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

Thurston’s Bounded image theorem

Cyril Lecuire and Ken’ichi Ohshika Laboratoire Emile Picard, Université Paul Sabatier, 118 Route de Narbonne, 31062 Toulouse Cedex 4, France, and Department of Mathematics, Faculty of Science, Gakushuin University, Toshima-ku, Tokyo 171-8588, Japan
Abstract.

Thurston’s bounded image theorem is one of the key steps in his proof of the uniformisation theorem for Haken manifolds. Thurston never published its proof, and no proof has been known up to today, although a proof of its weaker version, called the bounded orbit theorem is known. In this paper, we give a proof of the original bounded image theorem, relying on recent development of Kleinian group theory.

1. Introduction

From the late 1970s to the early 1980s, Thurston gave lectures on his uniformisation theorem for Haken manifolds ([20, 21]). The theorem states that every atoroidal Haken 3-manifold with its (possibly empty) boundary consisting only of incompressible tori admits a complete hyperbolic metric in its interior. His proof of this theorem is based on an induction making use of a hierarchy for Haken manifolds invented by Waldhausen [24], i.e., a system of incompressible surfaces cutting the manifold down to balls, together with Maskit’s combination theorem (see for instance [10, §VII]).

For simplicity, we now focus on the case of closed atoroidal Haken manifolds. In the last step of the induction, we are in the situation where NN is a closed atoroidal Haken manifold obtained from a 3-manifold MM with non-empty boundary (without torus components) by pasting M\partial M to itself by an orientation reversing involution. The induction hypothesis guarantees the existence of a convex cocompact hyperbolic structure on MM. There, Thurston used the so-called bounded image theorem to find a convex compact hyperbolic structure on MM, obtained by quasi-conformally deforming the given hyperbolic structure, which can be pasted up along M\partial M to give a hyperbolic structure on NN.

Let us explain the setting in more detail. Let MM be an atoroidal Haken manifold with an even number of boundary components all of which are incompressible. In the same way as we assumed that NN is closed, we assume that no boundary component of MM is a torus, for simplicity. Suppose that there is an orientation-reversing involution ι:MM\iota:\partial M\to\partial M taking each component of M\partial M to another one. Let NN be the closed manifold obtained from MM by identifying the points on M\partial M with their images under ι\iota. Suppose moreover that NN is also atoroidal.

We assume, as the hypothesis of induction, that MM admits a convex compact hyperbolic structure; in other words, that the interior of MM is homeomorphic to 3/Γ\mathbb{H}^{3}/\Gamma for a convex cocompact Kleinian group Γ\Gamma. The space of convex compact hyperbolic structures on MM, which is not empty by assumption, modulo isotopy is parameterised by 𝒯(M)\mathcal{T}(\partial M), as can be seen in the works of Ahlfors, Bers, Kra, Maskit, Marden and Sullivan. From each convex compact hyperbolic structure on MM, by taking the covering of MM associated with each component SS of M\partial M, we get a quasi-Fuchsian group isomorphic to π1(S)\pi_{1}(S), and by considering the second coordinate of the parameterisation 𝒯(S)×𝒯(S¯)\mathcal{T}(S)\times\mathcal{T}(\bar{S}) of the quasi-Fuchsian space, we obtain a map from 𝒯(M)\mathcal{T}(\partial M) to 𝒯(S¯)\mathcal{T}(\bar{S}), where 𝒯(S¯)\mathcal{T}(\bar{S}) denotes the Teichmüller space of SS with orientation reversed. By considering this for every component of M\partial M, we get a map σ:𝒯(M)𝒯(¯M)\sigma:\mathcal{T}(\partial M)\to\mathcal{T}(\bar{\partial}M) called the skinning map, where 𝒯(¯M)\mathcal{T}(\bar{\partial}M) denotes the product of 𝒯(S¯)\mathcal{T}(\bar{S}) for the components SS of M\partial M. Since ι\iota is orientation-reversing, it induces a homeomorphism ι:𝒯(¯M)𝒯(M)\iota_{*}\colon\mathcal{T}(\bar{\partial}M)\to\mathcal{T}(\partial M).

Then the bounded image theorem can be stated as follows.

Theorem 1.1.

Suppose that MM is a compact (orientable) atoroidal Haken manifold having an even number of boundary components all of which are incompressible and none of which are tori, and assume that MM is not homeomorphic to an II-bundle over a closed surface. Assume moreover that MM admits a convex compact hyperbolic structure. Suppose that there is an orientation reversing involution ι:MM\iota:\partial M\to\partial M taking each component of M\partial M to another component, and that by pasting each component of M\partial M to its image under ι\iota, we get a closed atoroidal manifold NN. Then there exists nn\in\mathbb{N} depending only on the topological type of MM such that the image of (ισ)n(\iota_{*}\circ\sigma)^{n} is bounded (i.e. precompact) in 𝒯(M)\mathcal{T}(\partial M).

There are several expository papers and books on Thurston’s uniformisation theorem ([16, 18, 7] among others). In all of them, a weaker version of the bounded image theorem called the bounded orbit theorem, which is sufficient for the proof of the uniformisation theorem, was proved and used, instead of this original one.

Up to now, no complete proof of the bounded image theorem as stated above was known. Kent [8] gave a proof of this theorem under the assumption that MM is acylindrical, in which case the deformation space of hyperbolic structures on MM is compact.

The purpose of this paper is to give a proof of the original bounded image theorem. Our argument relies on recent progress in Klenian group theory, in particular, the embedding of partial cores in the geometric limit from [4], the relation between the presence of short curves and their relative positions and the behaviour of ends invariant from [3], and criteria of convergence/divergence given in [2].

1.1. Outline

We are going to find nn such that if the image of (ισm)n(\iota_{*}\circ\sigma_{m})^{n} is unbounded then NN contains a non-peripheral incompressible torus, contradicting our assumption. For that purpose we shall use the invariant mm introduced in [2].

Given a simple closed curve dd on a closed surface SS equipped with a hyperbolic metric gg, we define

m(g,d,μ)=max{supY:dYdY(μ(g),μ),1lengthg(d)},m(g,d,\mu)=\max\left\{\sup_{\begin{subarray}{c}Y:\ d\subset\partial Y\end{subarray}}d_{Y}(\mu(g),\mu),{1\over\mathrm{length}_{g}(d)}\right\},

where μ(g)\mu(g) is a shortest marking for (S,g)(S,g), μ\mu is a full marking, and the supremum of the first term in the maximum is taken over all incompressible subsurfaces YY of SS whose boundaries Y\partial Y contain dd. See Definition 4.1 for more details.

It is not difficult to see that in the setting of Theorem 1.1, for a given sequence {mi}\{m_{i}\} in 𝒯(M)\mathcal{T}(\partial M), if the sequence {σ(mi)}\{\sigma(m_{i})\} is unbounded, then there is a simple closed curve dd such that m(σ(mi),d,μ)m(\sigma(m_{i}),d,\mu) is unbounded (see Lemma 4.3). The core of our argument consists in showing, with the help of arguments from [3] and [2], that in this situation, there is a simple closed curve dMd^{\prime}\subset\partial M such that {m(mi,d,μ)}\{m(m_{i},d^{\prime},\mu)\} is unbounded and that ddd\cup d^{\prime} bounds an essential annulus in MM. Using this argument repeatedly, we build (when {(ισ)n(mi)}\{(\iota_{*}\circ\sigma)^{n}(m_{i})\} is unbounded) an annulus in NN which goes through the interior of MM (viewed as a subset of NN) nn times. If nn is large enough, this annulus must create an essential torus in NN, and contradicts the assumption that NN is atoroidal.

Although this is the overall logic of the proof, in the following sections, we shall present the main steps in a different order. After setting up some preliminary definitions in Section 2, we shall discuss the topological part of the proof in Section 3. First we show that we can add some assumptions on the topology of MM which will simplify the arguments later on. Next, we study incompressible surfaces on M\partial M which can be extended multiple times through the characteristic submanifold of MM when it is viewed as a submanifold of NN. This will give us an integer nn which appears in Theorem 1.1. In Section 4 we shall discuss the relation between the behaviour of the invariant mm defined above, and the convergence and divergence of Kleinian groups. In Section 5 we shall prove our key proposition, and obtain the curve dd^{\prime} mentioned above. Finally in Section 6 we shall put these pieces together to prove our main theorem.

2. Preliminaries

2.1. Haken manifolds and characteristic submanifolds

An orientable irreducible compact 3-manifold which contains a non-peripheral incompressible surface is called a Haken manifold. We note that a compact irreducible 3-manifold with non-empty boundary is always Haken except for a 3-ball. We say that a Haken manifold is atoroidal when it does not contain a non-peripheral incompressible torus, and acylindrical when it does not contain a non-peripheral incompressible annulus. By the torus theorem for Haken manifolds ([25, 5, 6]), the former condition of the atoroidality is equivalent to the one that every monomorphism from ×\mathbb{Z}\times\mathbb{Z} into the fundamental group is peripheral, i.e. is conjugate to the image of the fundamental group of a boundary component.

The Jaco-Shalen-Johannson theory [5, 6] tells us that in a Haken manifold, incompressible tori and incompressible annuli can stay only in a very restricted place. Let us state what the theory says in the case when a Haken manifold MM is atoroidal and boundary-irreducible.

For an orientable atoroidal Haken boundary-irreducible 3-manifold MM, there exists a 3-submanifold XX of MM each of whose components is one of the following and which satisfies the following condition:

  1. (a)

    An II-bundle whose associated I\partial I-bundle coincides with its intersection with M\partial M. Such a component is called a characteristic II-pair.

  2. (b)

    A solid torus VV such that VMV\cap\partial M consists of annuli which are incompressible on both V\partial V and M\partial M. When VMV\cap\partial M is connected, it winds around the core curve of VV more than once.

  3. (c)

    A thickened torus S1×S1×IS^{1}\times S^{1}\times I at least one of whose boundary components lies on a component of M\partial M.

Every properly embedded essential annulus (i.e. an incompressible annulus which is not homotopic into the boundary) is properly isotopic into XX, and no component of XX is properly isotopic into another component.

Such XX is unique up to isotopy, and is called the characteristic submanifold of MM. We note that in the case when MM has no torus boundary component, which is the assumption of our main theorem, a component of the last type (c) does not appear.

Thurston’s celebrated uniformisation theorem for Haken manifolds says that every atoroidal Haken manifold whose boundary consists of incompressible tori admits a hyperbolic structure of finite volume. More generally, he proved that every atoroidal Haken manifold, including the case when it has non-torus boundary components, admits a (minimally parabolic) convex hyperbolic structure of finite volume. The term ‘convex hyperbolic structure’ will be explained in the following subsection.

2.2. Kleinian groups and their deformation spaces

A Kleinian group is a discrete subgroup of PSL2(){\rm PSL}_{2}(\mathbb{C}). In this paper, we always assume Kleinian groups to be torsion free, and finitely generated except for the case when we talk about geometric limits. For a Kleinian group Γ\Gamma, we can consider the complete hyperbolic 3-manifold 3/Γ\mathbb{H}^{3}/\Gamma. The convex core of 3/Γ\mathbb{H}^{3}/\Gamma is the smallest convex submanifold that is a deformation retract. The Kleinian group Γ\Gamma and the corresponding hyperbolic 3-manifold 3/Γ\mathbb{H}^{3}/\Gamma are said to be geometrically finite when the convex core of 3/Γ\mathbb{H}^{3}/\Gamma has finite volume. In particular, 3/Γ\mathbb{H}^{3}/\Gamma is said to be convex compact, and Γ\Gamma to be convex cocompact if the convex core is compact. We also say that Γ\Gamma is minimally parabolic when every parabolic element in Γ\Gamma is contained in a rank-2 parabolic subgroup. Any convex cocompact Kleinian group is automatically minimally parabolic since it does not have parabolic elements.

A 3-manifold MM is said to have a hyperbolic structure when IntM\operatorname{Int}M is homeomorphic to 3/Γ\mathbb{H}^{3}/\Gamma for a Kleinian group Γ\Gamma, and we regard the pull-back of the hyperbolic metric to IntM\operatorname{Int}M as a hyperbolic structure on MM. In particular if Γ\Gamma is taken to be geomerically finite or convex cocompact, we say that MM has a geometrically finite or convex compact hyperbolic structure. If MM admits a hyperbolic structure, then MM must be atoroidal.

The set of hyperbolic structures on MM modulo isotopy, which we denote by 𝖠𝖧(M)\mathsf{AH}(M), can be identified with a subset of the set of faithful discrete representations of π1(M)\pi_{1}(M) into PSL2(){\rm PSL}_{2}(\mathbb{C}) modulo conjugacy. We put on 𝖠𝖧(M)\mathsf{AH}(M) a topology induced from the weak topology on the representation space. We regard an element of 𝖠𝖧(M)\mathsf{AH}(M) both as a hyperbolic structure on MM and as a representation of π1(M)\pi_{1}(M) into PSL2(){\rm PSL}_{2}(\mathbb{C}) depending on the situation.

A Kleinian group GG is said to be a quasi-conformal deformation of another Kleinian group Γ\Gamma if there is a quasi-conformal homeomorphism f:^^f\colon\hat{\mathbb{C}}\to\hat{\mathbb{C}} such that G=fΓf1G=f\Gamma f^{-1} as Möbius transformations on ^\hat{\mathbb{C}}. When GG is a quasi-conformal deformation of Γ\Gamma, there is a diffeomorphism from 3/Γ\mathbb{H}^{3}/\Gamma to 3/G\mathbb{H}^{3}/G preserving the parabolicity in both directions, which induces an isomorphism between the fundamental groups coinciding with the isomorphism given by the conjugacy G=fΓf1G=f\Gamma f^{-1}. We note that a quasi-conformal deformation of geometrically finite (resp. convex cocompact, minimally parabolic geometrically finite) group is again geometrically finite (resp. convex cocompact, minimally parabolic geometrically finite).

Let MM be a compact 3-manifold admitting a minimally parabolic geometrically finite hyperbolic structure mm. Let 𝖰𝖧(M)\mathsf{QH}(M) denote the set of all minimally parabolic geometrically finite hyperbolic structures on MM modulo isotopy, which is regarded as a subset of 𝖠𝖧(M)\mathsf{AH}(M). Marden [9] showed that every minimally parabolic geometrically finite hyperbolic structures on MM is obtained as a quasi-conformal deformation of mm. Therefore we call 𝖰𝖧(M)\mathsf{QH}(M) the quasi-conformal deformation space. Furthermore, if M\partial M is incompressible, combined with the work of Ahlfors, Bers, Kra, Maskit and Sullivan, there is a parameterisation q:𝒯(M)𝖰𝖧(M)q\colon\mathcal{T}(\partial M)\to\mathsf{QH}(M), where 𝒯(M)\mathcal{T}(\partial M) denotes the Teichmüller space of M\partial M, i.e. the direct product of the Teichmüller spaces of the components of M\partial M. We shall refer to this map as the Ahlfors-Bers map.

In the case when MM is homeomorphic to S×[0,1]S\times[0,1] for a closed oriented surface SS, the deformation spaces 𝖠𝖧(M),𝖰𝖧(M)\mathsf{AH}(M),\mathsf{QH}(M) are denoted by 𝖠𝖧(S),𝖰𝖥(S)\mathsf{AH}(S),\mathsf{QF}(S) respectively. The quasi-conformal deformation space 𝖰𝖥(S)\mathsf{QF}(S) consists of quasi-Fuchsian representations of π1(S)\pi_{1}(S), i.e. quasi-conformal deformations of a Fuchsian representation, and is therefore called the quasi-Fuchsian space. The Ahlfors-Bers map can be expressed as qf:𝒯(S)×𝒯(S¯)𝖰𝖥(S)qf:\mathcal{T}(S)\times\mathcal{T}(\bar{S})\to\mathsf{QF}(S), where the second coordinate 𝒯(S¯)\mathcal{T}(\bar{S}) denotes the Teichmüller space of SS with orientation reversed, which is a more natural way for parametrisation since the boundary component S×{1}S\times\{1\} has the opposite orientation from the one given on S×{0}S\times\{0\} if we identify them with SS by dropping the second factor.

Now, let MM be an atoroidal Haken 3-manifold with non-empty incompressible boundary which does not contain a torus. Suppose that MM has a convex compact hyperbolic metric mm, and let SS be a component of M\partial M. Take a covering of MM associated with π1(S)π1(M)\pi_{1}(S)\subset\pi_{1}(M), and lift the hyperbolic structure mm to the hyperbolic structure m~\tilde{m} on S×[0,1]S\times[0,1]. It is known (see [16, Proposition 7.1]) that the lifted structure m~\tilde{m} is also convex cocompact, hence can be regarded as an element of 𝖰𝖥(S)\mathsf{QF}(S). Therefore m~\tilde{m} in turn corresponds to a point (gS(m),hS(m))(g_{S}(m),h_{S}(m)) in 𝒯(S)×𝒯(S¯)\mathcal{T}(S)\times\mathcal{T}(\bar{S}). Let S1,,SkS_{1},\dots,S_{k} be the components of M\partial M that are not tori, and we consider the point hSi(m)𝒯(S¯i)h_{S_{i}}(m)\in\mathcal{T}(\bar{S}_{i}) for each i=1,,ki=1,\dots,k. We define 𝒯(¯M)\mathcal{T}(\bar{\partial}M) to be 𝒯(S¯1)××𝒯(S¯k)\mathcal{T}(\bar{S}_{1})\times\dots\times\mathcal{T}(\bar{S}_{k}). The map taking g𝒯(M)g\in\mathcal{T}(\partial M) to (hS1(q(g)),,hSk(q(g)))𝒯(¯M)(h_{S_{1}}(q(g)),\dots,h_{S_{k}}(q(g)))\in\mathcal{T}(\bar{\partial}M) is called the skinning map, which we shall denote by σ\sigma.

2.3. Curve complexes and projections

Let SS be a connected compact orientable surface possibly with boundary, satisfying ξ(S)=3g+n4\xi(S)=3g+n\geq 4 where gg denotes the genus and nn denotes the number of the boundary components. The curve complex 𝒞𝒞(S)\mathcal{CC}(S) of SS with ξ(S)5\xi(S)\geq 5 is a simplicial complex whose vertices are isotopy classes of non-peripheral, non-contractible simple closed curves on SS such that n+1n+1 vertices span an nn-simplex when they are represented by pairwise disjoint simple closed curves. In the case when ξ(S)=4\xi(S)=4, we define 𝒞𝒞(S)\mathcal{CC}(S) to be a graph whose vertices are isotopy classes of simple closed curves such that two vertices have smallest possible intersection. In the case when SS is an annulus, we define 𝒞𝒞(S)\mathcal{CC}(S) to be a graph whose vertices are isotopy classes (relative to the endpoints) of non-peripheral simple arcs in SS such that two vertices are connected when they can be represented by arcs which are disjoint at their interiors. Masur-Minsky [11] proved that 𝒞𝒞(S)\mathcal{CC}(S) is Gromov hyperbolic with respect to the path metric for any SS.

A marking μ\mu on SS consists of a pants decomposition of SS, which is denoted by 𝐛𝐚𝐬𝐞(μ)\mathbf{base}(\mu) and whose components are called base curves, and a collection 𝐭(μ)\mathbf{t}(\mu) of simple closed curves, called transversals of 𝐛𝐚𝐬𝐞(μ)\mathbf{base}(\mu), such that each component of 𝐛𝐚𝐬𝐞(μ)\mathbf{base}(\mu) intersects at most one among them essentially. For two markings μ,ν\mu,\nu on SS and a subsurface YY, we define dY(μ,ν)d_{Y}(\mu,\nu) to be the distance between πY(𝐛𝐚𝐬𝐞(μ)𝐭(μ))\pi_{Y}(\mathbf{base}(\mu)\cup\mathbf{t}(\mu)) and πY(𝐛𝐚𝐬𝐞(ν)𝐭(ν))\pi_{Y}(\mathbf{base}(\nu)\cup\mathbf{t}(\nu)), where the projection πY:𝒞𝒞(S)𝔓(𝒞𝒞(Y))\pi_{Y}\colon\mathcal{CC}(S)\to\mathfrak{P}(\mathcal{CC}(Y)) is obtained by taking the intersection of curves on SS with YY and connecting the endpoints by arcs on FrY\operatorname{Fr}Y when the intersection contains arcs. In [12], a marking defined as such is called clean. In this paper, we only consider clean markings. A marking is called full when every base curve has a transversal. In general, for two sets of simple closed curves a,ba,b and a subsurface YY of SS, we define dY(a,b)d_{Y}(a,b) to be the distance in 𝒞𝒞(Y)\mathcal{CC}(Y) between πY(a)\pi_{Y}(a) and πY(b)\pi_{Y}(b) provided that both of them are non-empty. If one of them is empty, the distance is not defined.

For a point mm in 𝒯(S)\mathcal{T}(S), its shortest marking, which is a full marking and is denoted by μ(m)\mu(m), has a shortest pants decomposition of (S,m)(S,m) as 𝐛𝐚𝐬𝐞(μ(m))\mathbf{base}(\mu(m)), and 𝐭(μ(m))\mathbf{t}(\mu(m)) consisting of shortest transversals, one for each component of 𝐛𝐚𝐬𝐞(μ(m))\mathbf{base}(\mu(m)). When we talk about the distance dYd_{Y} between two points in 𝒯(S)\mathcal{T}(S) or between a point in 𝒯(S)\mathcal{T}(S) and a marking, we identify points m𝒯(S)m\in\mathcal{T}(S) with μ(m)\mu(m).

2.4. Geometric limits and compact cores

Let MM be an atoroidal boundary-irreducible Haken 3-manifold. Let {ρi}\{\rho_{i}\} be a sequence of faithful discrete representations of π1(M)\pi_{1}(M) into PSL2(){\rm PSL}_{2}(\mathbb{C}). We define a geometric limit of {ρi(π1(M))}\{\rho_{i}(\pi_{1}(M))\} to be a Kleinian group Γ\Gamma such that every element γ\gamma of Γ\Gamma is a limit of some sequence {giρi(π1(M))}\{g_{i}\in\rho_{i}(\pi_{1}(M))\}, and every convergent sequence {γijρij(π1(M))}\{\gamma_{i_{j}}\in\rho_{i_{j}}(\pi_{1}(M))\} has its limit in Γ\Gamma.

Fixing a point x3x\in\mathbb{H}^{3}, and considering its projections xix_{i} in 3/ρi(π1(M))\mathbb{H}^{3}/\rho_{i}(\pi_{1}(M)) and xx_{\infty} in 3/Γ\mathbb{H}^{3}/\Gamma, the geometric convergence implies the existence of pointed Gromov-Hausdorff convergence of ((3/ρi(π1(M)),xi))((\mathbb{H}^{3}/\rho_{i}(\pi_{1}(M)),x_{i})) to (3/Γ,x)(\mathbb{H}^{3}/\Gamma,x_{\infty}). This latter convergence means that there exist real numbers rir_{i} going to \infty, KiK_{i} converging to 11, and KiK_{i}-bi-Lipschitz diffeomorphisms fif_{i} (called approximate isometries) between rir_{i}-balls Bri(3/ρi(π1(M)),xi)B_{r_{i}}(\mathbb{H}^{3}/\rho_{i}(\pi_{1}(M)),x_{i}) and BKiri(3/Γ,x)B_{K_{i}r_{i}}(\mathbb{H}^{3}/\Gamma,x_{\infty}). Suppose that {ρi}\{\rho_{i}\} converges to ρ:π1(M)PSL2()\rho_{\infty}:\pi_{1}(M)\to{\rm PSL}_{2}(\mathbb{C}) as representations and that {ρi(π1(M))}\{\rho_{i}(\pi_{1}(M))\} converges to Γ\Gamma geometrically. Then, ρ(π1(M))\rho_{\infty}(\pi_{1}(M)) is a subgroup of the geometric limit Γ\Gamma.

For an open irreducible 3-manifold VV with finitely generated fundamental group, a compact 3-dimensional submanifold CC in VV is called a compact core when the inclusion induces an isomorphism between their fundamental groups. The existence of compact cores was proved by Scott [19]. The case which interests us is when VV is a hyperbolic 3-manifold.

Let 3/G\mathbb{H}^{3}/G be a hyperbolic 3-manifold associated with a finitely generated, torsion free Kleinian group GG. By Margulis’s lemma, there is a positive constant ε0\varepsilon_{0} such that the set of points of 3/G\mathbb{H}^{3}/G where the injectivity radii are less than ε0\varepsilon_{0} consists of a finite disjoint union of tubular neighbourhoods of closed geodesics of length less than ε0\varepsilon_{0}, called Margulis tubes, and cusp neighbourhoods each of which is stabilised by a maximal parabolic subgroup of GG, and whose quotient by its stabiliser is homeomorphic to S1×2S^{1}\times\mathbb{R}^{2} when the stabiliser has rank 1, and to S1×S1×S^{1}\times S^{1}\times\mathbb{R} when the stabiliser has rank 2. The former cusp neighbourhood is called a \mathbb{Z}-cusp neighbourhood, and the latter a torus cusp neighbourhood. The union of the cusp neighbourhoods is called the cuspidal part of 3/G\mathbb{H}^{3}/G. The complement of the cuspidal part is called the non-cuspidal part and is denoted by (3/G)0(\mathbb{H}^{3}/G)_{0}. Each boundary component of (3/G)0(\mathbb{H}^{3}/G)_{0} is either an open annulus or a torus. By the relative compact core theorem by McCullough [13], there is a compact core CG(3/G)0C_{G}\subset(\mathbb{H}^{3}/G)_{0} such that for each boundary component BB of (3/G)0(\mathbb{H}^{3}/G)_{0}, the intersection CGBC_{G}\cap B is a core annulus when BB is an open annulus, and is the entire BB when BB is a torus. We call such a compact core a relative compact core of (3/G)0(\mathbb{H}^{3}/G)_{0}.

Let p:3/ρ(π1(M))3/Γp\colon\mathbb{H}^{3}/\rho_{\infty}(\pi_{1}(M))\to\mathbb{H}^{3}/\Gamma be the covering map associated with the inclusion of ρ(π1(M))\rho_{\infty}(\pi_{1}(M)) into the geometric limit Γ\Gamma. Let CC be a relative compact core of (3/ρ(π1(M)))0(\mathbb{H}^{3}/\rho_{\infty}(\pi_{1}(M)))_{0}. Suppose that 3/Γ\mathbb{H}^{3}/\Gamma has a torus cusp neighbourhood TT. We say that 3/ρ(π1(M))\mathbb{H}^{3}/\rho_{\infty}(\pi_{1}(M)) wraps around TT when p|Cp|C is homotoped to an immersion which goes around TT non-trivially, and hence cannot be homotoped to an embedding.

3. Topological features

3.1. Coverings

In this subsection, we shall show that to prove Theorem 1.1, we can put an extra assumption that all the characteristic II-pairs of MM are product bundles.

We consider an atoroidal Haken manifold as is given in Theorem 1.1. Let p:M~Mp\colon\tilde{M}\to M be a finite-sheeted regular covering. Then pp induces the covering map between the boundaries p:M~Mp_{\partial}\colon\partial\tilde{M}\to\partial M. This map induces a proper embedding between Teichmüller spaces, p:𝒯(M)𝒯(M~)p_{\partial}^{*}\colon\mathcal{T}(\partial M)\to\mathcal{T}(\partial\tilde{M}) which is obtained by pulling back the conformal structures by pp_{\partial}. Also the involution ι\iota lifts to an orientation-reversing involution ι~:M~M~\tilde{\iota}\colon\partial\tilde{M}\to\partial\tilde{M} taking each component to another one. Since M~\tilde{M} is also an atoroidal boundary-irreducible Haken manifold, we can consider the skinning map σ~:𝒯(M~)𝒯(¯M~)\tilde{\sigma}\colon\mathcal{T}(\partial\tilde{M})\to\mathcal{T}(\bar{\partial}\tilde{M}).

Lemma 3.1.

If (ι~σ~)n(\tilde{\iota}\circ\tilde{\sigma})^{n} has bounded image for some nn\in\mathbb{N}, then so does (ισ)n(\iota\circ\sigma)^{n}.

Proof.

The map pp^{*}_{\partial} properly embeds 𝒯(M)\mathcal{T}(\partial M) into 𝒯(M~)\mathcal{T}(\partial\tilde{M}). By the definition of the maps σ~\tilde{\sigma} and ι~\tilde{\iota}, we have p(ισ)=(ι~σ~)pp^{*}_{\partial}\circ(\iota\circ\sigma)=(\tilde{\iota}\circ\tilde{\sigma})\circ p^{*}_{\partial}, and hence p(ισ)n=(ι~σ~)npp^{*}_{\partial}\circ(\iota\circ\sigma)^{n}=(\tilde{\iota}\circ\tilde{\sigma})^{n}\circ p^{*}_{\partial}. Therefore, if the image of (ι~σ~)n(\tilde{\iota}\circ\tilde{\sigma})^{n} is bounded, the properness of pp^{*}_{\partial} implies that the image of (ισ)n(\iota\circ\sigma)^{n} must also be bounded. ∎

This result allows us to work on manifolds with topological features that will make the arguments simpler.

Lemma 3.2.

Let MM be an orientable atoroidal Haken manifold with incompressible boundary. Then there is a double covering of MM all of whose characteristic II-pairs are product II-bundles.

Proof.

Let W1,,WpW_{1},\dots,W_{p} be the characteristic II-pairs of MM that are twisted II-bundles. Take their double coverings W~1,,W~p\tilde{W}_{1},\dots,\tilde{W}_{p} corresponding to the orientation double coverings of the base surfaces. For each WjW_{j} among W1,,WpW_{1},\dots,W_{p}, its frontier components (i.e. the closures of the components of WjM\partial W_{j}\setminus\partial M) are annuli. Each of such annuli has two pre-images in W~j\tilde{W}_{j} which are taken to each other by the unique non-trivial covering translation.

Let CC be the closure of a component of M(W1Wp)M\setminus(W_{1}\cup\dots\cup W_{p}). Let A1,,AkA_{1},\dots,A_{k} be the components of C(W1Wp)C\cap(W_{1}\cup\dots\cup W_{p}), which are annuli on l=1pWlM¯\cup_{l=1}^{p}\overline{\partial W_{l}\setminus\partial M}. We prepare two copies C+C^{+} and CC^{-} of CC. Each AjA_{j} among A1,AkA_{1},\dots A_{k}, which is contained some WiW_{i} among W1,,WpW_{1},\dots,W_{p}, has two lifts Aj+A_{j}^{+} and AjA_{j}^{-} in W~i\tilde{W}_{i}. Now we identify the copy of AjA_{j} in C+C^{+} to Aj+A_{j}^{+} in W~i\tilde{W}_{i} and the one in CC^{-} to AjA_{j}^{-} in W~i\tilde{W}_{i} for each annulus among A1,AkA_{1},\dots A_{k}. We repeat the same procedure for every component CC of M(W1Wp)M\setminus(W_{1}\cup\dots\cup W_{p}), and get a manifold M~\tilde{M}, which will turn out to be a double cover of MM as shown below.

Define a homeomorphism t:M~M~t\colon\tilde{M}\to\tilde{M} to be the covering translation on each W~i\tilde{W}_{i} and the map taking C±C^{\pm} to CC^{\mp} preserving the identification with CC for each of C±C^{\pm}. It is clear from the definition that this homeomorphism tt is a free involution. By taking the quotient of M~\tilde{M} under t2\langle t\rangle\cong\mathbb{Z}_{2}, we get a manifold naturally identified with MM. Thus we see that M~\tilde{M} is a double cover of MM. Since W~i\tilde{W}_{i} is a product II-bundle and all of the characteristic II-pairs contained in C±C^{\pm} are product bundles by our definition of W1,,WpW_{1},\dots,W_{p}, we see that M~\tilde{M} is a double cover as desired. ∎

For some of our arguments we will need a stronger assumption than having only product bundles:

Definition 3.3.

Let MM be a compact orientable Haken 33-manifold with incompressible boundary. We say that MM is strongly untwisted if and only if:

  1. (A)

    Every characteristic II-pair is a product bundle.

  2. (B)

    For any characteristic II-pair Ξ\Xi and any simple closed curve dMd\subset\partial M, the simple closed curve dd can be homotoped on M\partial M into at most one component of ΞM\Xi\cap\partial M.

We are going to construct a cover with the properties (A) and (B) above. In order to do that, we need to examine how characteristic II-pairs are attached to other components of the characteristic submanifold. In the following proof of Lemma 3.4, it will turn out that there are two situations ((a) and (b) below) where the second condition of ‘strong untwistedness’ breaks down.

Lemma 3.4.

Let MM be a compact orientable atoroidal Haken manifold with incompressible boundary. Then there is a finite-sheeted regular covering of MM which is strongly untwisted.

Proof.

By Lemma 3.2, we have a double covering all of whose characteristic II-pairs are product II-bundles. Therefore, we may assume that MM satisfies the first condition (A) of ‘strong untiwistedness’, and we shall construct a covering satisfying the second condition.

To construct such a covering, let us analyse how this second condition (B) can fail to hold. Let dMd\subset\partial M be a simple closed curve, and let WW a characteristic II-pair. Since no components of WMW\cap\partial M are annuli, dd can be homotoped on M\partial M into at most two components of WMW\cap\partial M. Furthermore, if dd can be homotoped into two such components, then dd lies (up to isotopy on M\partial M) on a component TjT_{j} (characteristic solid torus or a characteristic thickened torus) of TT, and there are two possibilities: (a) TjMT_{j}\cap\partial M is an annulus when TjT_{j} is a solid torus, and is the union of an annulus and a torus when TjT_{j} is a thickened torus; or (b) dd separates two consecutive components of TjMT¯T_{j}\cap\overline{M\setminus T} both lying in the same characteristic II-pair. We shall show that we can take a finite-sheeted covering of MM so that neither (a) nor (b) can happen.

First, we consider the condition (a). Let TjT_{j} be a component of TT such that TjMT_{j}\cap\partial M is an annulus (and TjT_{j} is a solid torus) or the union of an annulus and a torus (when TjT_{j} is a thickened torus). This implies that TjMTj¯T_{j}\cap\overline{M\setminus T_{j}} is connected, hence is an annulus, which we denote by AA. Since TjT_{j} is a characteristic solid torus or characteristic thickened torus, the annulus AA is essential, and hence is not homotopic to TjMT_{j}\cap\partial M fixing the boundary. Then, we can choose a simple closed curve α\alpha, which is not contractible in MM, on the component of Tj\partial T_{j} on which dd lies so that both αA\alpha\cap A and αM\alpha\cap\partial M are connected, i.e. arcs. Since π1(Tj)\pi_{1}(T_{j}) is either \mathbb{Z} or ×\mathbb{Z}\times\mathbb{Z}, we can take a kk-sheeted cyclic covering T~j\tilde{T}_{j} of TjT_{j} so that α\alpha cannot be lifted homeomorphically, whereas the annulus AA is homeomorphically lifted. (For instance, in the case when π1(Tj)\pi_{1}(T_{j})\cong\mathbb{Z}, we choose kk which is coprime with the element represented by α\alpha.) Then the preimage of the annulus AA is kk copies of AA, which we denote by A1,,AkA_{1},\dots,A_{k}. Let CC be MTj¯\overline{M\setminus T_{j}}. We prepare kk copies of CC, which we denote by C1,,CkC_{1},\dots,C_{k}. By pasting CjC_{j} along AjA_{j} to T~j\tilde{T}_{j}, we can make a kk-sheeted cyclic covering of MM in which T~j\tilde{T}_{j} does not satisfy the condition (a). If there is another component TjT_{j^{\prime}} of TT with the condition (a), we repeat the same process for all the kk lifts of TjT_{j^{\prime}} at the same time. Repeating the process, we get a finite-sheeted covering of MM in which there is no characteristic solid torus or a characteristic thickened torus with the condition (a). We use the same symbol MM and TT for this finite-sheeted covering, abusing the notation.

Now we turn to the condition (b). We consider three colours named the colour 0, 11 and 22 and we choose a colour for each annulus of TMT¯T\cap\overline{M\setminus T} so that, on T\partial T, no two consecutive annuli have the same colour. We take three copies of each component of TT and of MT¯\overline{M\setminus T} which we name the lift 0, 11 and 22. Consider a component UU of TT, a component VV of MT¯\overline{M\setminus T} and an annulus EUVE\subset U\cap V with the colour k{0,1,2}k\in\{0,1,2\}. For every j{0,1,2}j\in\{0,1,2\}, we glue the lift jj of VV to the lift (j+k)(j+k) mod 33 of UU along the appropriate lifts of EE. Using the same construction for each component of TMT¯T\cap\overline{M\setminus T}, we get a triple cover M^\hat{M} of MM in which any two consecutive components of T^M^T^¯\hat{T}\cap\overline{\hat{M}\setminus\hat{T}} lie in different components of M^T^¯\overline{\hat{M}\setminus\hat{T}}. In particular there is no characteristic solid torus or characteristic thickened torus in M^\hat{M} for which the condition (b) holds.

Thus, we have shown that by taking a finite-sheeted covering, we can make both of the situations (a) and (b) disappear, which means, as we saw above, that the covering is strongly untwisted. ∎

Lemmas 3.1 and 3.4 show that to prove Theorem 1.1, we have only to consider the case when MM is strongly untwisted.

3.2. Vertically extendible surfaces

Let MM be an atoroidal Haken manifold as in Theorem 1.1. Let XX be the characteristic submanifold of MM. Assume that every II-bundle in XX is a product II-bundle.

Definition 3.5.

Given an incompressible subsurface FMF\subset\partial M, we say that FF is one-time vertically extendible if there is an incompressible surface F1MF^{1}\subset\partial M and an essential II-bundle VFMV_{F}\subset M with VFM=FF1V_{F}\cap\partial M=F\cup F^{1} and F1XF^{1}\subset\partial X up to isotopy. We call F1F^{1} a first elevation of FF.

It follows from the definition of characteristic submanifold that there is an isotopy which takes VFV_{F} into the characteristic submanifold XX. From now on, we assume that if FF is one-time vertically extendible then FXF\subset X and VFXV_{F}\subset X.

We note solid torus components in XX may add some complications in the case when FF is an annulus. If FF is contained in such a component of XX, there may be more than one possible first elevation (even up to isotopy) and the II-bundles corresponding to two disjoint annuli may intersect (even up to isotopy).

We now define multiple elevations by induction.

Definition 3.6.

Given an incompressible subsurface FF in M\partial M and n2n\geq 2, we say that FF is nn-time vertically extendible if there is an essential surface F1MF^{1}\subset\partial M and an essential II-bundle VFMV_{F}\subset M with VFM=FF1V_{F}\cap\partial M=F\cup F^{1} and ι(F1)\iota(F^{1}) is (n1)(n-1)-time vertically extendible. An (n1)(n-1)-th elevation FnF^{n} of ι(F1)\iota(F^{1}) is defined to be an nn-th elevation of FF.

We say that two multi-curves c,dMc,d\subset\partial M intersect minimally if for every multicurves c,dc^{\prime},d^{\prime} homotopic to cc and dd respectively, {cd}{cd}\sharp\{c\cap d\}\leq\sharp\{c^{\prime}\cap d^{\prime}\}. Let F,GXMF,G\subset X\cap\partial M be two incompressible surfaces. We say that FF and GG intersect minimally if F\partial F intersects G\partial G minimally.

Lemma 3.7.

Let F,GMF,G\subset\partial M be connected incompressible subsurfaces which intersect minimally and are not disjoint. If FF and GG are nn-time vertically extendible, then so is FGF\cup G.

Proof.

If FF and GG are one-time vertically extendible, as was remarked before, we may assume that F,GXMF,G\subset X\cap\partial M. Since they intersect minimally and are not disjoint, they must lie in the same component HH of XMX\cap\partial M which is not an annulus. Then the component VV of XX containing HH is an II-bundle, which is a product II-bundle by assumption, and can be parametrised as H×[0,1]H\times[0,1].

Then, by moving FF and GG by isotopies, we have VF=F×[0,1]H×[0,1]V_{F}=F\times[0,1]\subset H\times[0,1] and VG=G×[0,1]H×[0,1]V_{G}=G\times[0,1]\subset H\times[0,1], and F1=F×{0,1}F,G1=G×{0,1}GF^{1}=F\times\{0,1\}\setminus F,\ G^{1}=G\times\{0,1\}\setminus G. Since F1F^{1}, resp. G1G^{1}, lies in the component of XMX\cap\partial M which does not contain FF and GG, F1F^{1} and G1G^{1} lie in the same component of XMX\cap\partial M. Therefore F1G1F^{1}\cup G^{1} lies in XMX\cap\partial M. Thus we have proved that if FF and GG are one-time vertically extendible then FGF\cup G is also one-time vertically extendible and F1G1F_{1}\cup G_{1} is its first elevation.

The case of n>1n>1 follows by induction. ∎

Corollary 3.8.

For any natural number nn, there is an (possibly empty) incompressible surface ΣnXM\Sigma^{n}\subset X\cap\partial M such that each component of Σn\Sigma^{n} is nn-time vertically extendible, no component of Σn\Sigma^{n} can be isotoped on M\partial M into another component of Σn\Sigma^{n} and every nn-time vertically extendible surface can be isotoped into Σn\Sigma^{n}.

Proof.

If there is no surface that is nn-time vertically extendible, we set Σn\Sigma^{n} to be \emptyset. Otherwise, let ΣX\Sigma\subset X be an nn-time vertically extendible incompressible surface. If every nn-time vertically extendible surface can be isotoped into Σ\Sigma, we are done, by taking Σn=Σ\Sigma^{n}=\Sigma.

Otherwise, there is an nn-time vertically extendible surface FF which cannot be isotoped into Σ\Sigma. Moving FF by an isotopy we can assume that FF intersects Σ\Sigma minimally. By Lemma 3.7, each connected component of FΣF\cup\Sigma is nn-time vertically extendible, and we replace Σ\Sigma with ΣF\Sigma\cup F, and call this enlarged surface Σ\Sigma. We repeat this operation as long as there is an nn-time vertically extendible surface which cannot be isotoped into Σ\Sigma. Every time we add a surface, either we decrease the Euler characteristic of Σ\Sigma or we add a disjoint annulus which cannot be isotoped into Σ\Sigma. Hence this process must terminate after finitely many steps. The final resulting surface is Σn\Sigma^{n}. ∎

Since an nn-time vertically extendible surface is mm-vertically extendible for any mnm\leq n, we have ΣnΣm\Sigma^{n}\subset\Sigma^{m} up to isotopy.

In the next lemma we show that, when NN is atoroidal, MM cannot contain an nn-time extendible surface for sufficiently large nn. In the last section, this result will lead us to the constant nn of Theorem 1.1.

Lemma 3.9.

There is LL depending only on the topological type of M\partial M such that if there is an LL-time vertically extendible surface, then NN is not atoroidal.

Proof.

Letting gg denote the genus of M\partial M, we set K=3g3K=3g-3, which is the number of curves in a pants decomposition of M\partial M. Since no components of Σn\Sigma^{n} can be isotoped into another component, Σn\partial\Sigma^{n} has at most 2K2K boundary components. Using this observation, we show in the following claim that Σn+2K\Sigma^{n+2K} must be a proper subsurface of Σn\Sigma^{n} even up to isotopy.

Claim 3.10.

For any nn\in\mathbb{N}, if Σn\Sigma^{n} is non-empty and any component of Σn\Sigma^{n} can be isotoped into Σn+2K\Sigma^{n+2K}, then NN cannot be atoroidal.

Proof.

Suppose that Σn\Sigma^{n}\neq\emptyset, and that any component of Σn\Sigma^{n} can be isotoped into Σn+2K\Sigma^{n+2K}. Since Σn+j\Sigma^{n+j} is contained in Σn\Sigma^{n} for any j0j\geq 0 up to isotopy as observed above, and no component of Σn+j\Sigma^{n+j} can be isotoped into another component, we have then Σn+j=Σn\Sigma^{n+j}=\Sigma^{n} for any j2Kj\leq 2K up to isotopy. Let FF be a component of Σn+2K\Sigma^{n+2K} with minimal Euler characteristic, and FjF^{j} its jj-th elevation. By definition, ι(Fj)\iota(F^{j}) is (n+2Kj)(n+2K-j)-time vertically extendible for any j2Kj\leq 2K. Therefore ι(Fj)\iota(F^{j}) can be isotoped into Σn\Sigma^{n}. Since Σn=Σn+2K\Sigma^{n}=\Sigma^{n+2K} and FF has minimal Euler characteristic, ι(Fj)\iota(F^{j}) is a component of Σn\Sigma^{n}, up to isotopy. In particular (ι(Fj))Σn\partial(\iota(F^{j}))\subset\partial\Sigma^{n} up to isotopy.

Let VjV^{j} be the II-bundle cobounded by ι(Fj1)\iota(F^{j-1}) and FjF^{j}. We note that by definition, FjF^{j} and ι(Fj)\iota(F^{j}) are identified in NN and that the interior of VjV^{j} is embedded in NN. Taking the union of the II-bundles VjV^{j} in NN for j2Kj\leq 2K, we get a map F×[0,2K]NF\times[0,2K]\rightarrow N such that F×{j}F\times\{j\} is sent to FjF^{j}. Let cc be a component of F\partial F. The image of the annulus c×[0,2K]c\times[0,2K] goes 2K+12K+1 times through Σn\partial\Sigma^{n}. Since Σn\partial\Sigma^{n} has at most 2K2K components, there is a component cc^{\prime} of Σn\partial\Sigma^{n} through which c×[0,2K]c\times[0,2K] goes at least twice. The image of the part of this annulus between two such instances forms a torus TT embedded in NN. Considering the component of M\partial M through which TT goes, we can construct an infinite cyclic covering of NN in which TT lifts to an infinite incompressible annulus. It follows that TT is incompressible and non-peripheral. Hence NN is not atoroidal. ∎

As mentioned before, we have ΣnΣm\Sigma^{n}\subset\Sigma^{m} for any mnm\leq n. Consider monotone increasing indices njn_{j} such that Σnj+1\Sigma^{n_{j}+1} is smaller than Σnj\Sigma^{n_{j}} in the sense that at least one component of Σnj\Sigma^{n_{j}} cannot be isotoped into Σnj+1\Sigma^{n_{j}+1}. Since no component of Σn\Sigma^{n} can be isotoped into another component, we have then either χ(Σnj+1)>χ(Σnj)\chi(\Sigma^{n_{j}+1})>\chi(\Sigma^{n_{j}}) or Σnj+1\Sigma^{n_{j}+1} has fewer connected components than Σnj\Sigma^{n_{j}}. It follows that there are at most KK such njn_{j}, namely, there is JKJ\leq K such that for any nnJ+1n\geq n_{J}+1, we have Σn=Σn+1\Sigma^{n}=\Sigma^{n+1}. By Claim 3.10, if njnj12Kn_{j}-n_{j-1}\geq 2K for some jJj\leq J or if ΣnJ\Sigma^{n_{J}}\neq\emptyset, then NN is not atoroidal. Since JKJ\leq K, we can now conclude the proof just by setting L=2K2L=2K^{2}. ∎

4. Convergence, divergence and subsurface projections

In this section, we shall review the relations between the invariant mm mentioned in the introduction and the convergence and divergence of Fuchsian and Kleinian groups.

4.1. Subsurface projections and Fuchsian groups

We first recall the definition of the invariant mm from [2], and see how it controls the behaviour of sequences of Fuchsian groups.

Definition 4.1.

Let SS be a (possibly disconnected) closed surface of genus at least 22 and gg a point in its Teichmüller space. Regarding gg as a hyperbolic structure on SS, we let μ(g)\mu(g) be a shortest marking for (S,g)(S,g) (See Section 2.3). Although there might be more than one shortest markings, its choice does not matter for our definition and arguments. We fix a full and clean marking μ\mu consisting of a pants decomposition and transversals on SS independent of gg. For any essential simple closed curve dd on SS, we define

m(g,d,μ)=max{supY:dYdY(μ(g),μ),1lengthg(d)},m(g,d,\mu)=\max\left\{\sup_{\begin{subarray}{c}Y:\ d\subset\partial Y\end{subarray}}d_{Y}(\mu(g),\mu),{1\over\mathrm{length}_{g}(d)}\right\},

where the supremum of the first term in the maximum is taken over all incompressible subsurfaces YY of SS whose boundaries Y\partial Y contain dd.

It follows from [2, Lemma 5.2] that two curves with unbounded mm cannot intersect:

Lemma 4.2.

Let {mi}\{m_{i}\} be a sequence in 𝒯(S)\mathcal{T}(S) and let c1,c2c_{1},c_{2} be simple closed curve on SS. If m(mi,ci,μ)m(m_{i},c_{i},\mu)\longrightarrow\infty for both i=1,2i=1,2, then i(c1,c2)=0i(c_{1},c_{2})=0.

Proof.

This is just a special case of [2, Lemma 5.2] for Fuchisan groups. We note that the assumption of bounded projections of end invariants is unnecessary in this special case for which end invariants are empty. ∎

The invariant mm is related to the divergence and convergence of a sequence by the following lemma:

Lemma 4.3.

Let μ\mu be a full marking on SS, and let {mi}\{m_{i}\} be a sequence in 𝒯(S)\mathcal{T}(S). Then every subsequence of {mi}\{m_{i}\} contains a convergent subsequence if and only if {m(mi,c,μ)}\{m(m_{i},c,\mu)\} is bounded for every essential simple closed curve cc on SS.

Proof.

Let μi\mu_{i} be a shortest marking for mim_{i}. It follows from classical results on Fenchel-Nielsen coordinates that any subsequence of {mi}\{m_{i}\} contains a converging subsequence if and only if the sequence {μi}\{\mu_{i}\} is a finite set and {lengthmi(μi)}\{\mathrm{length}_{m_{i}}(\mu_{i})\} is bounded. By [2, Lemma 2.3], the sequence {μi}\{\mu_{i}\} is infinite if and only if passing to a subsequence, there is an incompressible subsurface YY such that dY(μi,μ)d_{Y}(\mu_{i},\mu)\longrightarrow\infty (and hence m(mi,c,μ)m(m_{i},c,\mu)\longrightarrow\infty for any component cc of Y\partial Y).

On the other hand, if the sequence {μi}\{\mu_{i}\} is constant, then lengthmi(μi)\mathrm{length}_{m_{i}}(\mu_{i}) is unbounded if and only if passing to a subsequence, there is a curve cc with lengthmi(c)0\mathrm{length}_{m_{i}}(c)\longrightarrow 0 (and hence m(mi,c,μ)m(m_{i},c,\mu)\longrightarrow\infty). ∎

4.2. Relative convergence of Kleinian groups

We shall next establish a necessary condition on the invariant mm for algebraic convergence on a submanifold. We start with a fundamental result. Thurston proved in [23] the following which is the first half of the theorem often referred to as the ‘broken window only’ theorem. We note that the latter half of the broken window only theorem should need some rectification (see [17]) but is irrelevant to the present paper.

Theorem 4.4.

Let MM be an atoroidal Haken 3-manifold and XX its characteristic submanifold. Then for any curve γ\gamma in MXM\setminus X and any sequence {ρi𝖠𝖧(M)}\{\rho_{i}\in\mathsf{AH}(M)\}, the length of the closed geodesic in 3/ρi(π1(M))\mathbb{H}^{3}/\rho_{i}(\pi_{1}(M)) representing the free homotopy class of ρi(γ)\rho_{i}(\gamma) is bounded as ii\longrightarrow\infty.

Using arguments from [2], we establish the following necessary condition for algebraic convergence on a submanifold.

Theorem 4.5.

Let MM be an atoroidal Haken boundary-irreducible 33-manifold all of whose characteristic II-pairs are product II-bundles. Let {mi}\{m_{i}\} be a sequence in 𝒯(M)\mathcal{T}(\partial M), and {ρi:π1(M)PSL2()}\{\rho_{i}:\pi_{1}(M)\to{\rm PSL}_{2}(\mathbb{C})\} a sequence of representations corresponding to {q(mi)}\{q(m_{i})\}. Let μM\mu\subset\partial M be a full and clean marking, and WMW\subset M a submanifold with ‘paring locus’ PP which is a union of disjoint non-parallel essential annuli on W\partial W. We assume the following:

  1. (a)

    The closure of WMW\setminus\partial M is a union of essential annuli contained in PP.

  2. (b)

    For any non-contractible simple closed curve cc in PP, lengthρi(c)\mathrm{length}_{\rho_{i}}(c) is bounded as ii\longrightarrow\infty.

  3. (c)

    For any essential annulus EWE\subset W disjoint from PP, there is a component cc of E\partial E such that {m(mi|S,c,μ)}\{m(m_{i}|S,c,\mu)\} is bounded for the component SS of M\partial M on which cc lies.

  4. (d)

    If MM is an II-bundle, then PP\neq\emptyset.

Then the sequence of the restrictions {ρi|π1(W)}\{\rho_{i}|\pi_{1}(W)\} has a convergent subsequence up to conjugation.

Proof.

We follow the argument of [2, Proposition 6.1] with some modifications as below. Denote by cic_{i} a shortest pants decomposition of M\partial M with respect to mim_{i}. Note that {dY(mi,μ)=dY(μ(mi),μ)}\{d_{Y}(m_{i},\mu)=d_{Y}(\mu(m_{i}),\mu)\} is bounded for any essential subsurface YY that is not an annulus with its core curve in cic_{i} if and only if so is {dY(ci,μ)}\{d_{Y}(c_{i},\mu)\}. Let r0r_{0} be a multicurve consisting of core curves of PP, one taken from each component of PP. We define r0r_{0} to be empty if PP is empty. Let XX be the characteristic submanifold of (W,P)(W,P). Consider a multicurve rr of (WX)W(W\setminus X)\cap\partial W containing r0r_{0} which is maximal in the sense that any simple closed curve in (WX)W(W\setminus X)\cap\partial W either intersects r0r_{0} or is homotopic on W\partial W to a component of r0r_{0}. By our assumption (b) and Theorem 4.4, there is LL such that lengthρi(r)L\mathrm{length}_{\rho_{i}}(r)\leq L.

Let ZZ be the union of the characteristic II-pairs in XX. By assumption, ZZ is a product II-bundle in the form Σ×I\Sigma\times I (Σ\Sigma may be disconnected). We denote by f:ZΣf\colon Z\to\Sigma the projection along the fibres, and for a subsurface FΣF\subset\Sigma and for j=0,1j=0,1, we use the symbol FjF_{j} to denote f1(F)Σ×{j}f^{-1}(F)\cap\Sigma\times\{j\}. For each component FF of Σ\Sigma that is not a pair of pants, by the assumption (c), there is j{0,1}j\in\{0,1\} such that {dFj(ci,μ)}\{d_{F_{j}}(c_{i},\mu)\} is bounded. Let SjS_{j} be the component of M\partial M containing FjF_{j}, and denote by θi=ρiI:π1(Sj)PSL2()\theta_{i}=\rho_{i}\circ I_{*}:\pi_{1}(S_{j})\to{\rm PSL}_{2}(\mathbb{C}) the representation induced by the inclusion I:SjMI\colon S_{j}\hookrightarrow M. The quotient manifold 3/θi(π1(Sj))\mathbb{H}^{3}/\theta_{i}(\pi_{1}(S_{j})) covers 3/ρi(π1(M))\mathbb{H}^{3}/\rho_{i}(\pi_{1}(M)), and has end invariant mi|Sjm_{i}|_{S_{j}} on one side. Now, replacing ρi\rho_{i} with θi\theta_{i}, we can follow the proof of [2, Lemma 6.2] starting at the penultimate paragraph. This gives us a constant LL^{\prime} and a sequence of curve {ai}\{a_{i}\} on FF such that ρi(ai)L\ell_{\rho_{i}}(a_{i})\leq L^{\prime} and {dFj(f1(ai)Fj,ci)}\{d_{F_{j}}(f^{-1}(a_{i})\cap F_{j},c_{i})\} is bounded.

Up to isotopy, f(rZ)f(r\cap Z) consists of boundary components of Σ\Sigma. We denote f(rZ)f(r\cap Z) by ss. Still following [2], if {ai}\{a_{i}\} has a constant subsequence, then we pass to an appropriate subsequence of {ρi}\{\rho_{i}\}, and add aia_{i} (independent of ii) to ss. If not, by [2, Lemma 2.3], there is a subsurface YFY\subset F with dY(ai,μ)d_{Y}(a_{i},\mu)\longrightarrow\infty, passing to a subsequence. Since {dFj(ci,μ)}\{d_{F_{j}}(c_{i},\mu)\} and {dFj(f1(ai)Fj,ci)}\{d_{F_{j}}(f^{-1}(a_{i})\cap F_{j},c_{i})\} are bounded, YY must be a proper subsurface of FF (even up to isotopy). If, passing to a subsequence, there is k{0,1}k\in\{0,1\} such that Yk=f1(Y)SkY_{k}=f^{-1}(Y)\cap S_{k} is an annulus containing a component of cic_{i} for all ii, we add the projection by ff of this component of cic_{i} to ss. Otherwise, by the assumption (c), there exists k{0,1}k\in\{0,1\} with bounded {dYk(ci,μ)}\{d_{Y_{k}}(c_{i},\mu)\}. Hence, passing to a subsequence, dYk(ci,f1(ai)Sk)d_{Y_{k}}(c_{i},f^{-1}(a_{i})\cap S_{k})\longrightarrow\infty, and by [15, Theorem B], ρn(Y)0\ell_{\rho_{n}}(\partial Y)\to 0. In this case, we add Y\partial Y to ss. We repeat the above construction letting FF be a component of Σs\Sigma\setminus s until Σs\Sigma\setminus s becomes a union of annuli and pair of pants. Adding f1(s)f^{-1}(s) to rr, we obtain a pants decomposition of W\partial W, which we shall still denote by rr, such that {ρn(r)}\{\ell_{\rho_{n}}(r)\} is bounded.

Next we attach a transversal with bounded length to each component of ss. Let ee be a curve in ss, by the assumption (c), there is j{0,1}j\in\{0,1\} such that {m(mi,ej,μ)}\{m(m_{i},e_{j},\mu)\} is bounded, where ej=f1(e)Sje_{j}=f^{-1}(e)\cap S_{j}. We replace cic_{i} with a shortest pants decomposition not containing eje_{j}. Since {m(mi,ej,μ)}\{m(m_{i},e_{j},\mu)\} is bounded, there is a positive lower bound on {mi(ej)\{\ell_{m_{i}}(e_{j})}, and there is an upper bound on {mi(ci)}\{\ell_{m_{i}}(c_{i})\} by our definition of cic_{i}. Considering the covering associated with the inclusion SjMS_{j}\hookrightarrow M we can use the arguments of [2] (proof of Proposition 6.1, the part after the proof of Lemma 6.2) to obtain a transversal tet_{e} to ee with bounded length ρi(te)\ell_{\rho_{i}}(t_{e}).

Since the union of ss and all transversals defined for the components of ss is doubly incompressible in Thurston’s sense [23, Section 2], we can deduce from Thurston’s relative boundedness theorem [23, Theorem 3.1] that the restriction of {ρi|π1(W)}\{\rho_{i}|_{\pi_{1}(W)}\} has a convergent subsequence. ∎

5. Unbounded skinning and annuli

The following proposition is the main step of our proof of Theorem 1.1.

Proposition 5.1.

Let MM be an orientable atoroidal boundary-irreducible Haken 33-manifold that is strongly untwisted. Let {mi}\{m_{i}\} be a sequence in 𝒯(M)\mathcal{T}(\partial M), let σ\sigma be the skinning map, and assume that there is a simple closed curve dd on M\partial M such that m(σ(mi),d,μ)m(\sigma(m_{i}),d,\mu)\longrightarrow\infty for a full clean marking μ\mu. Then, passing to a subsequence, there is a properly embedded essential annulus AMA\subset M with A=dd\partial A=d\cup d^{\prime} such that m(mi,d,μ)m(m_{i},d^{\prime},\mu)\longrightarrow\infty.

We are going to show that any subsequence of {mi}\{m_{i}\} contains a further subsequence for which the conclusion holds. To simplify the notations we shall use the same subscript ii for all subsequences.

5.1. Re-marking

Our manifold MM is either connected or has two components. In the case when MM has two components, by considering the component on which dd lies, and abusing the symbol MM to denote this component, we can assume that MM is connected. Recall that, by the assumption throughout this section, MM is strongly untwisted. Let ρi:π1(M)PSL2()\rho_{i}\colon\pi_{1}(M)\to{\rm PSL}_{2}(\mathbb{C}) be a representation corresponding to q(mi)q(m_{i}).

As a first step for the proof of Proposition 5.1, we change the markings of MM so that the behaviour of the ρi\rho_{i} can be read more easily from the behaviour of their end invariants.

Lemma 5.2.

Let dd be an essential simple closed curve on M\partial M, and let d1,,dpd_{1},\dots,d_{p} be disjoint simple closed curves on M\partial M representing the homotopy classes of simple closed curves on M\partial M homotopic to dd in MM, where d1=dd_{1}=d. Furthermore, we assume that

  1. (*)

    {m(mi,dj,μ)}\{m(m_{i},d_{j},\mu)\} is bounded for every j=2,,pj=2,\dots,p.

Then there is a sequence of orientation-preserving homeomorphisms {ψi:MM}\{\psi_{i}:M\to M\} such that, passing to a subsequence, the following hold:

  1. (1)

    For any essential simple closed curve cMc\subset\partial M, either {m(ψi(mi),c,μ)}\{m(\psi_{i*}(m_{i}),c,\mu)\} is bounded or m(ψi(mi),c,μ)m(\psi_{i*}(m_{i}),c,\mu)\longrightarrow\infty,.

  2. (2)

    If AMA\subset M is an essential annulus disjoint from all the djd_{j} such that m(ψi(mi),kA,μ)m(\psi_{i*}(m_{i}),\partial_{k}A,\mu)\longrightarrow\infty for both boundary components 1A\partial_{1}A and 2A\partial_{2}A of AA, then ρiψi1(A)0\ell_{\rho_{i}\circ\psi^{-1}_{i*}}(\partial A)\longrightarrow 0.

  3. (3)

    For every djd_{j} among d1,,dpd_{1},\dots,d_{p} defined above,

    1. (i)

      ψi(dj)=dj\psi_{i}(d_{j})=d_{j} for every ii and jj,

    2. (ii)

      for every j=1,,pj=1,\dots,p, {m(ψi(mi),dj,μ)}\{m(\psi_{i*}(m_{i}),d_{j},\mu)\} is bounded if and only if {m(m1,dj,μ)}\{m(m_{1},d_{j},\mu)\} is bounded, and

    3. (iii)

      {m(σψi(mi),dj,μ)}\{m(\sigma\circ\psi_{i*}(m_{i}),d_{j},\mu)\} is bounded if and only if {m(σ(mi),dj,μ)}\{m(\sigma(m_{i}),d_{j},\mu)\} is bounded.

Proof.

We shall first define the homeomorphisms ψi\psi_{i}, and then verify the desired properties. Let Ξ\Xi be a component of the characteristic submanifold XX of MdM\setminus d. Suppose first that Ξ\Xi is a solid torus. The components of ΞM¯\overline{\partial\Xi\setminus\partial M} are incompressible annuli. We define ψi\psi_{i} on soid-torus components Ξ\Xi of XX to be a composition of Dehn twists along these frontier annuli with the following properties:

  1. (a)

    If Ξ\Xi is a solid torus, then πF(μ(ψi(mi)))\pi_{F}(\mu(\psi_{i*}(m_{i}))) is constant with respect to ii for every component FF of ΞM\Xi\cap\partial M except for at most one.

By the assumption (*), passing to a subsequence, we need not compose Dehn twists along annuli of the frontier components of Ξ\Xi to achieve the condition (a) when ΞM\Xi\cap\partial M contains an annular neighbourhood of dd (up to isotopy), and hence ψi\psi_{i}, as defined for the moment, also satisfies the following:

  1. (b)

    For every j=1,,pj=1,\dots,p, we have ψi(dj)=dj\psi_{i}(d_{j})=d_{j} and πAj(μ(ψi(mi)))=πAj(μ(mi))\pi_{A_{j}}(\mu(\psi_{i*}(m_{i})))=\pi_{A_{j}}(\mu(m_{i})) for an annulus AjA_{j} on M\partial M whose core curve is djd_{j}.

If Ξ\Xi is not a solid torus, Ξ\Xi is a product Ξ=F×I\Xi=F\times I. (Recall that we have an assumption that every characteristic II-pair of MM is a product bundle. This implies that an II-pair in the characteristic submanifold XX of MdM\setminus d is also a product II-bundle.) Let F0F_{0} be a component of ΞM\Xi\cap\partial M which does not contain a curve homotopic on M\partial M to d1d_{1} (there is always such a component since MM is strongly untwisted). Since the curve complex of F0F_{0} has finitely many orbits under the action of the mapping class group of F0F_{0} (relative to F0\partial F_{0}), there is a sequence of orientation-preserving homeomorphisms gi:F0F0g_{i}:F_{0}\to F_{0} fixing F0\partial F_{0} such that, passing to a subsequence, πF0(μ(gi(mi)))\pi_{F_{0}}(\mu(g_{i*}(m_{i}))) is constant. We then define ψi\psi_{i} on Ξ\Xi by extending gig_{i} along the fibres, i.e. ψi(x,t)=(gi(x),t)\psi_{i}(x,t)=(g_{i}(x),t) for any (x,t)Ξ=F0×I(x,t)\in\Xi=F_{0}\times I.

Thus we have the following.

  1. (c)

    there are R>0R>0 and a component F0F_{0} of ΞM\Xi\cap\partial M not containing any curve homotopic on M\partial M to d1d_{1} such that dY(μ(ψi(mi)),μ)Rd_{Y}(\mu(\psi_{i*}(m_{i})),\mu)\leq R for any incompressible subsurface YF0Y\subset F_{0}.

We note that since Ξ\Xi is a component of the characteristic submanifold of MdM\setminus d, if Ξ\partial\Xi contains a curve djd_{j}, then it must be peripheral, and hence the action of ψi\psi_{i} on Ξ\Xi does not affect the property (b).

We repeat the construction above for all the components of the characteristic submanifold XX, and we extend the resulting homeomorphisms to a homeomorphism of MM which is isotopic to the identity on the complement of the characteristic submanifold.

We now verify the properties (1, 2, 3) for ψi\psi_{i} thus constructed.

The first property (1) can be obtained by passing to a subsequence for any sequence of homeomorphisms. Therefore, we are done with (1).

We next turn to proving the property (3). By the assumption (*), taking a subsequence, we may assume that πF(μ(mi))\pi_{F}(\mu(m_{i})) is constant whenever FF is an annulus containing a curve djd_{j} for j1j\neq 1. Wet first show the following claim.

Claim 5.3.

For every j=1,,pj=1,\dots,p and for any sequence of incompressible subsurfaces YiMY_{i}\subset\partial M with its boundary containing djd_{j} which are not a pair of pants, {dYi(μ,ψi(μ))}\{d_{Y_{i}}(\mu,\psi_{i}(\mu))\} is bounded.

Proof.

Fix j=1,,pj=1,\dots,p, and consider a sequence of incompressible subsurfaces YiMY_{i}\subset\partial M each of which contains djd_{j} in its boundary. If all of the YiY_{i} are annulli after passing to a subsequence, the conclusion follows from the property (b). From now on, taking a subsequence, we assume that none of the YiY_{i} are annuli.

Assume first that there is a simple closed curve cMc\subset\partial M intersecting YiY_{i} which lies outside the characteristic submanifold XX. Then by our construction of ψi\psi_{i}, we have ψi(c)=c\psi_{i}(c)=c, and hence

dYi(μ,ψi(μ))\displaystyle d_{Y_{i}}(\mu,\psi_{i}(\mu)) dYi(μ,c)+dYi(c,ψi(μ))\displaystyle\leq d_{Y_{i}}(\mu,c)+d_{Y_{i}}(c,\psi_{i}(\mu))
dYi(μ,c)+dYi(ψi(c),ψi(μ))\displaystyle\leq d_{Y_{i}}(\mu,c)+d_{Y_{i}}(\psi_{i}(c),\psi_{i}(\mu))
dYi(μ,c)+dψi1(Yi)(μ,c)\displaystyle\leq d_{Y_{i}}(\mu,c)+d_{\psi_{i}^{-1}(Y_{i})}(\mu,c)
4i(c,μ)+2,\displaystyle\leq 4i(c,\mu)+2,

where the last inequality is due to Masur–Minsky [11, Lemma 2.1]. Thus we are done in this case.

Otherwise, taking a subsequence, we may assume that YiY_{i} is contained in ΞiM\Xi_{i}\cap\partial M for a component Ξi\Xi_{i} of the characteristic submanifold XX. Taking a further subsequence, we may assume that Ξi=Ξ\Xi_{i}=\Xi does not depend on ii. Since YiY_{i} is not an annulus, Ξ\Xi is a product II-pair F×IF\times I. Let F0F_{0} be the component of ΞM\Xi\cap\partial M given by the property (c). Let us denote by YiY_{i}^{\prime} the projection of YiY_{i} to F0F_{0} along the fibres, (setting Yi=YiY_{i}^{\prime}=Y_{i} if YiF0Y_{i}\subset F_{0}). By our definition of d1,,dpd_{1},\dots,d_{p}, the boundary of YiY^{\prime}_{i} contains some dkd_{k} with k2k\geq 2. Then, {m(mi,dk,μ)}\{m(m_{i},d_{k},\mu)\} is bounded by the assumption (*), and ψi(dk)=dk\psi_{i}(d_{k})=d_{k} by the property (b). In particular {dYi(μ(ψi(mi)),ψi(μ))=dψi1(Yi)(μ(mi),μ)}\{d_{Y^{\prime}_{i}}(\mu(\psi_{i}(m_{i})),\psi_{i}(\mu))=d_{\psi_{i}^{-1}(Y_{i}^{\prime})}(\mu(m_{i}),\mu)\} is bounded. On the other hand, by the property (c), {dYi(μ(ψi(mi)),μ)}\{d_{Y_{i}^{\prime}}(\mu(\psi_{i*}(m_{i})),\mu)\} is bounded. Thus we see that {dYi(μ,ψi(μ))dYi(μ,μ(ψi(mi))+dYi(μ(ψi(mi),ψi(μ))}\{d_{Y^{\prime}_{i}}(\mu,\psi_{i}(\mu))\leq d_{Y^{\prime}_{i}}(\mu,\mu(\psi_{i*}(m_{i}))+d_{Y^{\prime}_{i}}(\mu(\psi_{i*}(m_{i}),\psi_{i}(\mu))\} is bounded. It follows from the construction of ψi\psi_{i} that dYi(μ,ψi(μ))=dYi(μ,ψi(μ))d_{Y_{i}}(\mu,\psi_{i}(\mu))=d_{Y^{\prime}_{i}}(\mu,\psi_{i}(\mu)), and hence {dYi(μ,ψi(μ))}\{d_{Y_{i}}(\mu,\psi_{i}(\mu))\} is also bounded. ∎

Now we can show that the sequence {ψi}\{\psi_{i}\} satisfies the property (3) by the condition (*) and the following claim.

Claim 5.4.

For any j=1,,pj=1,\dots,p, the sequence {m(ψi(mi),dj,μ)}\{m(\psi_{i*}(m_{i}),d_{j},\mu)\} is bounded if and only if {m(mi,dj,μ)}\{m(m_{i},d_{j},\mu)\} is bounded, and {m(σψi(mi),dj,μ)}\{m(\sigma\circ\psi_{i*}(m_{i}),d_{j},\mu)\} is bounded if and only if {m(σ(mi),dj,μ)}\{m(\sigma(m_{i}),d_{j},\mu)\} is bounded.

Proof.

Let {YiM}\{Y_{i}\subset\partial M\} be a sequence of incompressible subsurfaces with djYid_{j}\subset\partial Y_{i} which are not pairs of pants. Since dYi(mi,μ)=dψi(Yi)(μ(ψi(mi)),ψi(μ))d_{Y_{i}}(m_{i},\mu)=d_{\psi_{i}(Y_{i})}(\mu(\psi_{i*}(m_{i})),\psi_{i}(\mu)), the triangle inequalities

dYi(μ(mi),μ)dψi(Yi)(μ(ψi(mi)),μ)+dψi(Yi)(μ,ψi(μ)), and dψi(Yi)(μ(ψi(mi)),μ)dψi(Yi)(μ(ψi(mi)),ψi(μ))+dψi(Yi)(ψi(μ),μ)\begin{split}&d_{Y_{i}}(\mu(m_{i}),\mu)\leq d_{\psi_{i}(Y_{i})}(\mu(\psi_{i*}(m_{i})),\mu)+d_{\psi_{i}(Y_{i})}(\mu,\psi_{i}(\mu)),\text{ and }\\ &d_{\psi_{i}(Y_{i})}(\mu(\psi_{i*}(m_{i})),\mu)\leq d_{\psi_{i}(Y_{i})}(\mu(\psi_{i*}(m_{i})),\psi_{i}(\mu))+d_{\psi_{i}(Y_{i})}(\psi_{i}(\mu),\mu)\end{split}

lead to

dψi(Yi)(μ(ψi(mi)),μ)dψi(Yi)(μ,ψi(μ))dYi(μ(mi),μ)dψi(Yi)(μ(ψi(mi)),μ)+dψi(Yi)(μ,ψi(μ)).\begin{split}&d_{\psi_{i}(Y_{i})}(\mu(\psi_{i*}(m_{i})),\mu)-d_{\psi_{i}(Y_{i})}(\mu,\psi_{i}(\mu))\\ &\leq d_{Y_{i}}(\mu(m_{i}),\mu)\leq d_{\psi_{i}(Y_{i})}(\mu(\psi_{i*}(m_{i})),\mu)+d_{\psi_{i}(Y_{i})}(\mu,\psi_{i}(\mu)).\end{split}

Thus by applying Claim 5.3, we see that {dYi(μ(mi),μ)}\{d_{Y_{i}}(\mu(m_{i}),\mu)\} is bounded if and only if {dψi(Yi)(μ(ψi(mi)),μ)}\{d_{\psi_{i}(Y_{i})}(\mu(\psi_{i*}(m_{i})),\mu)\} is bounded.

Since ψi(dj)=dj\psi_{i}(d_{j})=d_{j} by the property (b), we also have lengthmi(dj)=lengthψi(mi)(dj)\mathrm{length}_{m_{i}}(d_{j})=\mathrm{length}_{\psi_{i*}(m_{i})}(d_{j}), and we conclude that {m(ψi(mi),dj,μ)}\{m(\psi_{i*}(m_{i}),d_{j},\mu)\} is bounded if and only if {m(mi,dj,μ)}\{m(m_{i},d_{j},\mu)\} is bounded.

Since σ\sigma commutes with ψi\psi_{i*}, the same argument shows that that {m(σψi(mi),d,μ)}\{m(\sigma\circ\psi_{i*}(m_{i}),d,\mu)\} is bounded if and only if {(σ(mi),d,μ)}\{(\sigma(m_{i}),d,\mu)\} is bounded. ∎

To conclude the proof of Lemma 5.2, it remains to establish the property (2). We restate the property as a claim.

Claim 5.5.

Let AMA\subset M be an essential annulus with its boundary components denoted by 1A\partial_{1}A and 2A\partial_{2}A. Suppose that m(ψi(mi),kA,μ)m(\psi_{i*}(m_{i}),\partial_{k}A,\mu)\longrightarrow\infty for both k=1k=1 and k=2k=2. Then lengthρiψi1(1A)0\mathrm{length}_{\rho_{i}\circ\psi^{-1}_{i*}}(\partial_{1}A)\longrightarrow 0.

Proof.

Let a1,,aqa_{1},\dots,a_{q} be homotopically distinct simple closed curves on M\partial M representing all the homotopy classes (in M\partial M) homotopic to 1A\partial_{1}A in MM. By renumbering them, we can assume ak=kAa_{k}=\partial_{k}A for k=1,2k=1,2. If lengthψi(mi)(ak)0\mathrm{length}_{\psi_{i*}(m_{i})}(a_{k})\longrightarrow 0 for some k=1,,qk=1,\dots,q, we are done.

To deal with the remaining case, we now assume that there is a positive constant ϵ\epsilon such that lengthψi(mi)(ak)ϵ\mathrm{length}_{\psi_{i*}(m_{i})}(a_{k})\geq\epsilon for every ii\in\mathbb{N} and k=1,,qk=1,\dots,q. Then, there are a constant LL and simple closed curves ck,ic_{k,i} for every ii\in\mathbb{N} and k=1,,qk=1,\dots,q such that ck,ic_{k,i} intersects aka_{k} essentially and lengthψi(mi)(ck,i)L\mathrm{length}_{\psi_{i*}(m_{i})}(c_{k,i})\leq L. There is also K1K_{1} such that dY(ck,i,μ(ψi(mi)))K1d_{Y}(c_{k,i},\mu(\psi_{i*}(m_{i})))\leq K_{1} for any j,ij,i and any incompressible subsurface YMY\subset\partial M intersecting ck,ic_{k,i} that is neither an annulus nor a pair of pants, since by definition, the length of μ(ψi(mi))\mu(\psi_{i*}(m_{i})) is also bounded from above by a constant.

Since m(ψi(mi),ak,μ)m(\psi_{i*}(m_{i}),a_{k},\mu)\longrightarrow\infty and ψi(mi)(ak)ϵ\ell_{\psi_{i*}(m_{i})}(a_{k})\geq\epsilon for k=1,2k=1,2, there are incompressible subsurfaces Yk,iY_{k,i} such that akYk,ia_{k}\subset\partial Y_{k,i} and dYk,i(μ(ψi(mi)),μ)d_{Y_{k,i}}(\mu(\psi_{i*}(m_{i})),\mu)\longrightarrow\infty for k=1,2k=1,2. If, passing to a subsequence, Y1,iY_{1,i} and Y2,iY_{2,i} are both annuli, then, up to homotopy, they lie on the boundary of the same component Ξ\Xi of the characteristic submanifold (which is, up to passing to a further subsequence independent of ii). However, the assumption that m(ψi(mi),ak,μ)m(\psi_{i*}(m_{i}),a_{k},\mu)\longrightarrow\infty contradicts (a) when Ξ\Xi is a solid torus, and (c) when Ξ\Xi is an II-pair. Therefore, we can assume that one of the Yk,i(k=1,2)Y_{k,i}(k=1,2), say Y1,iY_{1,i} is not an annulus.

Suppose now that Y1,iY_{1,i} is not eventually contained in the characteristic submanifold XX (up to homotopy), even after passing to a subsequence. By taking a subsequence, we can assume that none of the Y1,iY_{1,i} are contained in XX. Then, there is a simple closed curve cMc\subset\partial M disjoint from XX which intersects Y1,iY_{1,i} for all ii, by passing to a further subsequence. By Theorem 4.4 there is a constant LL such that lengthρiψi1(c)L\mathrm{length}_{\rho_{i}\circ\psi_{i*}^{-1}}(c)\leq L. Since dY1,i(μ(ψi(mi)),μ)d_{Y_{1,i}}(\mu(\psi_{i*}(m_{i})),\mu)\longrightarrow\infty by our assumption, we have dY1,i(c1,i,c)d_{Y_{1,i}}(c_{1,i},c)\longrightarrow\infty. Then it follows from [15, Theorem B] that lengthρiψi1(Y1,i)0\mathrm{length}_{\rho_{i}\circ\psi_{i*}^{-1}}(\partial Y_{1,i})\longrightarrow 0, and hence in particular, we have lengthρiψi1(1A)0\mathrm{length}_{\rho_{i}\circ\psi_{i*}^{-1}}(\partial_{1}A)\longrightarrow 0.

Next suppose that Y1,iY_{1,i} eventually lies in XX. Taking a subsequence, we can assume that all the surfaces Y1,iY_{1,i} lie in the same component Ξ\Xi of XX. Since Y1,iY_{1,i} is not an annulus, Ξ\Xi must be an II-bundle, which has a form of Ξ=F×I\Xi=F\times I. By (c), there is another surface Y3,iΞY_{3,i}\subset\partial\Xi such that Y1,iY_{1,i} and Y3,iY_{3,i} bound an II-bundle compatible with the II-bundle structure of Ξ\Xi, and are projected along the fibres of Ξ=F×I\Xi=F\times I to the same surface ZiZ_{i} in FF and dY3,i(μ(ψi(mi)),μ)Rd_{Y_{3,i}}(\mu(\psi_{i*}(m_{i})),\mu)\leq R. We note that by our definition of a1,,aqa_{1},\dots,a_{q}, there is k02k_{0}\geq 2 such that ak0a_{k_{0}} lies on Y3,i\partial Y_{3,i}. Then since dY3,i(μ(ψi(mi)),ck0,i)K1d_{Y_{3,i}}(\mu(\psi_{i*}(m_{i})),c_{k_{0},i})\leq K_{1}, we have dY3,i(ck0,i,μ)R+K1d_{Y_{3,i}}(c_{k_{0},i},\mu)\leq R+K_{1}. We shall make use of {c1,i}\{c_{1,i}\} and {ck0,i}\{c_{k_{0},i}\} to apply [15, Theorem B] as before. Since they do not lie on the same surface, we first need to project them to FF. This leads to the following claim:

Claim 5.6.

There are K>0K>0 and two sequences of simple closed curves {d1,i}\{d_{1,i}\} and {dk0,i}\{d_{k_{0},i}\} on FF such that lengthρiψi1(dk,i)K\mathrm{length}_{\rho_{i}\circ\psi_{i*}^{-1}}(d_{k,i})\leq K for all ii and k=1,k0k=1,k_{0}, and dZi(d1,i,dk0,i)d_{Z_{i}}(d_{1,i},d_{k_{0},i})\longrightarrow\infty.

Proof.

Let kk be either 11 or k0k_{0}. If ck,ic_{k,i} is contained in Ξ\Xi for sufficiently large ii, then we let dk,id_{k,i} be the projection of ck,ic_{k,i} to FF. We also note that lengthρiψi1(dk,i)L\mathrm{length}_{\rho_{i}\circ\psi_{i*}^{-1}}(d_{k,i})\leq L then.

Suppose that this is not the case. We let SS be the component of M\partial M containing ck,ic_{k,i}. Following [15, page 138] we extend the multicurve B:=Fr(ΞS)B:=\operatorname{Fr}(\Xi\cap S) to a complete geodesic lamination λ\lambda by performing Dehn twists around BB infinitely many times to ck,ic_{k,i} and adding finitely many isolated leaves spiralling around BB. There is a unique pleated surface hk,i:S3/ρi(π1(S))h_{k,i}\colon S\to\mathbb{H}^{3}/\rho_{i}(\pi_{1}(S)) realising λ\lambda which induces ρiψi1\rho_{i}\circ\psi_{i*}^{-1} between the fundamental groups. Let RλR_{\lambda} be the ϵ\epsilon-thick part of SS with respect to the hyperbolic metric induced by hk,ih_{k,i}. By the efficiency of pleated surfaces ([22, Theorem 3.3], [15, Theorem 3.5]), there is a constant K2K_{2} such that lengthhk,i(ck,iRλ)L+K2i(ck,i,B)\mathrm{length}_{h_{k,i}}(c_{k,i}\cap R_{\lambda})\leq L+K_{2}i(c_{k,i},B) (the relation between the alternation and intersection numbers comes from (4.3) in [15]). In particular, there is an arc κk,i\kappa_{k,i} in ck,i(ΞS)Rλc_{k,i}\cap(\Xi\cap S)\cap R_{\lambda} intersecting Yk,iY_{k,i} and having length at most L+K2L+K_{2}. By Theorem 4.4, the length of each component of BB on hk,ih_{k,i} is bounded by a constant LL^{\prime} independent of ii. By joining one or two copies of κk,i\kappa_{k,i} (depending on whether κk,i\kappa_{k,i} intersects one or two components of BFrRλB\cup\operatorname{Fr}R_{\lambda}) with arcs on BFrRλB\cup\operatorname{Fr}R_{\lambda}, we can construct in SΞS\cap\Xi a simple closed curve dk,id_{k,i} such that lengthhk,i(dk,i)2(L+K2+L+ϵ)\mathrm{length}_{h_{k,i}}(d_{k,i})\leq 2(L+K_{2}+L^{\prime}+\epsilon). Furthermore, this construction implies that there is a constant K3K_{3} such that dY(dk,i,ck,i)K3d_{Y}(d_{k,i},c_{k,i})\leq K_{3} for any incompressible subsurface YSΞY\subset S\cap\Xi intersecting both dk,id_{k,i} and ck,ic_{k,i}, and in particular for Y=Yk,iY=Y_{k,i}. We use the same symbol dk,id_{k,i} to denote the projection of dk,id_{k,i} on FF along the fibres of Ξ=F×I\Xi=F\times I.

Thus we have lengthρiψi1(dk,i)2(L+K2+L+ϵ),\mathrm{length}_{\rho_{i}\circ\psi_{i*}^{-1}}(d_{k,i})\leq 2(L+K_{2}+L^{\prime}+\epsilon), and dZi(d1,i,dk0,i)dY1,j(c1,i,μ)dYk0,i(ck0,i,μ)2K3dY1,j(c1,i,μ)RK12K3d_{Z_{i}}(d_{1,i},d_{k_{0},i})\geq d_{Y_{1,j}}(c_{1,i},\mu)-d_{Y_{k_{0},i}}(c_{k_{0},i},\mu)-2K_{3}\geq d_{Y_{1,j}}(c_{1,i},\mu)-R-K_{1}-2K_{3}\longrightarrow\infty. ∎

Continuation of Proof of 5.5. Set ϑi=ρiψi1I:π1(S)PSL2()\vartheta_{i}=\rho_{i}\circ\psi_{i*}^{-1}\circ I_{*}:\pi_{1}(S)\to{\rm PSL}_{2}(\mathbb{C}) where I:π1(S)π1(M)I_{*}:\pi_{1}(S)\to\pi_{1}(M) is the homomorphism induced by the inclusion. Following [15], we denote by 𝒞0(ϑi,K)\mathcal{C}_{0}(\vartheta_{i},K) the set of simple closed curves on SS whose translation lengths with respect to ϑi\vartheta_{i} are less than or equal to KK. By the claim above, we see that both d1,id_{1,i} and dk0,id_{k_{0},i} lie in 𝒞0(ϑi,K)\mathcal{C}_{0}(\vartheta_{i},K) and that dY1,i(d1,i,dk0,i)d_{Y_{1,i}}(d_{1,i},d_{k_{0},i})\longrightarrow\infty. In particular, diamY1,i(𝒞0(ϑi,K))\mathrm{diam}_{Y_{1,i}}({\mathcal{C}}_{0}(\vartheta_{i},K))\longrightarrow\infty. It follows from [15, Theorem B] that lengthϑi(Y1,i)0\mathrm{length}_{\vartheta_{i}}(\partial Y_{1,i})\longrightarrow 0. In particular, lengthϑi(1A)0\mathrm{length}_{\vartheta_{i}}(\partial_{1}A)\longrightarrow 0, and hence lengthρiψi1(1A)0\mathrm{length}_{\rho_{i}\circ\psi_{i*}^{-1}}(\partial_{1}A)\longrightarrow 0. ∎

This also concludes the proof of Lemma 5.2. ∎

By Claim 5.4, proving Proposition 5.1 for {ρi}\{\rho_{i}\} is equivalent to proving it for {ρiψi1}\{\rho_{i}\circ\psi_{i*}^{-1}\}. Thus we may assume that {ρi}\{\rho_{i}\} satisfies the following.

  1. (I)

    For any simple closed curve cMc\subset\partial M, either {m(mi,c,μ)}\{m(m_{i},c,\mu)\} (resp. {m(σ(mi),c,μ)}\{m(\sigma(m_{i}),c,\mu)\},) is bounded or m(mi,c,μ)m(m_{i},c,\mu)\longrightarrow\infty (resp. m(σ(mi),c,μ))m(\sigma(m_{i}),c,\mu)\longrightarrow\infty).

  2. (II)

    If AMA\subset M is an essential annulus such that m(mi,kA,μ)(k=1,2)m(m_{i},\partial_{k}A,\mu)\longrightarrow\infty\ (k=1,2) for both boundary components 1A\partial_{1}A and 2A\partial_{2}A of AA, then lengthρi(A1)0\mathrm{length}_{\rho_{i}}(\partial A_{1}^{*})\longrightarrow 0.

5.2. End invariants and wrapping

In this subsection, we shall discuss how algebraic limits projects to geometric limits and how this is reflected in the behaviour of the end invariants.

Let us now fix the assumptions and notations which will be used in most results of this section.

Setting 5.7.

We consider an orientable atoroidal compact boundary-irreducible Haken 33-manifold MM without torus boundary components, and a sequence of representations ρiQH(M)\rho_{i}\in QH(M) corresponding to Ahlfors-Bers coordinates mi𝒯(M)m_{i}\in\mathcal{T}(\partial M). We have a non-contractible simple closed curve dMd\subset\partial M, and we denote by d1,,dpMd_{1},...,d_{p}\subset\partial M simple closed curves representing all homotopy classes of M\partial M on M\partial M which are homotopic to dd in MM, with d=d1d=d_{1}. We assume that ρi(d)0\ell_{\rho_{i}}(d^{*})\longrightarrow 0.

We also assume that we have a submanifold VdV_{d} of MM whose frontier consists of incompressible annuli and which has the following three properties:

  1. (i)

    VdV_{d} contains all the curves dj(j=1,,p)d_{j}\ (j=1,\dots,p), and djd_{j} is not peripheral in VdMV_{d}\cap\partial M for every j=1,,pj=1,\dots,p.

  2. (ii)

    The restriction of ρi\rho_{i} to π1(Vd)\pi_{1}(V_{d}) converges to a representation ρ:π1(Vd)PSL2()\rho_{\infty}\colon\pi_{1}(V_{d})\to{\rm PSL}_{2}(\mathbb{C}).

  3. (iii)

    If AVdA\subset V_{d} is an essential annulus disjoint from dd with core curve aa which is not homotopic to dd in MM, then lengthρi(a)0\mathrm{length}_{\rho_{i}}(a)\longrightarrow 0 if and only if AA is properly homotopic to the closure of a component of VdM\partial V_{d}\setminus\partial M.

Suppose first p2p\geq 2. If a component of the characteristic submanifold containing dd (up to isotopy) is a solid torus, then it contains all of d1,,dpd_{1},\dots,d_{p} up to isotopy. We let TT be this characteristic solid torus in this case. If the component is an II-pair, then p=2p=2, and it contains d2d_{2} up to isotopy. In this case, we let TT be A×[0,1]A\times[0,1] such that A×{0}A\times\{0\} is an annular neighbourhood of dd whereas A×{1}A\times\{1\} is that of d2d_{2}. Since FrVd\operatorname{Fr}V_{d} consists of annuli, by the condition (i) above, TT can be assumed to be contained in VdV_{d} by moving it by an isotopy in both cases. If p=1p=1, we set T=T=\emptyset.

Given j=1,,pj=1,\dots,p, we denote by FjF_{j} the component of VdMkjdkV_{d}\cap\partial M\setminus\bigcup_{k\neq j}d_{k} containing djd_{j}.

The sequence of groups {ρi(π1(Vd))}\{\rho_{i}(\pi_{1}(V_{d}))\} converges geometrically to a Kleinian group Γ\Gamma containing ρ(π1(Vd))\rho_{\infty}(\pi_{1}(V_{d})), passing to a subsequence.

In the next section, we shall construct VdV_{d} having the properties above, which shows that our argument in the present section really works.

Assuming the existence of VdV_{d} for the moment, we now prove that every component of VdTV_{d}\setminus T has a compact core which is embedded in the geometric limit 3/Γ\mathbb{H}^{3}/\Gamma making use of the work of [4].

Lemma 5.8.

In 5.7, let WW be a submanifold of VdV_{d} which is the closure of a component of VdTV_{d}\setminus T. Then there is a relative compact core CW3/ρ(π1(W))C_{W}\subset\mathbb{H}^{3}/\rho_{\infty}(\pi_{1}(W)) which is homeomorphic to WW and on which the restriction of the covering projection 3/ρ(π1(W))3/Γ\mathbb{H}^{3}/\rho_{\infty}(\pi_{1}(W))\to\mathbb{H}^{3}/\Gamma induced by the inclusion is injective. Furthermore, for the closures of two components W1,W2W_{1},W_{2} of VdTV_{d}\setminus T (in the case when TT is non-empty and separates WW), the compact cores CW1C_{W_{1}} and CW2C_{W_{2}} can be taken so that their images in 3/Γ\mathbb{H}^{3}/\Gamma are disjoint.

Proof.

Our conditions in 5.7 imply the assumptions of [4, Proposition 4.4], and applying this proposition, we see that there is a compact submanifold of 3/Γ\mathbb{H}^{3}/\Gamma which lifts to a compact core CWC_{W} of 3/ρ(π1(W))\mathbb{H}^{3}/\rho_{\infty}(\pi_{1}(W)) such that the restriction of the covering projection 3/ρ(π1(W))3/Γ\mathbb{H}^{3}/\rho_{\infty}(\pi_{1}(W))\to\mathbb{H}^{3}/\Gamma to CWC_{W} is injective. Let ΓWΓ\Gamma_{W}\subset\Gamma be the geometric limit of {ρi(π1(W))}\{\rho_{i}(\pi_{1}(W))\}. Then the restriction of the covering projection 3/ρ(π1(W))3/ΓW\mathbb{H}^{3}/\rho_{\infty}(\pi_{1}(W))\to\mathbb{H}^{3}/\Gamma_{W} to CWC_{W} must also be injective.

By [4, Lemma 4.6], ρ(π1(W))\rho_{\infty}(\pi_{1}(W)) is either a generalised web group or a degenerate group without accidental parabolic elements. It follows then from [1, Corollary C and Theorem E] that CWC_{W} is homeomorphic to WW. The last sentence of our lemma also follows from [4, Proposition 4.4]. ∎

We next show that by performing Dehn twists along embedded annuli bounded by dd and dj(j=2,,p)d_{j}\ (j=2,\dots,p), we can make each FjF_{j} embedded in the algebraic limit and mapped injectively in the geometric limit by the covering projection.

In the next lemma and the following, we shall use the expression ‘the outward side of a cusp’. We say that an embedding of the surface FjVdF_{j}\subset\partial V_{d} into the geometric limit 3/Γ\mathbb{H}^{3}/\Gamma lies on the outward side of a cusp if the cusp lies on the same side of the embedding of FjF_{j} as the embeddings of the components of VdTV_{d}\setminus T intersecting FjF_{j}. Otherwise we say that the embedding of FjF_{j} lies on the inward side of the cusp.

Lemma 5.9.

In 5.7, we denote by DjD_{j} the right-hand Dehn twist along an embedded annulus bounded by d=d1d=d_{1} and djd_{j} (j=2,,pj=2,\dots,p). Then for each jj, there is a sequence {ai(j)}\{a_{i}(j)\} of integers with the following properties:

  1. The sequence {θi=ρiDjai(j)|π1(Fj)}\{\theta_{i}=\rho_{i}\circ{D_{j*}^{a_{i}(j)}}|_{\pi_{1}(F_{j})}\} converges algebraically to a representation θ:π1(Fj)PSL2()\theta_{\infty}:\pi_{1}(F_{j})\to{\rm PSL}_{2}(\mathbb{C}).

  2. There is an embedding hj:Fj3/θ(π1(Fj))h_{j}:F_{j}\to\mathbb{H}^{3}/\theta_{\infty}(\pi_{1}(F_{j})) inducing θ\theta_{\infty} such that the restriction of the covering projection ΠFj:3/θ(π1(Fj))3/Γ\Pi_{F_{j}}:\mathbb{H}^{3}/\theta_{\infty}(\pi_{1}(F_{j}))\to\mathbb{H}^{3}/\Gamma to hj(Fj)h_{j}(F_{j}) is an embedding and its image ΠFjhj(Fj)\Pi_{F_{j}}\circ h_{j}(F_{j}) lies on the outward side of the cusp corresponding to ρ(d)=θ(d)\rho_{\infty}(d)=\theta_{\infty}(d) when the latter is a rank-2 cusp.

Proof.

This is a relative version of [2, Lemma 4.5].

Let WW^{\prime} and W′′W^{\prime\prime} be the components of VdTV_{d}\setminus T intersecting FjF_{j} (we set W=W′′W^{\prime}=W^{\prime\prime} if there is only one such component), and set Fj=FjWF_{j}^{\prime}=F_{j}\cap W^{\prime} and Fj′′=FjW′′F_{j}^{\prime\prime}=F_{j}\cap W^{\prime\prime}. By Lemma 5.8, there are compact cores CW3/ρ(π1(W))C_{W^{\prime}}\subset\mathbb{H}^{3}/\rho_{\infty}(\pi_{1}(W^{\prime})) and CW′′3/ρ(π1(W′′))C_{W^{\prime\prime}}\subset\mathbb{H}^{3}/\rho_{\infty}(\pi_{1}(W^{\prime\prime})), homeomorphic to WW^{\prime} and W′′W^{\prime\prime} respectively, on which the restrictions of the covering projections to 3/Γ\mathbb{H}^{3}/\Gamma are injective. The inclusions induce embeddings f:FjCWf^{\prime}:F_{j}^{\prime}\hookrightarrow\partial C_{W^{\prime}} and f′′:Fj′′CW′′f^{\prime\prime}:F_{j}^{\prime\prime}\hookrightarrow\partial C_{W^{\prime\prime}} which lift to embeddings g:Fj3/ρ(π1(Fj))g^{\prime}:F_{j}^{\prime}\hookrightarrow\mathbb{H}^{3}/\rho_{\infty}(\pi_{1}(F_{j})) and g′′:Fj′′3/ρ(π1(Fj))g^{\prime\prime}:F_{j}^{\prime\prime}\hookrightarrow\mathbb{H}^{3}/\rho_{\infty}(\pi_{1}(F_{j})). The restrictions of the covering projection ΠFj:3/ρ(π1(Fj))3/Γ\Pi_{F_{j}}:\mathbb{H}^{3}/\rho_{\infty}(\pi_{1}(F_{j}))\to\mathbb{H}^{3}/\Gamma to g(FjW)g^{\prime}(F_{j}\cap W^{\prime}) and to g′′(FjW′′)g^{\prime\prime}(F_{j}\cap W^{\prime\prime}) are embeddings.

If TT does not separate FjF_{j}, we set gˇ=g=g′′\check{g}=g^{\prime}=g^{\prime\prime}, otherwise, we put gg^{\prime} and g′′g^{\prime\prime} together to get an embedding gˇ:FjT3/ρ(π1(Fj))\check{g}:F_{j}\setminus T\to\mathbb{H}^{3}/\rho_{\infty}(\pi_{1}(F_{j})). Moving CW,CW′′C_{W^{\prime}},C_{W^{\prime\prime}}, ff^{\prime} and f′′f^{\prime\prime} by isotopies, we may assume that they send the boundary of FjTF_{j}\setminus T into the ϵ\epsilon-thin part. Then for an appropriate choice of ϵ\epsilon, the map gˇ\check{g} sends the boundary of FjTF_{j}\setminus T to the boundary of the ϵ1\epsilon_{1}-thin part of 3/ρ(π1(Fj))\mathbb{H}^{3}/\rho_{\infty}(\pi_{1}(F_{j})), where ϵ1\epsilon_{1} is smaller than the three-dimensional Margulis constant. It is then easy to extend gˇ\check{g} to an embedding g:Fj3/ρ(π1(Fj))g:F_{j}\to\mathbb{H}^{3}/\rho_{\infty}(\pi_{1}(F_{j})) such that g(TFj)g(T\cap F_{j}) lies on the boundary of the ϵ2\epsilon_{2}-thin part with ϵ2ϵ1\epsilon_{2}\leq\epsilon_{1}. By Lemma 5.8 and by our construction, the restriction of ΠFjg\Pi_{F_{j}}\circ g to FjTF_{j}\setminus T, which is ΠFjg^\Pi_{F_{j}}\circ\hat{g}, is an embedding and with an appropriate choice of ϵ\epsilon, the composition ΠFjg\Pi_{F_{j}}\circ g maps FjTF_{j}\cap T to the boundary of the ϵ0\epsilon_{0}-thin part of 3/Γ\mathbb{H}^{3}/\Gamma.

If ρ(d)\rho_{\infty}(d) belongs to a rank-11 maximal parabolic subgroup of Γ\Gamma, then it is easy to change gg on FjTF_{j}\cap T so that ΠFjg\Pi_{F_{j}}\circ g is an embedding. In this case, we simply take aia_{i} to be 0.

Otherwise, ρ(d)\rho_{\infty}(d) belongs to a rank-22 maximal parabolic subgroup of Γ\Gamma. We denote by T0T_{0} the boundary of the corresponding torus cusp-neighbourhood in 3/Γ\mathbb{H}^{3}/\Gamma, i.e. the boundary of the corresponding component of the ϵ2\epsilon_{2}-thin part. Let ZZ be the union of ΠFjg(FjT)\Pi_{F_{j}}\circ g(F_{j}\setminus T) and T0T_{0}. Then ΠFjg(Fj)\Pi_{F_{j}}\circ g(F_{j}) is contained in ZZ by our way of extending gˇ\check{g} to gg as described above. As is explained in [2, Lemma 3.1], ΠFjg\Pi_{F_{j}}\circ g is homotopic to a standard map fkf_{k} wrapping kk times around T0T_{0} for some kk\in\mathbb{Z}, and there are two standard embeddings f0,f1:FjZf_{0},f_{1}:F_{j}\to Z such that f0(Fj)f_{0}(F_{j}) lies on the outward side of the cusp associated with dd and f1(Fj)f_{1}(F_{j}) lies on its inward side, both without wrapping around T0T_{0}.

Let {qi:Bri(3/ρi(π1(M)),xi)BKiri(3/Γ,x)}\{q_{i}\colon B_{r_{i}}(\mathbb{H}^{3}/\rho_{i}(\pi_{1}(M)),x_{i})\to B_{K_{i}r_{i}}(\mathbb{H}^{3}/\Gamma,x_{\infty})\} be a sequence of KiK_{i}-bi-Lipschitz approximate isometry on the rir_{i}-ball with ri,Ki1r_{i}\longrightarrow\infty,K_{i}\longrightarrow 1 given by the geometric convergence as explained in Section 2.4. By [2, Lemma 3.1], there is sis_{i}\in\mathbb{Z} such that qi1f0q_{i}^{-1}\circ f_{0} is homotopic to qi1ΠFigDjsiq_{i}^{-1}\circ\Pi_{F_{i}}\circ g\circ D_{j}^{s_{i}}. The conclusion follows, taking ai(j)=sia_{i}(j)=s_{i} and setting hjh_{j} to be the lift of f0f_{0} to 3/θ(π1(Fj))\mathbb{H}^{3}/\theta_{\infty}(\pi_{1}(F_{j})). ∎

Next we study how the embedding of a compact core in the geometric limit as above affects the end invariants.

Lemma 5.10.

In 5.7, for each j=1,,pj=1,\dots,p, suppose that there is an embedding hj:Fj3/ρ(π1(Fj))h_{j}:F_{j}\to\mathbb{H}^{3}/\rho_{\infty}(\pi_{1}(F_{j})) inducing ρ|π1(Fj)\rho_{\infty}|_{\pi_{1}(F_{j})} such that the restriction of the covering projection ΠFj:3/ρ(π1(Fj))3/Γ\Pi_{F_{j}}:\mathbb{H}^{3}/\rho_{\infty}(\pi_{1}(F_{j}))\to\mathbb{H}^{3}/\Gamma to hj(Fj)h_{j}(F_{j}) is an embedding.

If ΠFj(hj(Fj))\Pi_{F_{j}}(h_{j}(F_{j})) lies on the outward side of the cusp associated with ρ(d)ρ(π1(M))Γ\rho_{\infty}(d)\in\rho_{\infty}(\pi_{1}(M))\subset\Gamma, then {m(mi,dj,μ)}\{m(m_{i},d_{j},\mu)\} is bounded whereas m(σ(mi),dj,μ)m(\sigma(m_{i}),d_{j},\mu)\longrightarrow\infty. If ΠFj(hj(Fj))\Pi_{F_{j}}(h_{j}(F_{j})) lies on the inward side of the cusp associated with ρ(d)\rho_{\infty}(d) then {m(σ(mi),dj,μ)}\{m(\sigma(m_{i}),d_{j},\mu)\} is bounded whereas m(mi,dj,μ)m(m_{i},d_{j},\mu)\longrightarrow\infty.

Proof.

Suppose that ΠFj(hj(Fj))\Pi_{F_{j}}(h_{j}(F_{j})) lies on the outward side of the cusp associated with ρ(d)ρ(π1(M))Γ\rho_{\infty}(d)\in\rho_{\infty}(\pi_{1}(M))\subset\Gamma. Let cFjc\subset F_{j} be a simple closed curve intersecting djd_{j} essentially, cc^{*} the closed geodesic homotopic to ΠFj(hj(c))\Pi_{F_{j}}(h_{j}(c)), and denote by ψi:BKiri(3/ρi(π1(M)),xi)Bri(3/Γ,x)\psi_{i}\colon B_{K_{i}r_{i}}(\mathbb{H}^{3}/\rho_{i}(\pi_{1}(M)),x_{i})\to B_{r_{i}}(\mathbb{H}^{3}/\Gamma,x_{\infty}) an approximate isometry associated with the geometric convergence of {ρi(π1(M))}\{\rho_{i}(\pi_{1}(M))\} to Γ\Gamma as explained in Section 2.4. For ii large enough, ψi1(c)\psi_{i}^{-1}(c^{*}) is a quasi-geodesic lying outside the thin part. on the same side as FjF_{j} of the Margulis tube associated with ρi(d)\rho_{i}(d). Let SjS_{j} be the component of M\partial M containing FjF_{j}. In the covering 3/ρi(π1(Sj))\mathbb{H}^{3}/\rho_{i}(\pi_{1}(S_{j})) of 3/ρi(π1(M))\mathbb{H}^{3}/\rho_{i}(\pi_{1}(M)), the closed geodesic homotopic to ρi(c)\rho_{i}(c) lies above the Margulis tube associated with ρi(d)\rho_{i}(d). Therefore, by [3, Theorem 1.3] there is a constant DD such that dY(c,μ(mi))Dd_{Y}(c,\mu(m_{i}))\leq D for any surface YSjY\subset S_{j} with djFrYd_{j}\subset\operatorname{Fr}Y. Thus for any full marking μ\mu, there is DD^{\prime} such that dY(μ,μ(mi))Dd_{Y}(\mu,\mu(m_{i}))\leq D^{\prime} for any surface YSY\subset S with djFrYd_{j}\subset\operatorname{Fr}Y.

To conclude that {m(mi,dj,μ)}\{m(m_{i},d_{j},\mu)\} is bounded, it remains to show that lengthmi(dj)\mathrm{length}_{m_{i}}(d_{j}) is bounded away from 0. Assume the contrary, that lengthmi(dj)0\mathrm{length}_{m_{i}}(d_{j})\longrightarrow 0 after passing to a subsequence. Then, there is an annulus joining the closed geodesic dj3/ρi(π1(S))d^{*}_{j}\subset\mathbb{H}^{3}/\rho_{i}(\pi_{1}(S)) representing ρi(d)\rho_{i}(d) with dj+C(3/ρi(π1(S)))d_{j}^{+}\subset\partial C(\mathbb{H}^{3}/\rho_{i}(\pi_{1}(S))) corresponding to djd_{j}, which lies entirely in the ϵi\epsilon_{i}-thin part with ϵi0\epsilon_{i}\longrightarrow 0. Since ψi1(c)\psi_{i}^{-1}(c^{*}) has bounded length, it cannot intersect such an annulus, whereas ψi1(c)\psi_{i}^{-1}(c^{*}) lies in a uniformly bounded neighbourhood of the convex core for large ii. Since cc^{*} and ΠFj(hj(Fj))\Pi_{F_{j}}(h_{j}(F_{j})) lie on the same side of the cusp associated with ρ(d)\rho_{\infty}(d), this contradicts the assumption that ΠFj(hj(Fj))\Pi_{F_{j}}(h_{j}(F_{j})) lies on the outward side of the cusp associated with ρ(d)Γ\rho_{\infty}(d)\in\Gamma.

Since lengthρi(d)0\mathrm{length}_{\rho_{i}}(d)\longrightarrow 0 and {m(mi,dj,μ)}\{m(m_{i},d_{j},\mu)\} is bounded, it follows from [14, Short Curve Theorem] that m(σ(mi),dj,μ)m(\sigma(m_{i}),d_{j},\mu)\longrightarrow\infty.

A quite similar argument works also when ΠFj(hj(Fj))\Pi_{F_{j}}(h_{j}(F_{j})) lies on the inward side of the cusp associated with ρ(d)Γ\rho_{\infty}(d)\in\Gamma. ∎

Corollary 5.11.

In 5.7, assume that p2p\geq 2, and consider jpj\leq p such that {m(mi,dj,μ)}\{m(m_{i},d_{j},\mu)\} is bounded. Then there is an embedding h:Fj3/ρ(π1(Fj))h:F_{j}\to\mathbb{H}^{3}/\rho_{\infty}(\pi_{1}(F_{j})) inducing ρ\rho_{\infty} such that the restriction of the covering projection ΠFj:3/ρ(π1(Fj))3/Γ\Pi_{F_{j}}:\mathbb{H}^{3}/\rho_{\infty}(\pi_{1}(F_{j}))\to\mathbb{H}^{3}/\Gamma to h(Fj)h(F_{j}) is an embedding whose image lies on the outward side of the cusp corresponding to ρ(d)\rho_{\infty}(d).

Proof.

As can be seen in the proof of Lemma 5.9, if ρ(d)\rho_{\infty}(d) belongs to a rank-11 maximal parabolic subgroup of Γ\Gamma, then ai(j)=0a_{i}(j)=0 for any ii and θ=ρ\theta_{\infty}=\rho_{\infty}. Therefore, our claim of this corollary follows immediately from Lemmas 5.9 and 5.10.

Otherwise, ρ(d)\rho_{\infty}(d) belongs to a rank-22 maximal parabolic subgroup of Γ\Gamma. By Lemma 5.9, there is a sequence of integers {ai(j)}\{a_{i}(j)\} and an embedding hj:Fj3/θ(π1(Fj))h_{j}\colon F_{j}\to\mathbb{H}^{3}/\theta_{\infty}(\pi_{1}(F_{j})) inducing θ\theta_{\infty} between the fundamental groups such that the restriction of the covering projection ΠFj:3/θ(π1(Fj))3/Γ\Pi_{F_{j}}:\mathbb{H}^{3}/\theta_{\infty}(\pi_{1}(F_{j}))\to\mathbb{H}^{3}/\Gamma to hj(Fj)h_{j}(F_{j}) is an embedding and its image ΠFjhj(Fj)\Pi_{F_{j}}\circ h_{j}(F_{j}) lies on the outward side of the cusp corresponding to θ(d)=ρ(d)\theta_{\infty}(d)=\rho_{\infty}(d). By Lemma 5.10, {m(Djaimi,dj,μ)}\{m(D_{j*}^{a_{i}}m_{i},d_{j},\mu)\} is bounded. Since {m(mi,dj,μ)}\{m(m_{i},d_{j},\mu)\} is bounded by assumption, this is possible only when {ai(j)}\{a_{i}(j)\} is bounded. Then we may take ai(j)=0a_{i}(j)=0 for any ii in Lemma 5.9 so that θ=ρ\theta_{\infty}=\rho_{\infty}, and the conclusion follows. ∎

We now put these results together to get the result which we shall use to prove Proposition 5.1.

Lemma 5.12.

In 5.7, suppose that {m(mi,dj,μ)}\{m(m_{i},d_{j},\mu)\} is bounded for j1j\neq 1. Then, there is a relative compact core for 3/ρ(π1(Vd))\mathbb{H}^{3}/\rho_{\infty}(\pi_{1}(V_{d})) homeomorphic to VdV_{d} on which the restriction of the covering projection Πd:3/ρ(π1(Vd))3/Γ\Pi_{d}:\mathbb{H}^{3}/\rho_{\infty}(\pi_{1}(V_{d}))\to\mathbb{H}^{3}/\Gamma is injective. Furthermore, a cusp neighbourhood corresponding to ρ(d)\rho_{\infty}(d) intersects the compact core in an annular neighbourhood of d1d_{1}.

Proof.

By Lemma 5.8, for the components WW of VdTV_{d}\setminus T, we have embeddings gW:W3/ρ(π1(Vd))g_{W}:W\to\mathbb{H}^{3}/\rho_{\infty}(\pi_{1}(V_{d})) inducing ρ|π1(W)\rho_{\infty}|_{\pi_{1}(W)}, on the union of which the restriction of Πd\Pi_{d} is injective.

If p=1p=1, then T=T=\emptyset by definition, and hence W=VdW=V_{d}. We can take a cusp neighbourhood corresponding to ρ(d)\rho_{\infty}(d) intersecting gW(W)g_{W}(W) along an annulus in the homotopy class of dd. Since p=1p=1, such an annulus is isotopic on Vd\partial V_{d} to an annular neighbourhood of d1=dd_{1}=d.

Suppose that p2p\geq 2, and assume that {m(mi,dj,μ)}\{m(m_{i},d_{j},\mu)\} is bounded for every j1j\neq 1. Then by Corollary 5.11, for every j1j\neq 1, there is an embedding gj:Fj3/ρ(π1(Vd))g_{j}:F_{j}\to\mathbb{H}^{3}/\rho_{\infty}(\pi_{1}(V_{d})) inducing ρ|π1(Fj)\rho_{\infty}|_{\pi_{1}(F_{j})} on which the restriction of Πd\Pi_{d} is injective. Furthermore, it follows from the construction that gjg_{j} and gWg_{W} agree on FjWF_{j}\cap W. Putting together the maps gWg_{W} for all the components WW of VdTV_{d}\setminus T and the gjg_{j} for all j1j\neq 1, we get an embedding g:Vd3/ρ(π1(Vd))g:V_{d}\to\mathbb{H}^{3}/\rho_{\infty}(\pi_{1}(V_{d})) inducing ρ|π1(Vd)\rho_{\infty}|_{\pi_{1}(V_{d})} on which the restriction of Πd\Pi_{d} is injective.

Changing gg by an isotopy, we may assume that g(Vd)g(V_{d}) intersects a cusp neighbourhood CC associated with ρ(d)\rho_{\infty}(d) along an annulus Ag(Vd)A\subset g(\partial V_{d}) which is a regular neighbourhood of g(dk)g(d_{k}) for some k=1,,pk=1,\dots,p. Then g(Fk)g(F_{k}) lies on the inward side of CC. This is possible only if Πdg(Fk)\Pi_{d}\circ g(F_{k}) lies on the inward side of CC; for the restriction of Πd\Pi_{d} is injective on g(Vd)g(V_{d}), and hence it cannot wrap around CC.

By assumption, for every j1j\neq 1, {m(mi,dj,μ)}\{m(m_{i},d_{j},\mu)\} is bounded. It follows then from Corollary 5.11 that Πd(g(Fj))\Pi_{d}(g(F_{j})) lies on the outward side of CC for j1j\neq 1. Hence the only possibility is that AA is a regular neighbourhood of g(d1)g(d_{1}). ∎

5.3. Completion of the proof of Proposition 5.1

Proof of Proposition 5.1.

If MM is an II-bundle, then m(σ(mi),d,μ)=m(mi,d2,μ)m(\sigma(m_{i}),d,\mu)=m(m_{i},d_{2},\mu) and the conclusion follows. In the other cases, we shall prove the proposition by contradiction. Assume that MM is not an II-bundle, that m(σ(mi),d,μ)m(\sigma(m_{i}),d,\mu)\longrightarrow\infty, and that {m(mi,dj,μ)}\{m(m_{i},d_{j},\mu)\} is bounded for every j=2,,pj=2,\dots,p.

By Lemma 5.2, after re-marking and passing to a subsequence, we may assume that {ρi=q(mi)}\{\rho_{i}=q(m_{i})\} satisfies:

  1. (1)

    for any simple closed curve cMc\subset\partial M, either {m(mi,c,μ)}\{m(m_{i},c,\mu)\} (resp. {m(σ(mi),c,μ)}\{m(\sigma(m_{i}),c,\mu)\}) is bounded or m(mi,c,μ)m(m_{i},c,\mu)\longrightarrow\infty (resp. m(σ(mi),c,μ)m(\sigma(m_{i}),c,\mu)\longrightarrow\infty);

  2. (2)

    if AMA\subset M is an essential annulus such that A\partial A does not intersect dd (hence any of djd_{j}) and m(mi,jA,μ)m(m_{i},\partial_{j}A,\mu)\longrightarrow\infty for both boundary components 1A\partial_{1}A and 2A\partial_{2}A of AA, then lengthρi(A1)0\mathrm{length}_{\rho_{i}}(\partial A_{1}^{*})\longrightarrow 0.

We note that by Lemma 5.2, {m(mi,dj,μ)}\{m(m_{i},d_{j},\mu)\} is bounded for every j=2,,pj=2,\dots,p and m(σ(mi),d,μ)m(\sigma(m_{i}),d,\mu)\longrightarrow\infty even after re-marking.

Taking a further subsequence, we can also assume that for any essential annulus EE of MM, either lengthρi(E)0\mathrm{length}_{\rho_{i}}(\partial E)\longrightarrow 0 or lengthρi(E)\mathrm{length}_{\rho_{i}}(\partial E) is bounded away from 0. Let 𝒜=kAk\mathcal{A}=\bigcup_{k}A_{k} be a maximal family of pairwise disjoint non-isotopic essential annuli such that

  1. [i]

    the length of the core curve of each annulus AkA_{k} tends to 0 (lengthρi(Ak)0\mathrm{length}_{\rho_{i}}(\partial A_{k})\longrightarrow 0 for any kk),

  2. [ii]

    𝒜\partial\mathcal{A} does not intersect dd, and

  3. [iii]

    no component of 𝒜\mathcal{A} contains a curve homotopic to dd.

Denote by VdV_{d} the component of MN(𝒜)M\setminus N(\mathcal{A}) containing dd, where N(𝒜)N(\mathcal{A}) denotes a thin regular neighbourhood of 𝒜\mathcal{A}. Let PP be the closure of VdMV_{d}\setminus\partial M, which is a union of annuli. Next we shall control the geometry of VdV_{d} and the length of dd.

Claim 5.13.

Passing to a subsequence, the restrictions {ρi|π1(Vd)}\{\rho_{i|\pi_{1}(V_{d})}\} converge and lengthρi(d)0\mathrm{length}_{\rho_{i}}(d)\longrightarrow 0.

Proof.

Let us first assume that m(mi,d1,μ)m(m_{i},d_{1},\mu)\longrightarrow\infty, and verify the hypotheses of Theorem 4.5 with W=VdW=V_{d}. The hypothesis (a) follows from the construction of VdV_{d}. The hypothesis (b) follows from the property (2) above. By Lemma 4.2, {m(mi,c,μ)}\{m(m_{i},c,\mu)\} is bounded for any simple closed curve cc intersecting dd. This observation combined with the assumption that {m(mi,dj,μ)}\{m(m_{i},d_{j},\mu)\} is bounded for any j=2,pj=2,\dots p, the property (2) above and the maximality of 𝒜\mathcal{A} yields the hypothesis (c). Now by Theorem 4.5 we can take a subsequence in such a way that the restrictions {ρi|π1(Vd)}\{\rho_{i|\pi_{1}(V_{d})}\} converge.

If lengthmi(d)0\mathrm{length}_{m_{i}}(d)\longrightarrow 0, we are done. Otherwise, since we are assuming that m(mi,d,μ)m(m_{i},d,\mu)\longrightarrow\infty, there is a sequence of subsurfaces YiMY_{i}\subset\partial M such that dYi(mi,μ)d_{Y_{i}}(m_{i},\mu)\longrightarrow\infty and dYid\subset\partial Y_{i}. Consider a simple closed curve cVdMc\subset V_{d}\cap\partial M intersecting dd. Since {ρi|π1(Vd)}\{\rho_{i|\pi_{1}(V_{d})}\} converges, {lengthρi(c)}\{\mathrm{length}_{\rho_{i}}(c)\} is bounded. Then we have dYi(mi,c)d_{Y_{i}}(m_{i},c)\longrightarrow\infty (for dYi(mi,μ)d_{Y_{i}}(m_{i},\mu)\longrightarrow\infty) and it follows from [15, Theorem 2.5] that lengthρi(d)0\mathrm{length}_{\rho_{i}}(d)\longrightarrow 0.

Suppose now that {m(mi,d1,μ)}\{m(m_{i},d_{1},\mu)\} is bounded. Since {m(σ(mi),d1,μ)}\{m(\sigma(m_{i}),d_{1},\mu)\}\longrightarrow\infty by assumption, lengthρi(d)0\mathrm{length}_{\rho_{i}}(d)\longrightarrow 0 by [14, Short Curve Theorem]. We add to PP a thin regular neighbourhood of dd on Vd\partial V_{d} and we can verify as above that the hypotheses of Theorem 4.5 are satisfied for (Vd,P)(V_{d},P). ∎

Now we are in the situation of 5.7, and we use its notations. By Lemma 5.12, g(F1)g(F_{1}) lies on the inward side of the cusp corresponding to ρ(d)\rho_{\infty}(d), and g(Fj)g(F_{j}) lies on the outward side for every j=2,,pj=2,\dots,p. Then Lemma 5.10 implies that {m(σ(mi),d1,μ)}\{m(\sigma(m_{i}),d_{1},\mu)\} is bounded. This contradicts our assumption. ∎

6. The proof of Theorem 1.1

Now we shall complete the proof of Theorem 1.1. By Lemmas 3.1 and 3.4, we can assume that every MM is strongly untwisted. Let LL be the number provided by Lemma 3.9, and consider a sequence {mi}\{m_{i}\} such that {miL+1=(ισ)L+1mi}\{m_{i}^{L+1}=(\iota_{*}\circ\sigma)^{L+1}m_{i}\} has no convergent subsequence. By Lemma 4.3, passing to a subsequence, there is a simple closed curve dL+1Md_{L+1}\subset\partial M such that m(miL+1,dL+1,μ)m(m_{i}^{L+1},d_{L+1},\mu)\longrightarrow\infty. Then we have m(σ(miL),ι(dL+1),μ)m(\sigma(m_{i}^{L}),\iota(d_{L+1}),\mu)\longrightarrow\infty. By Proposition 5.1, passing to a further subsequence, there is an incompressible annulus ALA_{L} bounded by ι(dL+1)\iota(d_{L+1}) and another simple closed curve dLMd_{L}\subset\partial M with m(miL,dL,μ)m(m_{i}^{L},d_{L},\mu)\longrightarrow\infty. Repeating this, we get a family of simple closed curves {dk,0kL+1}\{d_{k},0\leq k\leq L+1\} such that dkι(dk+1)d_{k}\cup\iota(d_{k+1}) bounds an incompressible annulus. This means that an annular neighbourhood of ι(dL+1)\iota(d_{L+1}) is LL-time vertically extendible, contradicting Lemma 3.9. This completes the proof of Theorem 1.1.

References

  • [1] Anderson, J. W., and Canary, R. D. Cores of hyperbolic 33-manifolds and limits of Kleinian groups. Amer. J. Math. 118, 4 (1996), 745–779.
  • [2] Brock, J., Bromberg, K., Canary, R., and Lecuire, C. Convergence and divergence of Kleinian surface groups. Journal of Topology 8, 3 (2015), 811–841.
  • [3] Brock, J. F., Bromberg, K. W., Canary, R. D., and Minsky, Y. N. Convergence properties of end invariants. Geom. Topol. 17, 5 (2013), 2877–2922.
  • [4] Brock, J. F., Bromberg, K. W., Canary, R. D., and Minsky, Y. N. Windows, cores and skinning maps. Ann. Sci. Éc. Norm. Supér. (4) 53, 1 (2020), 173–216.
  • [5] Jaco, W. H., and Shalen, P. B. Seifert fibered spaces in 3-manifolds. Memoirs of the American Mathematical Society 21, 220 (1979), viii–192.
  • [6] Johannson, K. Homotopy equivalences of 3-manifolds with boundaries, vol. 761 of Lecture Notes in Mathematics. Springer, Berlin, 1979.
  • [7] Kapovich, M. Hyperbolic manifolds and discrete groups. Modern Birkhäuser Classics. Birkhäuser Boston, Inc., Boston, MA, 2009. Reprint of the 2001 edition.
  • [8] Kent, R. P. Skinning maps. Duke Mathematical Journal 151, 2 (2010), 279–336.
  • [9] Marden, A. The geometry of finitely generated kleinian groups. Annals of Mathematics 99, 3 (1974), 383–462.
  • [10] Maskit, B. Kleinian groups, vol. 287 of Grundlehren der Mathematischen Wissenschaften. Springer-Verlag, Berlin, 1988.
  • [11] Masur, H. A., and Minsky, Y. N. Geometry of the complex of curves. I. Hyperbolicity. Inventiones mathematicae 138, 1 (1999), 103–149.
  • [12] Masur, H. A., and Minsky, Y. N. Geometry of the complex of curves. II. Hierarchical structure. Geometric and Functional Analysis 10, 4 (2000), 902–974.
  • [13] McCullough, D. Compact submanifolds of 3-manifolds with boundary. The Quarterly Journal of Mathematics. Oxford. Second Series 37, 147 (1986), 299–307.
  • [14] Minsky, Y. The classification of Kleinian surface groups. I. Models and bounds. Ann. of Math. (2) 171, 1 (2010), 1–107.
  • [15] Minsky, Y. N. Kleinian groups and the complex of curves. Geom. Topol. 4 (2000), 117–148.
  • [16] Morgan, J. W. On Thurston’s uniformization theorem for three-dimensional manifolds. In The Smith conjecture (New York, 1979), vol. 112 of Pure Appl. Math. Academic Press, Orlando, FL, 1984, pp. 37–125.
  • [17] Ohshika, K. Degeneration of marked hyperbolic structures in dimensions 2 and 3. In Handbook of group actions. Vol. III, vol. 40 of Adv. Lect. Math. (ALM). Int. Press, Somerville, MA, 2018, pp. 13–35.
  • [18] Otal, J.-P. Thurston’s hyperbolization of Haken manifolds. In Surveys in differential geometry, Vol. III (Cambridge, MA, 1996). Int. Press, Boston, MA, 1998, pp. 77–194.
  • [19] Scott, G. P. Compact submanifolds of 3-manifolds. Journal of the London Mathematical Society. Second Series 7, 2 (1973), 246–250.
  • [20] Thurston, W. The Geometry and Topology of Three-manifolds: Lecture Notes from Princeton University 1978-80. Mathematical Sciences Research Institute, notes taken by Kerckhoff, S. and Floyd, W.J., 1978-1980.
  • [21] Thurston, W. P. Hyperbolic structures on 33-manifolds. I. Deformation of acylindrical manifolds. Ann. of Math. (2) 124, 2 (1986), 203–246.
  • [22] Thurston, W. P. Hyperbolic Structures on 3-manifolds, II: Surface groups and 3-manifolds which fiber over the circle. arXiv.org (Jan. 1998).
  • [23] Thurston, W. P. Hyperbolic Structures on 3-manifolds, III: Deformations of 3-manifolds with incompressible boundary. https://arxiv.org/abs/math/9801058 (1998).
  • [24] Waldhausen, F. On irreducible 3-manifolds which are sufficiently large. Annals of Mathematics 87, 1 (1968), 56–88.
  • [25] Waldhausen, F. On the determination of some bounded 3-manifolds by their fundamental groups alone. Topol. Appl., Symp. Herceg-Novi (Yugoslavia) 1968, 331-332 (1969)., 1969.