk±)k(\psi^{\pm}_{k})_{k\in\mathbb{N}} converges strongly to ψ±\psi^{\pm} and ψk±1\psi^{\pm}_{k}\equiv 1 in supp(σDk±)supp(\sigma^{\mp}_{\partial D^{\pm}_{k}}). Hence in the notation of Theorem 5.5

𝒩ψk,ψk+(t)=𝒩Dk,Dk+(t),σDk±(ψk±)=σDk±,\mathcal{N}_{\psi^{-}_{k},\psi^{+}_{k}}(t)=\mathcal{N}_{D^{-}_{k},D^{+}_{k}}(t),\quad\sigma^{\mp}_{D^{\pm}_{k}}(\psi^{\pm}_{k})=\|\sigma^{\mp}_{D^{\pm}_{k}}\|,

and in particular σDk+(ψk)σDk+(ψk+)δ(Γk)mBMk0\frac{\sigma^{+}_{D^{-}_{k}}(\psi^{-}_{k})\cdot\sigma^{-}_{D^{+}_{k}}(\psi^{+}_{k})}{\delta(\Gamma_{k})||m^{k}_{\rm{BM}}||}\neq 0. Then we can restate the conclusion of Theorem 5.3 by a multiplicative error uniformly close to 1 along the sequence as tt gets larger.

References

  • [ABB+17] Miklos Abert, Nicolas Bergeron, Ian Biringer, Tsachik Gelander, Nikolay Nikolav, Jean Raimbault, and Iddo Samet, On the growth of L2L^{2}-invariants for sequences of lattices in Lie groups, Annals of Mathematics 185 (2017), no. 3, 711 – 790.
  • [Bab02] Martine Babillot, On the mixing property for hyperbolic systems, Israel J. Math. 129 (2002), 61–76. MR 1910932
  • [BBB19] Martin Bridgeman, Jeffrey Brock, and Kenneth Bromberg, Schwarzian derivatives, projective structures, and the Weil-Petersson gradient flow for renormalized volume, Duke Math. J. 168 (2019), no. 5, 867–896. MR 3934591
  • [BBP21] Martin Bridgeman, Kenneth Bromberg, and Franco Vargas Pallete, The Weil-Petersson gradient flow of renormalized volume on a Bers slice has a global attracting fixed point, 2021.
  • [Bow93] Brian H. Bowditch, Geometrical finiteness for hyperbolic groups, J. Funct. Anal. 113 (1993), no. 2, 245–317. MR 1218098
  • [Bow95] by same author, Geometrical finiteness with variable negative curvature, Duke Mathematical Journal 77 (1995), no. 1, 229–274.
  • [CT99] Richard D. Canary and Edward C. Taylor, Hausdorff dimension and limits of Kleinian groups, Geom. Funct. Anal. 9 (1999), no. 2, 283–297. MR 1692482
  • [DOP00] Françoise Dal’bo, Jean-Pierre Otal, and Marc Peigné, Séries de Poincaré des groupes géométriquement finis, Israel J. Math. 118 (2000), 109–124. MR 1776078
  • [EO21] Sam Edwards and Hee Oh, Spectral gap and exponential mixing on geometrically finite hyperbolic manifolds, Duke Math. J. 170 (2021), no. 15, 3417–3458. MR 4324181
  • [Ham89] Ursula Hamenstädt, A new description of the Bowen-Margulis measure, Ergodic Theory Dynam. Systems 9 (1989), no. 3, 455–464. MR 1016663
  • [Ham04] by same author, Small eigenvalues of geometrically finite manifolds, J. Geom. Anal. 14 (2004), no. 2, 281–290. MR 2051688
  • [LP82] Peter D. Lax and Ralph S. Phillips, The asymptotic distribution of lattice points in Euclidean and non-Euclidean spaces, J. Functional Analysis 46 (1982), no. 3, 280–350. MR 661875
  • [McM99] Curtis T. McMullen, Hausdorff dimension and conformal dynamics. I. Strong convergence of Kleinian groups, J. Differential Geom. 51 (1999), no. 3, 471–515. MR 1726737
  • [Pat76] Samuel J. Patterson, The limit set of a Fuchsian group, Acta Math. 136 (1976), no. 3-4, 241–273. MR 450547
  • [PP17] Jouni Parkkonen and Frédéric Paulin, Counting common perpendicular arcs in negative curvature, Ergodic Theory Dynam. Systems 37 (2017), no. 3, 900–938. MR 3628925
  • [Rob00] Thomas Roblin, Sur l’ergodicité rationnelle et les propriétés ergodiques du flot géodésique dans les variétés hyperboliques, Ergodic Theory Dynam. Systems 20 (2000), no. 6, 1785–1819. MR 1804958
  • [Rob03] by same author, Ergodicité et équidistribution en courbure négative, Mém. Soc. Math. Fr. (N.S.) (2003), no. 95, vi+96. MR 2057305
  • [Sul79] Dennis Sullivan, The density at infinity of a discrete group of hyperbolic motions, Inst. Hautes Études Sci. Publ. Math. (1979), no. 50, 171–202. MR 556586
  • [Sul84] by same author, Entropy, Hausdorff measures old and new, and limit sets of geometrically finite Kleinian groups, Acta Math. 153 (1984), no. 3-4, 259–277. MR 766265
  • [Sul87] by same author, Related aspects of positivity in Riemannian geometry, J. Differential Geom. 25 (1987), no. 3, 327–351. MR 882827