A Universal Cannon-Thurston map and the surviving curve complex.
Abstract.
Using the Birman exact sequence for pure mapping class groups, we construct a universal Cannon-Thurston map onto the boundary of a curve complex for a surface with punctures we call surviving curve complex. Along the way we prove hyperbolicity of this complex and identify its boundary as a space of laminations. As a corollary we obtain a universal Cannon-Thurston map to the boundary of the ordinary curve complex, extending earlier work of the second author with Mj and Schleimer.
1. introduction
Given a closed hyperbolic –manifold that fibers over the circle with fiber a surface , Cannon and Thurston [CT07] proved that the lift to the universal covers of the inclusion extends to a continuous -equivariant map of the compactifications. This is quite remarkable as the ideal boundary map is a –equivariant, sphere–filling Peano curve. A Cannon-Thurston map, , for a type-preserving, properly discontinuous actions of the fundamental group of hyperbolic surfaces (closed or punctured) acting on hyperbolic –space was shown to exist in various situations (see [Min94, ADP99, McM01, Bow07]), with Mj [Mj14a] proving the existence in general (see Section 1.1 for a discussion of even more general Cannon-Thurston maps).
Suppose that is a hyperbolic surface with basepoint , and write . The curve complex of is a –hyperbolic space on which acts via the Birman exact sequence. In [LMS11], the second author, Mj, and Schleimer constructed a universal Cannon-Thurston map when is a closed surface of genus at least . Here we complete this picture, extending this to all surfaces with complexity .
Theorem 1.1 (Universal Cannon-Thurston Map).
Let be a connected, orientable surface with . Then there exists a subset and a continuous, –equivariant, finite-to-one surjective map . Moreover, if is any Cannon-Thurston map for a proper, type-preserving, isometric action on without accidental parabolics, then there exists a map so that factors as
For the reader familiar with Cannon-Thurston maps in the setting of cusped hyperbolic surfaces, the finite-to-one condition may seem unnatural. We address this below in the process of describing the subset . First, we elaborate on the universal property of the theorem (that is, the ``moreover" part).
Let denote the universal cover 111We will mostly be interested in real hyperbolic space in dimension , so will simply write .. A proper, type-preserving, isometric action of on a has quotient hyperbolic –manifold homeomorphic to . Each of the two ends (after removing cusp neighborhoods) is either geometrically finite or simply degenerate. In the latter case, there is an associated ending lamination that records the asymptotic geometry of the end; see [Thu78, Bon86, Min10, BCM12]]. The Cannon-Thurston map for such an action is an embedding if both ends are geometrically finite; see [Flo80]. If there are one or two degenerate ends, the Cannon-Thurston map is a quotient map onto a dendrite or the entire sphere , respectively, where a pair of points are identified if and only if and are ideal endpoints of a leaf or complementary region of the for (one of) the ending lamination(s) ; see [CT07, Min94, Bow07, Mj14b]. A more precise version of the universal property is thus given by the following. Here is the space of ending laminations of , which are all possible ending laminations of ends of hyperbolic –manifolds as above; see Section 2.6 for definitions.
Theorem 1.2.
Given two distinct points , if and only if and are the ideal endpoints of a leaf or complementary region of for some .
When has punctures, is not the most natural ``receptacle" for a universal Cannon-Thurston map. Indeed, there is another hyperbolic space whose boundary naturally properly contains . The surviving curve complex of , denoted is the subcomplex of spanned by curves that ``survive" upon filling back in. In section 4, we prove that is hyperbolic. One could alternatively verify the axioms due to Masur and Schleimer [MS13], or try to relax the conditions of Vokes [Vok] to prove hyperbolicity; see Section 4.
The projection was studied by the second author with Kent and Schleimer in [KLS09] where it was shown that for any vertex , the fiber is –equivariantly isomorphic to the Bass-Serre tree dual to the splitting of defined by the curve determined by ; see also [Har86],[HV17]. As such, there is a –equivariant map ; see §2.4. As we will see, the first part of Theorem 1.1 is a consequence of the following; see Section 8.
Theorem 1.3.
For any vertex , the map has a continuous –equivariant extension
and the induced map
is surjective and does not depend on . Moreover, is equivariant with respect to the action of the pure mapping class group .
The subset is defined analogously to the set in [LMS11]. Specifically, if and only if any geodesic ray starting at any point and limiting to at infinity has the property that every essential simple closed curve has nonempty intersection with ; see Section 8. It is straightforward to see that is the largest set on which a Cannon-Thurston map can be defined to .
As we explain below, and a pair of points in are identified by if and only if they are identified by , and thus is also finite-to-one on . It turns out that this precisely describes the difference between and . Let be the set of points for which is infinite.
Proposition 1.4.
We have .
The analogue of Theorem 1.2 is also valid for .
Theorem 1.5.
Given two distinct points , if and only if and are the ideal endpoints of a leaf or complementary region of for some .
It is easy to see that for any ending lamination , the endpoints at infinity of any leaf of (and hence also the non parabolic fixed points of complementary regions) are contained in , though this a fairly small subset; for example, almost-every point has the property that any geodesic ray limiting to has dense projection to . The complementary regions that contain parabolic fixed points are precisely the regions with infinitely many ideal endpoints. Together with Proposition 1.4 provides another description of the difference ; see Corollary 8.10.
A important ingredient in the proofs of the above theorems is an identification of the Gromov boundary , analogous to Klarreich's Theorem [Kla99b]; see Theorem 2.12. Specifically, we let denote the space of ending laminations on together with ending laminations on all proper witnesses of ; see Section 2.3. We call the space of surviving ending laminations. A more precise statement of the following is proved in Section 6; see Theorem 6.1
Theorem 1.6.
There is a –equivariant homeomorphism .
To describe the map in Theorem 1.1 we consider the map from Theorem 1.3, composed with the homeomorphism from Theorem 1.6. Since is a subset of , we can simply take to be the subset that maps onto , and compose the restriction to this subset with the homeomorphism from Klarreich's Theorem. The more geometric description of is obtained by a more detailed analysis of the map carried out in Section 8.
1.1. Historical discussion
Existence of the Cannon-Thurston map in the context of Kleinian groups is proved by several authors starting with Floyd [Flo80] for geometrically finite Kleinian groups and then by Cannon and Thurston for fibers of closed hyperbolic 3-manifolds fibering over the circle. Cannon and Thurston's work was circulated as a preprint around 1984 and inspired works of many others before it was published in 2007 [CT07]. The existence of the Cannon-Thurston map was proven by Minsky [Min94] for closed surface groups of bounded geometry and by by Mitra and Klarreich [Mit98b, Kla99a] for hyperbolic 3-manifolds of bounded geometry with an incompressible core and without parabolics. Alperin-Dicks-Porti [ADP99] proved the existence of the Cannon-Thurston map for figure eight knot complement, McMullen [McM01] for punctured torus groups, and then Bowditch [Bow07, Bow13] for more general punctured surface groups of bounded geometry. Mj completed the investigation for all finitely generated Kleinian surface groups without accidental parabolics, first for closed and then for punctured surfaces in a series of papers that culminated in the two papers [Mj14a] and [Mj14b], the latter with an appendix by S. Das. For general Kleinian groups, see Das-Mj [DM16] and Mj [Mj17], and the survey [Mj18].
Moving beyond real hyperbolic spaces, it is now classical that a quasi-isometric embedding of one Gromov hyperbolic space into another extends to an embedding of the Gromov boundaries. One of the first important generalizations of Cannon and Thurston's work outside the setting of Kleinian groups is due to Mitra in [Mit98a] who proved that given a short exact sequence
of infinite word hyperbolic groups, the Cannon–Thurston map exists and it is surjective. In this case the Cannon-Thurston map is defined between the Gromov boundary of the fiber group and the Gromov boundary of its extension . Mitra defined an algebraic ending lamination associated to points in the Gromov boundary of the base group in [Mit97], and recent work of Field [Fie20] proves that the quotient of in terms of such an ending lamination is a dendrite (compare the Kleinian discussion above).
In a different direction, Mitra later extended his existence result to trees of hyperbolic spaces; see [Mit98b]. In 2013 Baker and Riley gave the first example example of a hyperbolic subgroup of a hyperbolic group with no continuous Cannon-Thurston map ([BR13]); see also Matsuda [MO14]. On the other hand, Baker and Riley ([BR20]) proved existence of Cannon-Thurston maps even under arbitrarily heavy distortion of a free subgroup of a hyperbolic group.
For free groups and their hyperbolic extensions, Cannon-Thurston maps are better understood than arbitrary hyperbolic extensions. Kapovich and Lustig characterized the Cannon-Thurston maps for hyperbolic free-by-cyclic groups with fully irreducible monodromy [KL15]. Later Dowdall, Kapovich and Taylor characterized Cannon-Thurston maps for hyperbolic extensions of free groups coming from convex cocompact subgroups of outer automorphism group of the free group [DKT16].
1.2. Outline
In Section 2, we give preliminaries on curve complexes, witnesses and Gromov boundary of a hyperbolic space along with basics of spaces of laminations. In particular, subsection 2.4 is devoted to the construction of the survival map and in subsection 2.5 the relation between cusps and witnesses via the survival map is given. In Section 3, we define survival paths in and give an upper bound on the survival distance in terms of projection distances into curve complexes of witnesses. In Section 4 we prove the hyperbolicity of . Section 5 is devoted to the distance formula for , a-la Masur-Minsky, and as a result we prove that survival paths are uniform quasi-geodesics in . In Section 6 we explore the boundary of the survival curve complex and prove that it is homeomorphic to the space of survival ending laminations on , a result analogous to that of Klarreich [Kla99b]. In Section 7 we extend the definition of survival map to the closures of curve complexes. Finally in Section 8, we prove Theorem 1.3 and the rest of the theorems from the introduction. Specifically, we prove the existence and continuity of the map in Section 8.1 and its surjectivity in Section 8.2. Finally, we Section 8.3 we prove the universal property of as well as constructing the map .
Acknowledgements
The authors would like to thank Saul Schleimer for helpful conversations in the early stages of this work. The second author would also like to thank Autumn Kent, Mahan Mj, and Saul Schleimer for their earlier collaborations that served as partial impetus for this work.
2. Preliminaries
Throughout what follows, we assume is surface of genus with punctures, and complexity . We fix a complete hyperbolic metric of finite area on and a locally isometric universal covering . We also fix a point , and write to denote either the punctured surface or the surface with an additional marked point , with the situation dictating the intended meaning when it makes a difference. We sometimes refer to the puncture produced by removing as the –puncture. We further choose and use this to identify with the covering group of , acting by isometries.
2.1. Notation and conventions
Let with . We write to mean . We also write
When the constants are clear from the context or independent of any varying quantities and unimportant, we also write as well as . In addition, we will use the shorthand notation denote the cut-off function giving value if and otherwise.
Any connected simplicial complex will be endowed with a path metric obtained by declaring each simplex to be a regular Euclidean simplex with side lengths equal to . The vertices of a connected simplicial complex will be denoted with a subscript , and the distance between vertices will be an integer computed as the minimal length of a path in the –skeleton. By a geodesic between a pair of vertices in a simplicial complex, we mean either an isometric embedding of an interval into the –skeleton with endpoints and or the vertices encountered along such an isometric embedding, with the situation dictating the intended meaning.
2.2. Curve complexes
By a curve on a surface , we mean an essential (homotopically nontrivial and nonperipheral), simple closed curve. We often confuse a curve with its isotopy class. When convenient, we take the geodesic representative with respect to a complete finite area hyperbolic metric on the surface with geodesic boundary components (if any) and realize an isotopy class by its unique geodesic representative. A multi-curve is a disjoint union of pairwise non-isotopic curves, which we also confuse with its isotopy class and geodesic representative when convenient.
The curve complex of a surface with is the complex whose vertices are curves (up to isotopy) and whose –simplices are multi-curves with components. According to work of Masur-Minsky [MM99], curve complexes are Gromov hyperbolic. For other proofs, see [Bow06, Ham07] as well as [Aou13, Bow14, CRS14, HPW15] which prove uniform bounds on .
Theorem 2.1.
For any surface , is –hyperbolic, for some .
The surviving complex is defined to be the subcomplex of the curve complex , spanned by those curves that do not bound a twice-punctured disk, where one of the punctures is the –puncture. Given curves , we write for the distance between and (in the –skeleton).
2.3. Witnesses for and subsurface projection to witnesses
A subsurface of is either itself or a component of the complement of a small, open, regular neighborhood of a (representative of a) multi-curve ; we assume is not a pair of pants (a sphere with three boundary components/punctures). The boundary of , denoted , is the sub-multi-curve of consisting of those components that are isotopic into . As with (multi-)curves, subsurfaces is considered up to isotopy, in general, but when convenient we will choose a representative of the isotopy class without comment.
Definition 2.2.
A witness for is a subsurface such that for every curve in , no representative of the isotopy class of can be made disjoint from .
Remark 2.3.
Witnesses were introduced in a more general setting by Masur and Schleimer in [MS13] where they were called holes.
Clearly, is a witness. Note that if is the boundary of a twice-punctured disk , one of which is the –puncture, and the complementary component with is a witness. To see this, we observe that any curve in that can be isotoped disjoint from must be contained in , but the only such curve in is . It is clear that these two types of subsurfaces account for all witnesses. We let denote the set of witnesses and the set of proper witnesses. We note that any proper witness is determined by its boundary curve, : if , then is the closure of the component of not containing the –puncture.
An important tool in what follows is the subsurface projection of curves in to witnesses; see [MM00].
Definition 2.4.
(Projection to witnesses) Let be a witness for and a curve. We define the projection of to , as follows. If , then . If , then is the set of curves
where (1) we have taken representatives of and so that and intersect transversely and minimally, (2) the union is over all complementary arcs of that meet , (3) is a small regular neighborhood of of the union, and (4) we have discarded any components of that are not essential curves in . The projection is always a subset of with diameter at most ; see [MM00]. We note that is never empty by definition of a witness.
Given and a witness , we define the distance between and in by
Note that if , then is simply the usual distance between and in . According to [MM00, Lemma 2.3], projections satisfy a –Lipschitz projection bound.
Proposition 2.5.
For any two distinct curves , . In fact, for any path in connecting to , such that for all , .
We should mention that in [MM00] Masur and Minsky consider the map from and proves the second statement. Since is a subcomplex of for which every curve has non-empty projection, the first statement follows from the second.
We will also need the following important fact about projections from [MM00].
Theorem 2.6 (Bounded Geodesic Image Theorem).
Let be a geodesic in for some surface , all of whose vertices intersect a subsurface . Then, there exists a number such that,
where denotes the image of the geodesic in .
We assume (as we may) that , as this makes some of our estimates cleaner. In fact, there is a uniform that is independent of in this theorem, given by Webb [Web15].
2.4. Construction of the survival map
Consider the forgetful map
induced from the inclusion . By definition of , is well defined since every curve in determines a curve in . Each point determines a weighted multi-curve: is contained in the interior of a unique simplex, which corresponds to a multi-curve on , and the barycentric coordinates determine weights on the components of the multi-curve. According to [KLS09], the fiber of the map is naturally identified with the Bass-Serre tree associated to the corresponding weighted multi-curve: .
An important tool in our analysis is the survival map
The construction of the analogous map when is closed is described in [LMS11]. Since there are no real subtleties that arise, we describe enough of the details of the construction for our purposes, and refer the reader to that paper for details. Before getting to the precise definition of , we note that for every , the restriction of to will be denoted , and this is simpler to describe: is a –equivariantly factors as , where the action of on comes from our reference hyperbolic structure on , the associated covering map , and choice of basepoint .
To describe in general, it is convenient to construct a more natural map from which is defined as the descent to a quotient. Specifically, we will define a map
where is the component of the group of diffeomorphisms that of containing the identity (all diffeomorphisms of are assumed to extend to to diffeomorphisms of the closed surface obtained by filling in the punctures). To define , first for each curve , we let denote the geodesic representative in our fixed hyperbolic metric on , and choose once and for all so that for any two vertices , is equal to the number of components of . If is disjoint from the interior of , then , viewed as a curve on . If is contained in the interior of , then we let denote the two boundary components of this neighborhood, and define to be a point of the edge between the curves and determined by the relative distance to and . For a point inside a simplex of dimension greater than , we use the neighborhoods as well as the barycentric coordinates of inside to define ; see [LMS11, Section 2.2] for details.
Next we note that the isotopy from the identity to lifts to an isotopy from the identity to a canonical lift of . The map is then defined from our choice and the canonical lift by the equation
Alternatively, the we have the evaluation map , , which lifts to a map (given by , where again is the canonical lift), and then is defined as the descent by :
Note that every is for some (indeed, defines a locally trivial fiber bundle). As is shown in [LMS11], is well-defined independent of the choice of such a diffeomorphism with since any two differ by an isotopy fixing , and is –equivariant (where the points is used to identify the fundamental group with the group of covering transformations). It is straightforward to see that is constant on components of : two points in such a component are given by and where and are isotopic by an isotopy , so that remains outside for all .
2.5. Cusps and witnesses
The following lemma relates to the proper witnesses. Let denote the set of parabolic fixed points. Assume that for each , we choose a horoball invariant by the parabolic subgroup , the stabilizer of in . We further assume, as we may, that (1) the union of the horoballs is –invariant, (2) the horoballs are pairwise disjoint (so all projected to pairwise disjoint cusp neighborhoods of the punctures), and (3) the horoballs all project disjoint from for all curves . Recall that any proper witness is determined by its boundary.
Lemma 2.7.
There is a –equivariant bijection determined by
(1) |
for any with in the interior of the horoball . Moreover, , we have for all , and , acts trivially on .
From the lemma (and as illustrated in the proof) defines an isomorphism inverting the isomorphism .
Proof.
For any with and any curve , we have . On the other hand, is the boundary of a twice punctured disk containing the puncture, and hence is the boundary of a witness we denote . Since and are disjoint,
The same proof that is well-defined (independent of the choice of with ), shows is independent of such a choice of (up to isotopy). Therefore, is well defined by (1). Since was arbitrary and is constant on components of the complement of , we have
Given , we view as a curve disjoint from and hence is disjoint from . There is an isotopy of to fixing (since this is just a neighborhood of the cusp) and hence an isotopy of to disjoint from . This implies , proving that , as well as the formula for all .
Next observe that for any proper witness , the subcomplex uniquely determines . Therefore, the property that , together with the –equivariance of implies that is –equivariant. All that remains is to show that is a bijection. Let be the pairwise disjoint horoball cusp neighborhoods of the punctures obtained by projecting the horoballs for all .
For any proper witness , there is a diffeomorphism , isotopic to the identity by an isotopy which is the identity on for all , and so that , for some . Note that there is an arc connecting to the puncture which is disjoint from both and . It follows that and are isotopic, and thus by further isotopy (no longer the identity on ) we may assume that . Therefore, . Observe that the canonical lift has for some with . Therefore, , and so , so is surjective.
To see that is injective, suppose are such that . The two punctures surrounded by and by are therefore the same, hence there exists an element so that . By –equivariance, we must have
Choose a representative loop for with minimal self-intersection and denote this . If is simple closed, then the mapping class associated to is the product of Dehn twists (with opposite signs) in the boundary curves of a regular neighborhood of . Otherwise, fills a subsurface and is pseudo-Anosov on this subsurface by a result of Kra [Kra81] (see also [KLS09]). It follows that if and only if is disjoint from , which happens if and only if (up to isotopy relative to ). In the action of on , the element sends to , and these are the initial and terminal endpoints of the –image of the lift of with initial point . On the other hand, , and hence so is , which means that is fixes . Therefore, , and thus is injective. ∎
2.6. Spaces of laminations
We refer the reader to [Thu88], [CEG06], [FLP12], and [CB88] for details about the topics discussed here. By a lamination on a surface we mean a compact subset of the interior of foliated by complete geodesics with respect to some complete, hyperbolic metric of finite area, with (possibly empty) geodesic boundary; the geodesics in the foliation are uniquely determined by the lamination and are called the leaves. For example, any simple closed geodesic is a lamination with exactly one leaf. For a fixed complete, finite area, hyperbolic metric on , all geodesic laminations are all contained in a compact subset of the interior of . For any two complete, hyperbolic metrics of finite area, laminations that are geodesic with respect to the first are isotopic to laminations that are geodesic with respect to the second. In fact, we can remove any geodesic boundary components, and replace the resulting ends with cusps, and this remains true. We therefore sometimes view laminations as well-defined up to isotopy, unless a hyperbolic metric is specified in which case we assume they are geodesic.
A complementary region of a lamination is the image in of the closure of a component of the preimage in the universal covering; intuitively, it is the union of a complementary component together with the ``leaves bounding this component''. We view the complementary regions as immersed subsurfaces with (not necessarily compact) boundary consisting of arcs and circles (for a generic lamination, the immersion is injective, though in general it is only injective on the interior of the subsurface). We will also refer to the closure of a complementary component in the universal cover of as a complementary region (of the preimage of a lamination).
We write for the set of laminations on the surface , dropping the reference to when it is clear from the context. The set of essential simple closed curves, up to isotopy (i.e. the vertex set of ) is thus naturally a subset of . A lamination is minimal if every leaf is dense in it, and it is filling if its complementary regions are ideal polygons, or one-holed ideal polygons where the hole is either a boundary component or cusp of . A sublamination of a lamination is a subset which is also a lamination. Every lamination decomposes as a finite disjoint union of simple closed curves, minimal sublaminations without closed leaves (called the minimal components), and biinfinite isolated leaves (leaves with a neighborhood disjoint from the rest of the lamination).
There are several topologies on that will be important for us (in what follows, and whenever discussing convergence in the topologies, we view laminations as geodesic laminations with respect to a fixed complete hyperbolic metric of finite area; the resulting topology and convergence is independent of the choice of metric). The first is a metric topology called the Hausdorff topology (also known as the Chabauty topology), induced by the Hausdorff metric on the set of all compact subsets of a compact space (in our case, the compact subset of the surface that contains all geodesic laminations) defined by
If a sequence of converges to in this topology, we write . The following provides a useful characterization of convergence in this topology; see [CEG06].
Lemma 2.8.
We have if and only if
-
(1)
for all there is a sequence of points so that , and
-
(2)
for every subsequence , if , and , then .
This lemma holds not just for Hausdorff convergence of laminations, but for any sequence of compact subsets of a compact metric space with respect to the Hausdorff metric.
The set can also be equipped with a weaker topology called the coarse Hausdorff topology, [Ham06], introduced by Thurston in [Thu78] where it was called the geometric topology (see also [CEG06] where it was Thurston topology). If a sequence converges to in the coarse Hausdorff topology, then we write . The following describes convergence in this topology; see [CEG06].
Lemma 2.9.
We have if and only if condition (1) holds from Lemma 2.8.
The next corollary gives a useful way of understanding coarse Hausdorff convergence.
Corollary 2.10.
We have if and only if every Hausdorff convergent subsequence converges to a lamination containing .
Since any lamination has only finitely many sublaminations, from the corollary we see that while limits are not necessarily unique in the coarse Hausdorff topology, a sequence can have only finitely many limits. We let denote the space of ending laminations on , which are minimal, filling laminations, equipped with the coarse Hausdorff topology. As suggested by the name, these are precisely the laminations that occur as the ending laminations of a type preserving, proper, isometric action on hyperbolic –space without accidental parabolics as discussed in the introduction.
A measured lamination is a lamination together with an invariant transverse measure ; that is, an assignment of a measure on all arcs transverse to the lamination, satisfying natural subdivision properties which is invariant under isotopy of arcs preserving transversality with the lamination. The support of a measured lamination is the sublamination with the property that a transverse arc has positive measure if and only if the intersection with is nonempty, and is a union of minimal components and simple closed geodesics. We often assume that has full support, meaning . In this case, we sometimes write instead of .
The space of measured laminations on is the set of all measured laminations of full support equipped with the weak* topology on measures on an appropriate family of arcs transverse to all laminations. Given an arbitrary measured lamination, , we have is an element of , and so every measured lamination determines a unique point of . We let denote the subspace of measured laminations whose support is an ending lamination (i.e. it is in ). We write and for the respective projectivizations of and , obtained by taking the quotient by scaling measures, with the quotient topologies. The following will be useful in the sequel; see [Thu78, Chapter 8.10].
Proposition 2.11.
The map , given by , is continuous with respect to the coarse Hausdorff topology on .
For the surface , we consider the subspace
which is the union of ending laminations of all witnesses of . Similarly, we will write for those measured laminations supported on laminations in , and for its projectivization.
2.7. Gromov Boundary of a hyperbolic space
A –hyperbolic space can be equipped with a boundary at infinity, as follows. Given and a basepoint , the Gromov product of and based at is given by
Up to a bounded error (depending only on ), is the distance from to a geodesic connecting and . The quantity is estimated by the distance from the basepoint to a quasi-geodesic between and . There is an additive and multiplicative error in the estimate that depends only on the hyperbolicity constant and the quasi-geodesic constants. Using and slim triangles, we also note that for all ,
where the constants in the coarse lower bound depend only on the hyperbolicity constant.
A sequence is said to converge to infinity if . Two sequences and are equivalent if . The points in are equivalence classes of sequences converging to infinity, and if , then we say converges to and write in . The topology on the boundary is such that a sequence converges to a point if there exist sequences representing for all , and representing so that
Klarreich [Kla99b] proved that the Gromov boundary of the curve complex is naturally homeomorphic to the space of ending laminations equipped with the quotient topology from using the geometry of the Teichmuller space222In fact, Klarreich worked with the space of measured foliations, an alter ego of the space of measured laminations.. Hamenstädt [Ham06] gave a new proof, endowing with the coarse Hausdorff topology (which for is the same topology as the quotient topology), also providing additional information about convergence. Yet another proof of the version we use here was given by Pho-On [PO17].
Theorem 2.12.
For any surface equipped with a complete hyperbolic metric of finite area (possibly having geodesic boundary), there is a homeomorphism so that if and only if .
2.8. Laminations and subsurfaces
The following lemma relates coarse Hausdorff convergence of a sequence to coarse Hausdorff convergence of its projection to witnesses in important special case.
Lemma 2.13.
If and for some witness , then if and only if .
Note that for each , is a union of curves, which are not necessarily disjoint. In particular, is not necessarily a geodesic laminations, so we should be careful in discussing its coarse Hausdorff convergence. However, viewing the union as a subset of , it has diameter at most , and hence if are any two curves, for each , and , then and either both coarse Hausdorff converge to or neither does (by Theorem 2.12). Consequently, it makes sense to say that coarse Hausdorff converges to a lamination in .
Proof.
For the rest of this proof we fix a complete hyperbolic metric on and realize as an embedded subsurface with geodesic boundary. Let us first assume . After passing to an arbitrary convergent subsequence, we may assume . It suffices to show that .
Let be the decomposition into isotopy classes of arcs of intersection: that is, each is a union of all arcs of intersection of with so that any two arcs of are isotopic if and only if they are contained in the same set (we may have to pass to a further subsequence so that each intersection consists of the same number of isotopy classes, which we do). For each , let be the geodesic multi-curve produced from the isotopy class by surgery in the definition of projection. Note that and have no transverse intersections. Pass to a further subsequence so that and ; here, each is a compact subset of so Hausdorff convergence to a closed set still makes sense, though are not necessarily geodesic laminations. By Corollary 2.10 (and the discussion in the paragraph preceding this proof), , for each . Appealing to Lemma 2.8, it easily follows that . Since has no transverse intersections with , has no transverse intersections with , for each . Therefore, has no transverse intersections with , and since , has no transverse intersections with . Since , it follows that , as required.
Now in the opposite direction we assume that . Let and be as above, so that for each (after passing to a subsequence) we have
Similar to the above, and since has no transverse intersections with , has no transverse intersections with . Since is an ending lamination, . Since the convergent subsequence was arbitrary, it follows that . ∎
Finally, we note that just as curves can be projected to subsurfaces, whenever a lamination minimally intersects a subsurface in a disjoint union of arcs, we may use the same procedure to project laminations.
3. Survival paths
To understand the geometry of , the Gromov boundary, and the Cannon-Thurston map we eventually construct, we will make use of some special paths we call survival paths. To describe their construction, we set the following notation. Given a witness and , let denote a geodesic between and .
The following definition is reminiscent of hierarchy paths from [MM00], though our situation is considerably simpler.
Definition 3.1.
Given , let be any –geodesic. If is a proper witness such that is a vertex of , we say that is a witness for . Note that if is a witness for , then the immediate predecessor and successor to in are necessarily contained in (hence also in ) and we let be a geodesic (which we also view as a path in ). Replacing every consecutive triple with the path produces a path from to in which we call a survival path from to , and denote it . We call the main geodesic of and, if is witness for , we call the corresponding –geodesic the witness geodesic of for , and also say that is a witness for .
An immediate corollary of Theorem 2.6, we have
Corollary 3.2.
For any and proper witness , if , then is a witness for , for any geodesics between and .
Proof.
Since , it follows by Theorem 2.6 that some vertex of has empty projection to . But the only multi-curve in with empty projection to is , hence is a vertex of . ∎
No two consecutive vertices of can be boundaries of witness (since any two such boundaries nontrivially intersect). Therefore, the next lemma follows.
Lemma 3.3.
For any and geodesics , there are at most witnesses for . ∎
The following lemma estimates the lengths of witness geodesics on a survival path.
Lemma 3.4.
Given a survival path and a witness for , the initial and terminal vertices and of the witness geodesic segment satisfy
Consequently, of satisfies
Proof.
By Theorem 2.6 applied to the subsegments of from to and to proves the first inequality. The second is immediate from the triangle inequality. ∎
Finally we have the easy half of a distance estimate (c.f. [MM00]).
Lemma 3.5.
For any and , we have
Recall that denotes the set of all witnesses for and that is the cut-off function giving value if and otherwise.
Proof.
Since is a path from to , it suffices to prove that the length of is bounded above by the right-hand side. For each witness of whose boundary appears in , we have replaced the length two segment with , which has length . By Lemma 3.4 we have
If , this implies the length , of is less than . Otherwise, the length is less than . Let denote the witnesses for whose boundaries appear in . By Lemma 3.3, , half the length of . Further note that by Corollary 3.2, if , then is one of the witnesses , for some .
Combining all of these (and fact that ) we obtain the following bound on the length of , and hence :
∎
Lemma 3.6.
Given , if is not a witness for , then
Proof.
Lemma 3.7.
Suppose is a survival path and with , with respect to the ordering from . Then if lie on the main geodesic , then the subpath of from to is a survival path.
If and/or for proper witnesses for , respectively, then the same conclusion holds, provided the subsegments of and/or in between and has length at least .
Proof.
When are on the main geodesic, this is straightforward, since in this case, the subsegment of the main geodesic between and serves as the main geodesic for a survival path between and .
There are several cases for the second statement. The proofs are all similar, so we just describe one case where, say, with , and is in the main geodesic. The assumption in this case means that in , the distance between and is at least . Lemma 3.4 implies that , and so by the triangle inequality, . Therefore, by Theorem 2.6 any geodesic from to must pass through . In particular, the path that starts at , travels to , then continues along the subsegment of from to , is a geodesic in . We can easily build a survival path from to using this geodesic that is a subsegment of , as required. The other cases are similar. ∎
3.1. Infinite survival paths
Masur-Minsky proved that for any surface and any two points in , where is the Gromov boundary, there is a geodesic ``connecting" these points; see [MM00]. Given , we let denote such a geodesic.
The construction of survival paths above can be carried out for geodesic lines and rays in , replacing any length two path with a geodesic from to to produce a survival ray or survival line, respectively. More generally, to a geodesic segment or ray of we can construct other types of survival rays and survival lines. Specifically, first construct a survival path as above or as just described, then append to one or both endpoints an infinite witness ray (or rays). For example, for any two distinct witnesses and and points in the Gromov boundaries of and , respectively, we can construct a survival line starts and ends with geodesic rays in and , limiting to and , respectively, and having main geodesic being a segment. In this way, we see that survival lines can thus be constructed for any pair of distinct points in
and we denote such by , as in the finite case. From this discussion, we have the following.
Lemma 3.8.
For any distinct pair of elements
there exists a (possibly infinite) survival path ``connecting" these points.∎
The next proposition allows us to deduce many of the properties of survival paths to infinite survival paths.
Proposition 3.9.
Any infinite survival path (line or ray) is an increasing union of finite survival paths.
Proof.
This follows just as in the proof of Lemma 3.7. ∎
Remark 3.10.
Unless otherwise stated, the term ``survival path" will be reserved for finite survival paths. ``Infinite survival path" will mean either survival ray or survival line.
4. Hyperbolicity of the surviving curve complex
In this section we prove the following theorem using survival paths. The proof appeals to Proposition 4.3, due to Masur-Schleimer [MS13] and Bowditch ([Bow14]), which gives criteria for hyperbolicity.
Theorem 4.1.
The complex is Gromov-hyperbolic.
Remark 4.2.
There are alternate approaches to proving Theorem 4.1. For example, Masur and Schleimer provide a collection of axioms in [MS13] whose verification would imply hyperbolicity. Another approach would be to show that Vokes' condition for hyperbolicity in [Vok] which requires an action of the entire mapping class group can be relaxed to requiring an action of the stabilizer of , which is a finite index subgroup of the mapping class group. We have chosen to give a direct proof using survival paths since it is elementary and illustrates their utility.
Proposition 4.3.
Given , there exists with the following property. Suppose that is a connected graph and for each there is an associated connected subgraph including . Suppose that,
-
(1)
For all ,
-
(2)
For any with , the diameter of in is at most .
Then, is –hyperbolic.
We will apply Proposition 4.3 to the graph , and for vertices , the required subcomplex is a (choice of some) survival path . Note that if are distance one apart, then , which has diameter . Therefore, as long as , condition (2) in Theorem 4.3 will be satisfied. We therefore focus on condition (1), and express this briefly by saying that span an –slim survival triangle. The next lemma verifies condition (1) in a special case.
Lemma 4.4.
Given , there exists with the following property. If are any three points such that for all proper witness and every , then span an –slim survival triangle.
Proof.
First note that by Lemma 3.4, the length of any witness geodesic of any one of the three sides is at most ; we will use this fact throughout the proof without further mention. We also observe that by Theorem 2.6, for any and any proper witness , at least one of or is at most .
Next suppose is on a subsegment for some proper witness of . Observe that is within from either or and so by Theorem 2.6 and the triangle inequality, one of or is at most . If is any other proper witness, we claim that or is at most . To see this, note that either lies in , in , or neither. In the first two cases, or , respectively, by Theorem 2.6, while in the third case both of these inequalities hold. Therefore, since and are disjoint, , and hence or is at most .
Now let be any vertex and the nearest vertex along , and observe that . Since is –hyperbolic (for some ), there is a vertex with . Without loss of generality, we assume . Choose to be if or one of the adjacent vertices of if is the boundary of a witness. Then , so
Now suppose is a proper witness. Then at least one of or is at most as is at least one of or . If , then applying the triangle inequality, we see that
If instead, and , then the triangle inequality implies
The other two possibilities are similar, and hence .
A standard argument subdividing an –gon into triangles proves the following.
Corollary 4.5.
Given let be as in Lemma 4.4. If and are such that for all , then for all , there exists and (with all indices taken modulo ) such that .
For the remainder of the proof (and elsewhere in the paper) it is useful to make the following definition.
Definition 4.6.
Given and , consider the proper witnesses with projection at least :
and set
In words, is the set of all proper witness for which and have distance greater than .
Lemma 4.7.
For any three points and , there is at most one such that
Proof.
Suppose there exist two distinct
Then by Theorem 2.6, are (distinct) vertices in any –geodesic between any two vertices in . Choose geodesics , , and , and note that concatenating any two of these (with appropriate orientations) produces a geodesic between a pair of vertices in . Since must also lie on all –geodesics between these three vertices, it must lie on at least one of the geodesic segments to ; without loss of generality, suppose . If is not a vertex of either or , then our geodesic from to does not contain , a contradiction. Without loss of generality, we may assume . But then the geodesic subsegment between and in together with the geodesic subsegment between and in is also a geodesics (as above) and does not pass through , a contradiction. ∎
Proof of Theorem 4.1.
Let . By the triangle inequality, if , then at least one of or is greater than . By Lemma 4.7, there is at most one such that both are greater than . If such exists, denote it and write ; otherwise, write . Defining
(and defining , similarly), we can express as a disjoint union
By Theorem 2.6, the –geodesics and contain for all , and we write
so that appear in this order along and . Similarly write
The –geodesic triangle between , , and must appear as in the examples illustrated in Figure 1.
We now subdivide each of the survival paths , , and into subsegments as follows. In this subdivision, is a concatenation of witness geodesics for each witness in and complementary subsegments connecting consecutive such witness geodesics. The complementary segments are themselves survival paths obtained as concatenations of –geodesic segments and witness geodesic segments for witnesses for which . The paths and are similarly described concatenations. Applying Lemma 3.4, all of the witness segments that appear in the complementary segments (and are thus not from witnesses in ) have length at most .
Let be any point. We must show that there is some so that is uniformly bounded. There are two cases (which actually divide up further into several sub-cases), depending on whether or not lies on a witness geodesics for a witness .
Suppose first that lies on a witness geodesic for . By definition of , , and so there is also a witness geodesic . Since there are –geodesics so that every vertex has a nonempty projection to , and since (again, by definition of ), Theorem 2.6 and the triangle inequality imply
(2) |
So and are –geodesics whose starting and ending points are within distance of each other. Since is –hyperbolic for some , it follows that there is some so that . Since is a subgraph of , . We can similarly find the required if is in a witness geodesic segment for a witness .
Next suppose lies in the witness geodesic , for (if ). The argument in this sub-case is similar to the previous one, as we now describe. Let and be the –geodesic segments. Arguing as in the proof of (2), we see that the endpoints of these three geodesic segments in satisfy
Since is –hyperbolic, we can again easily deduce that for some
we have .
Finally, we assume lies in a complementary subsegment of one of the –witness subsegments of as described above. Note that both lie in one of the ``bigons'' in Figure 1 (cases (1) and (2) below) or in the single central ``triangle" (case (3) below, which happens when ). Thus, depending on which complementary subsegment we are looking at, we claim that one of the following must hold:
-
(1)
there exists so that ,
-
(2)
there exists so that , or
-
(3)
there exists and so that
.
The proofs of these statements are very similar to the proof in the case that or . If is a complementary segment which is part of a bigon and is in for some (or ), then we are in case (1) and we take the corresponding complementary segment of the bigon with (or ). It follows that all vertices of , , and have non-empty projections to , so by Theorem 2.6 and the triangle inequality we have
On the other hand, for some and similarly
and so the conclusion of (1) holds. If for some , then a symmetric argument proves (2) holds. The only other possibility is that , , and , where and , so that is a segment of the ``triangle". A completely analogous argument proves that condition (3) holds.
In any case, note that the two subsegments of the bigon (respectively, three segments of the central triangle), together with segments in curve complexes of proper witnesses give a quadrilateral (respectively, hexagon) of survival paths. Furthermore, by the triangle inequality and application of Theorem 2.6, we see that there is a uniform bound to the projections to all proper witnesses of the vertices of this quadrilateral (respectively, hexagon). Let be the constant from Lemma 4.4 for this . By Corollary 4.5, there is some on one of the other sides of this quadrilateral/hexagon so that . It may be that is in or , or that it lies in one of the witness segments. As described above, these segments have length at most , and so in this latter case, we can find with .
Combining all the above, we see that there is always some with bounded above by
This provides the required uniform bound on thinness of survival paths, and completes the proof of the theorem. ∎
5. Distance Formula
In this section we prove the following theorem.
Theorem 5.1.
For any , there exists , so that
for all .
Recall that here is shorthand for the condition and that if and , otherwise. Note that we have already proved an upper bound on of the required form in Corollary 3.5 and thus we need only prove the lower bound.
Remark 5.2.
One of the main ingredients in our proof is the following due to Behrstock [Beh04] (see [Man10] for the version here).
Lemma 5.3 (Behrstock Inequality).
Assume that and are witnesses for and with nonempty projection to both and . Then
We will also need the following application which we use to provide an ordering on the witnesses for a pair having large enough projection distances. A more general version was proved in [BKMM12] (see also [CLM12]) and is related to the partial order on domains of hierarchies from [MM00]. The version we will use is the following.
Proposition 5.4.
Suppose and are witnesses in the set . Then the following are equivalent:
(1) | (2) |
(3) | (4) |
Proof.
By Lemma 5.3 we have and . To prove we use triangle inequality:
since . The proof of is identical to the proof that . ∎
Definition 5.5.
For any , we define a relation on , declaring for , if any of the equivalent statements of the Proposition 5.4 is satisfied.
Lemma 5.6.
For any , the relation is a total order on .
Proof.
We first prove that any two element are ordered. If not, then that means Proposition 5.4 (3) fails to hold as stated, or with replacing , and thus we have and . Hence,
which contradicts the assumption that .
The relation is clearly anti-symmetric, so it remains to prove that it is transitive. To that end, let in , and we assume , hence . Since and , we have and . So by the triangle inequality
Then by Lemma 5.3, we have
So, appealing to the fact that and and Proposition 5.4 the triangle inequality implies
a contradiction. ∎
The next lemma is also useful in the proof of Theorem 5.1.
Lemma 5.7.
Let , with , and . Then,
Proof.
From our assumptions, the definition of the order on , and the triangle inequality we have
By Lemma 5.3, we have . Thus, by the definition of the order on and the triangle inequality, we have
∎
We are now ready to prove the lower bound in Theorem 5.1, which we record in the following proposition.
Proposition 5.8.
Fix . Given we have
Proof.
Let be a geodesic between in , and denote its vertices
So, is the length of . Let , suppose , and write
For each let be such that and for all . That is, is the last vertex for which . Then, if , so , Lemma 5.7 implies and so
Since the projection is –Lipschitz (see Proposition 2.5) and and are distance in , we have
Therefore, . Set and .
Given , and again appealing to Proposition 2.5, we have
Observe this inequality is trivially true for since and so the left hand side is at most in this case. Another application of Lemma 5.7 implies for all (the case is similarly trivial). Therefore
(3) |
for all .
Next, observe that since we have
Since is a subcomplex, we have and so
∎
Proof of Theorem 5.1.
As a consequence of the Theorem 5.1 we have the following two facts.
Corollary 5.9.
Given a witness , the inclusion map is a quasi-isometric embedding.
Corollary 5.10.
Survival paths are uniform quasi-geodesics in .
Moreover, we have
Lemma 5.11.
Survival paths can be reparametrized to be uniform quasi-geodesics in .
Proof.
Let be a survival path with main geodesic . For every proper witness , if there is a –witness geodesic segment in , we reparametrize along this segment so that it is traversed along an interval of length . Since such –witness geodesic segments replaced geodesic subsegments of of length , and since they lie in the –neighborhood of , this clearly defines the required reparametrization. ∎
Corollary 5.12.
Any infinite survival path is a uniform quasi-geodesic.
6. Boundary of the surviving curve complex
Recall that we denote the disjoint union of ending lamination spaces of all witnesses by
We call this the space of surviving ending laminations of , and give it the coarse Hausdorff topology.
In this section we will prove Theorem 1.6 from the introduction. In fact, we will prove the following more precise version, that will be useful for our purposes.
Theorem 6.1.
There exists a homeomorphism such that for any sequence , in if and only if .
We denote the Gromov product of based at by , and recall that the Gromov boundary of is defined to be the set of equivalence classes of sequences which converge at infinity with respect to . Throughout the rest of this section we will use (without explicit mention) the fact that the Gromov product between a pair of point (in any hyperbolic space) is uniformly estimated by the minimal distance from the basepoint to a point on a uniform quasi-geodesic between the points.
For each proper witness , Corollary 5.9 implies that embeds into . Likewise, Corollary 5.12, combined together with the fact that implies that also embeds . Using these embeddings, we view as a subspace of , for all witnesses . The next proposition says that the subspaces are all disjoint.
Proposition 6.2.
For any two witnesses for , .
Proof.
We now have a natural inclusion of the disjoint union of Gromov boundaries
In fact, this disjoint union accounts for the entire Gromov boundary.
Lemma 6.3.
We have
Proof.
Let and , and we assume without loss of generality that is a quasi-geodesic in and that the first vertex is the basepoint . If as , then given , let be such that for all . For any , the subsegment of the quasi-geodesics, , is some uniformly bounded distance to in , by hyperbolicity and Corollary 5.10. Therefore, the distance in from any point of to is at least . So the distance from any point of to is at least . Letting , it follows that in . Consequently, converges to a point in , so . For the rest of the proof, we may assume that is bounded by some constant for all .
We would like to show that there is a unique witness such that . To do that, let be the maximal Hausdorff distance in between and , for all (which is finite by hyperbolicity of and Corollary 5.10).
Claim 6.4.
Given , if for some witness , then is a witness for , for all .
Proof.
Let be such that . If is not a witness for , then every vertex of the main geodesic of has nonempty project to . Furthermore, the geodesic in from to of length at most can be extended to a path in ) to of length at most such that every vertex has nonempty project to . By Proposition 2.5 and the triangle inequality we have,
If , then since every vertex of this geodesic has nonempty projection to , it follows that . If , then there is a witness for such that , and as a result . In any case,
which is a contradiction. This proves the claim. ∎
Since , there exists such that , and hence for all , is a witness for . Let be the set of proper witnesses for , and set
Note that for all and that is nonempty for all . Since each contains no more than elements by Corollary 5.10, the (nested) union is given by for some . The boundaries of the witnesses in lie on the geodesic for all , and we let be the one furthest from . Without loss of generality, we may assume that and all agree on , for all .
For any and any witness of with further from , note that : otherwise, by Claim 6.4 would be a witness for for all and so with further from than , a contradiction to our choice of .
For any , let be the last vertex of in . By the previous paragraph together with Theorem 2.6 and the bound
we see that the subpath of from to has length bounded above by some constant , independent of . In particular, . Therefore, and converge to the same point on the Gromov boundary of . Since , which is quasi-isometrically embedded in , it follows that , as required. ∎
The next Lemma provides a convenient tool for deciding when a sequence in converges to a point in , for some proper witness .
Lemma 6.5.
Given and for some witness , then if and only if .
Proof.
Throughout, we assume , the basepoint, which without loss of generality we assume lies in , and let be any sequence converging to , so that for the Gromov product in we have as .
Since is a uniform quasi-geodesic by Corollary 5.10 it follows that
with uniform constants (where the distance on the left is the minimal distance from to the survival path).
Let be the first point of intersection of with (starting from ). By Lemma 3.4, . Consequently, because and , this means
Therefore, by hyperbolicity, the )–geodesic from (any curve in) to lies in a uniformly bounded neighborhood of , and so
If , then the right-hand side of the above coarse inequality tends to infinty, and hence so does the left-hand side. This implies .
Next suppose that . As above, we have
and so it suffices to show that the right-hand side tends to infinity as . Since , we have , and setting to be the first point of in , Lemma 3.4 implies that . Therefore, passes within –distance of on its way to . Since is a uniform quasi-geodesic by Corollary 5.10, it follows that is uniformly Hausdorff close to the initial segment from to . Since the closest point of to is, coarsely, the point , which is uniformly close to , we have
Therefore, and converge together to . This completes the proof. ∎
Proof of Theorem 6.1.
By Lemma 6.3, for any , there exists a witness so . Let , where is the homeomorphism given by Theorem 2.12. This defines a bijection .
We let with in , and prove that coarse Hausdorff converges to . Let be the witness with . According to Proposition 6.5, in . By Theorem 2.12, , and by Lemma 2.13, , as required.
To prove the other implication, we suppose that , for some , and prove that in where . Let be the witness with . By Lemma 2.13, . By Theorem 2.12, in where . By Proposition 6.5, in and .
All that remains is to show that is a homeomorphism. Throughout the remainder of this proof, we will frequently pass to subsequences, and will reindex without mention. We start by proving that is continuous. Let with as . Pass to any Hausdorff convergence subsequence so that for some lamination . If we can show that , then this will show that the original sequence coarse Hausdorff converges to , and thus will be continuous.
For each , let be a sequence with as . Since , we may pass to subsequences so that for any sequence , we have as . From the first part of the argument, as , for all . For each , pass to a subsequence so that , thus . By passing to yet a further subsequence for each , we may assume for all ; in particular, this holds for . Now pass to a subsequence of so that for some lamination , it follows that , as . Since (from the above, setting for all ), this implies that .
Since we have . If then it is the unique minimal sublamination of , and since , we have . If for some proper witness, then either is the unique minimal sublamination of , or contains . Since does not intersect the interior of , whereas is a sublamination that does nontrivially intersects the interior of , it follows that . Therefore we have in both cases, and so is continuous.
To prove continuity of , suppose , and we must show . We first pick a sequence of curves such that in . Then, as , by the first part of the proof, and after passing to subsequences as necessary, we may assume: (i) as , and hence for all ; (ii) for all ; and (iii) , for all .
Now pass to any Hausdorff convergent subsequence . It suffices to show that for this subsequence . Observe that we also have and by (ii) above we also have as , for any sequence . Thus, for example, we can conclude that , and so by the first part of the proof we have .
As equivalence classes of sequences, we thus have and . We further observe that by hyperbolicity and the conditions above, for all we have
Therefore,
from which it follows that , as required. This completes the proof. ∎
Proof of Theorem 1.6.
Let be the homeomorphism from Theorem 6.1. It suffices to show that is –equivariant. For this, let be any mapping class and any boundary point. If is any sequence with in , then since acts by isometries on . Applying Theorem 6.1 to the sequence we see that . On the other hand we also have , since acts by homeomorphisms on the space of laminations with the coarse Hausdorff topology. Therefore, , as required. ∎
7. Extended survival map
We start by introducing some notation before we define the extended survival map. First observe that there is an injection given by sending a point in the interior of the simplex with barycentric coordinates to the projective class, ; here we are viewing as a measured geodesic lamination with support and with the transverse counting measure scaled by on the component, for each . We denote the image by , which by construction admits a bijective map (inverse to the inclusion above).
By Theorem 2.12, , and so it is natural to define
and we extend the bijection to a surjective map
By Proposition 2.11 and Theorem 2.12, this is continuous at every point of .
Similar to the survival map defined in Section 2.4, we can define a map
This is defined by exactly the same procedure as in Section 2.4 of [LMS11], which goes roughly as follows: If is a measured lamination with no closed leaves in its support , and if , then . When contains closed leaves we replace those with the foliated annular neighborhoods of such curves defined in Section 2.4). When the lies on a leaf of (or the modified when there are closed leaves) we ``split apart at ", then take the –image. The same proof as that given in [LMS11, Proposition 2.9] shows that is continuous.
As in Section 2.4 (and in [LMS11]) via the lifted evaluation map , given by (for the canonical lift), the map descends to a continuous, –equivariant map making the following diagram commute:
By construction, the restriction and agree after composing with the bijection between and in the first factor. Since maps onto , if we define to be the image of via the map defined similarly to the one above, then the following diagram of –equivariant maps commutes, with the vertical arrows being bijections
(4) |
Similar to above, we define
where, recall, is the space of measured laminations on whose support is contained in . Then extends to a map
The fact that is in for any and is straightforward from the definition (c.f. [LMS11, Proposition 2.12]): for generic , is obtained from by adding the –puncture in one of the complementary components of and adjusting by a homeomorphism. With this, it follows that the map extends to a map making the following diagram, extending (4), commute.
We will call the map the extended survival map. Vertical maps in the diagram are natural maps which take projective measured laminations to their supports and they send onto and onto .
Lemma 7.1.
The extended survivial map is –equivariant and is continuous at every point of .
Proof.
To prove the continuity statement, we use the homeomorphism from Theorem 6.1 to identity with . Now suppose , , and be a sequence in such that . Passing to a subsequence, there is a measure on and a measure on such that in . Since is continuous on
By Proposition 2.11 this implies,
On the other hand, , and by Theorem 6.1 this means
in , since and .
The –equivariance follows from that of on and continuity at the remaining points. ∎
The following useful fact and it's proof are identical to the statement and proof of [LMS11, Lemma 2.14].
Lemma 7.2.
Fix . Then if and only if and are on the same leaf of, or in the same complementary region of, .
Suppose that , is a parabolic fixed point, is the horoball based at as in Section 2.5, and is the complementary region of containing . Given , choose any so that , so that . Observe that is a complementary region of containing a puncture (corresponding to ), and hence is a lamination with two punctures in the complementary component (one of which is the -puncture). Therefore, is an ending lamination in a proper witness. More precisely, by Lemma 7.2, we may assume without changing the image , and then as in the proof of Lemma 2.7, is the boundary of the witness which is disjoint from . Thus, .
In fact, every ending lamination on a proper witness arises as such an image as the next lemma shows.
Lemma 7.3.
Suppose is an ending lamination in a proper witness . Then there exists , , a complementary region of containing , and so that and .
Proof.
Note that the inclusion of is homotopic through embeddings to a diffeomorphism, after filling in (since after filling in , is peripheral). Consequently, after filling in , is isotopic to a geodesic ending lamination on . Let be a diffeomorphism isotopic to the identity with . Then where is the projective class of any transverse measure on .
Next, observe that lies in a complementary region of which is a punctured polygon (since is a simple closed curve disjoint from bounding a twice punctured disk including the -puncture). Let be the complementary region of that projects to . Then is an infinite sided polygon invariant by a parabolic subgroup fixing some . Now let be the canonical lift as in Section 2.4 and let , so that by definition . By Lemma 7.2, for any , it follows that . From the remarks preceeding this lemma, it follows that . Since , unless , it follows that , completing the proof. ∎
8. Universal Cannon–Thurston maps
In this section we will prove the following.
Theorem 1.3 For any vertex , the map has a continuous –equivariant extension
and the induced map
is surjective and does not depend on . Moreover, is equivariant with respect to the action of the pure mapping class group .
Before proceeding, we describe the subset in Theorem 1.3.
Definition 8.1.
Let be a subsurface. A point fills if,
-
•
The image of every geodesic ending in projected to intersects every curve which projects to ,
-
•
There is a geodesic ray ending at with .
Now let be the set of points that fill .
We note that when , there is a ray ending at so that is contained in a proper subsurface . The boundary of this subsurface is an essential curve and is a bounded diameter set. Thus, restricting to the set is necessary (c.f. [LMS11, Lemma 3.4]).
Given the modifications to the setup, the existence of the extension of Theorem 1.3 follows just as in the case that is closed in [LMS11]; this is outlined in Section 8.1. The surjectivity requires more substantial modification, however, and this is carried out in Section 8.2. The proof of the universal property of , as well as the discussion of , Theorem 1.1, and the relationship to Theorem 1.3 is carried out in Section 8.3.
8.1. Quasiconvex nesting and existence of Cannon–Thurston maps
In this section we will prove the existence part of the Theorem 1.3.
Theorem 8.2.
For any vertex , the induced survival map has a continuous, –equivariant extension to
Moreover, the restriction does not depend on the choice of .
By the last claim, we may denote the restriction as , without reference to the choice of . To prove this theorem, we will use the following from [LMS11, Lemma 1.9], which is a mild generalization of a lemma of Mitra in [Mit98a].
Lemma 8.3.
Let and be two hyperbolic metric spaces, and a continuous map. Fix a basepoint and a subset . Then there is a –Cannon-Thurston map
if and only if for all there is a neighborhood basis of and a collection of uniformly quasiconvex sets such that;
-
•
, and
-
•
as .
Moreover,
determines uniquely, where .
Given the adjustments already made to our setup, the proof of Theorem 8.2 is now nearly identical to [LMS11, Theorem 3.6], so we just recall the main ingredients, and explain the modifications necessary in our setting.
We fix a bi-infinite geodesic in so that is a closed geodesic that fills (i.e. nontrivially intersects every essential simple closed curve or arc on ). As in [LMS11], we construct quasi-convex sets from such as follows. Define
where is the survival map. Let denote the two half spaces bounded by and define the sets
The proofs of the following two facts about these sets are identical to the quoted results in [LMS11].
-
•
[LMS11, Proposition 3.1]: , are simplicial subcomplexes of spanned by their vertex sets and are weakly convex (meaning every two points in the set are joined by some geodesic contained in the set).
- •
Now we consider a set of pairwise disjoint translates of in so that the corresponding (closed) half spaces nest:
Since the action is properly discontinuously on , there is a such that
(5) |
Here, is the closure in . For such a sequence, we say nests down on .
On the other hand, if is a geodesic ray ending in some point which is not a parabolic fixed point, intersects infinitely many times. Hence, we can find a sequence which nests down on . In particular, for any element has a sequence that nests down on .
The main ingredient in the proof of existence of the extension is the following.
Proposition 8.4.
If nests down to , then for a basepoint , the sets are quasiconvex and we have
The proof is nearly identical to that of [LMS11, Proposition 3.5], but since it's the key to the proof of existence, we sketch it for completeness.
Sketch of proof.
Because of the nesting in , we have nesting in ,
We must show that for any , there exists so that , for all . The first observation is that because is simplicial (hence –Lipschitz), it suffices to find so that for all .
To prove this, one can use an inductive argument to construct an increasing sequence so that
Before explaining the idea, we note that this implies that are properly nested: a path from to inside must pass through a vertex of , for each , before entering the next set. Therefore, it must contain at least vertices, and so have length at least . This completes the proof by taking , since then a geodesic from to a point of will have length at least (if it leaves , then it's length is greater than ).
The main idea to find the sequence is involved in the inductive step. If we have already found , and we want to find , we suppose there is no such , and derive a contradiction. For this, assume
for all , and let be a vertex in this intersection. Set , and recall that is a component of the complement a small neighborhood of the preimage in of the geodesic representative of in . The fact that translates into the fact that and . After passing to subsequences and extracting a limit, we find a geodesic from a point on (or one of its endpoints in ) to , which projects to have empty transverse intersection with in . Since is contained in the bounded set , any subsequential Hausdorff limit does not contain an ending lamination on , by Theorem 2.12, and so any ray with no transverse intersections is eventually trapped in a subsurface (a component of the minimal subsurface of the maximal measurable sublamination of the Hausdorff limit). This contradicts the fact that , and completes the sketch of the proof. ∎
We are now ready for the proof of the existence part of Theorem 1.3.
Proof of Theorem 8.2.
The existence and continuity of follows by verifying the hypotheses in Lemma 8.3.
Fix a basepoint and let be a sequence nesting to a point . The collection of sets
is a neighborhood basis of in . By definition of
for all . By Proposition 8.4, . Therefore, by Lemma 8.3 we have a -Cannon–Thurston map defined on by
Since the sets on the right-hand side do not depend on the choice of , and since , we also write , and note that does not depend on . ∎
Observe that for all , we have
(6) |
where is any sequence nesting down on , because the intersection of the closure is in fact the intersection of the boundaries.
8.2. Surjectivity of the Cannon-Thurston map
We start with the following lemma.
Lemma 8.5.
For any we have
The analogous statement for closed is [LMS11, Lemma 3.12], but the proof there does not work in our setting. Specifically, the proof in [LMS11] appeals to Klarreich's theorem about the map from Teichmüller space to the curve complex, and extension to the boundary of that; see [Kla99b]. In our situation, the analogue would be a map from Teichmüller space to , to which Klarreich's result does not apply.
Proof.
We first claim that if is closed and –invariant then either or . This is true since the set of fixed points of pseudo-Anosov elements of is dense in and is dense in . As a result, is dense in . Since any nonempty, closed, pure mapping class group invariant subset of has to include , the claim is true.
Now we will show that contains a –invariant set. For this, first let be the set of pseudo-Anosov fixed points for elements in . Since the action leaves invariant, and since pseudo-Anosov elements act with north-south dynamics on , it follows that . Next, we need to show that for . For any point , let be a pseudo-Anosov element with . Then fixes , but is also a pseudo-Anosov element of , since is a normal subgroup of . So, , since was arbitrary. Applying the same argument to , we find . Since was arbitrary, is –invariant.
Therefore, is a nonempty closed –invariant subset of , and so both of these sets equal . ∎
To prove the surjectivity, we will need the following proposition. The exact analogue for closed is much simpler, but is false in our case (as the second condition suggests); see [LMS11, Proposition 3.13]. To state the proposition, recall that denotes the set of parabolic fixed points; see Section 2.5.
Proposition 8.6.
If is a sequence of points in with limit , then one of the following holds:
-
(1)
does not converge to a point of ; or
-
(2)
and accumulates only on points in .
To prove this, we will need the following lemma. For the remainder of this paper, we identify the points of with via Theorem 6.1.
Lemma 8.7.
Suppose and with . If in , then .
Proof.
We suppose in . Let be the horoball based at disjoint from all chosen neighborhoods of geodesics used to define as in Sections 2.4 and 2.5. Applying an isometry if necessary, we can assume that in the upper-half plane model and is stabilized by the cyclic, parabolic group . By Lemma 2.7, the –image of is a single point . Note that if for some , then remains a bounded distance from , and hence does not converge to any . Therefore, it must be that and consequently .
Proof of Proposition 8.6.
We suppose and argue as in [LMS11]. Specifically, the assumption that means that a ray ending at , after projecting to , is eventually trapped in some proper, –injective subsurface , and fills if is not an annulus. If is not an annular neighborhood of a puncture, then we arrive at the same contradiction from [LMS11, Proposition 3.13]. On the other hand, if is an annular neighborhood of a puncture, then by Lemma 8.7, , as required. ∎
Theorem 8.8.
The Cannon–Thurston map
is surjective and –equivariant.
Proof.
Let . Then, by Lemma 8.5 for some sequence . Passing to a subsequence, assume that in . If we are done since by continuity at every point of we have,
If , then by Proposition 8.6, and , where . By Lemma 6.5, . Let be the generator of . As in the proof of Lemma 8.7, is not entirely contained in any horoball based at , and hence it must be that there exists a sequence such that where is some point such that . Since is a Dehn twist in , it does not affect . Thus and hence by another application of Lemma 6.5. But in this case since does not satisfy either of the possibilities given in Lemma 8.6, and hence . But this implies
again appealing to continuity of .
The proof of –equivariance is identical to the proof of [LMS11, Theorem 1.2]. The idea is to use –equivariance, and prove for and in the dense subset of consisting of attracting fixed points of elements whose axes project to filling closed geodesics on . The point is that such points are attracting fixed points in of , but their images are also attracting fixed points in since is pseudo-Anosov by Kra's Theorem [Kra81], when viewed as an element of . The fact that and are the attracting fixed points of in and , respectively, finishes the proof. ∎
8.3. Universality and the curve complex
The following theorem on the universality is an analog of [LMS11, Corollary3.10]. While the statement is similar, it should be noted that in [LMS11], the map is finite-to-one, though this is not the case here since some of the complementary regions of the preimage in of laminations in are infinite sided ideal polygons, and whose sides accumulate to a parabolic fixed point. We follow [LMS11] where possible, and describe the differences when necessary.
Theorem 1.5.
Given two distinct points , if and only if and are the ideal endpoints of a leaf or complementary region of for some .
The proof will require a few additional facts. The first is the analogue of [LMS11, Proposition 3.8] which states that the intersections at infinity of the images of the half-spaces satisfy
(7) |
where as above, is a geodesics that projects to a closed, filling geodesic in . The next is the analogue in our setting of [LMS11, Lemma 3.9]. To describe this, recall that the element stabilizing is a pseudo-Anosov mapping class when viewed in by a theorem of Kra [Kra81]. Let be the attracting and repelling fixed points (i.e. the stable/unstable laminations). Then we have
(8) |
The proofs of these facts are identical to those in [LMS11], and we do not repeat them.
Proof of Theorem 1.5.
Given , first suppose that there is an ending lamination and which is either a leaf or complementary region of , so that and are ideal vertices of . Let be –translates of that nest down on and , respectively. Then by (6), we have
By Lemma 7.2, is a single point, which we denote . Now observe that because intersects for all sufficiently large , (8) implies
Therefore, . The exact same argument shows , and hence
as required.
Now suppose . Again by (6) there are sequences and (-translates of ) nesting down to and respectively so that
Because the intersections are nested, this implies that for all we have
Passing to a subsequence if necessary, we may assume that for all , and . Therefore, for all we have
The last equality here is an application of (7). Combining this with the description of above and (8), we have
For the last equation where we have applied (8), we have used the fact that the stable/unstable laminations of the pseudo-Anosov mapping classes corresponding to and in stabilizing and , respectively, are all distinct, hence is not one of the stable/unstable laminations.
From the equation above, we have and so that , for all . According to Lemma 7.2, there exists so that for all , and there exists a leaf or complementary region of so that . Since and nest down on and , respectively, it follows that and as . Therefore, are endpoints of a leaf of or ideal endpoints of a complementary region of , as required. ∎
We can now easily deduce the following, which also proves Proposition 1.4.
Proposition 8.9.
Given , is infinite if and only if for some proper witness .
Proof.
Theorem 1.5 implies that for , contains more than two points if and only if there is a lamination and a complementary region of so that is precisely the set of ideal points of . Moreover, in this case the proof above shows that .
On the other hand, Lemma 7.3 and the paragraph preceding it tell us that is contained in for a proper witness if and only if it is given by where and is the complementary region of containing , where with .
Finally, we note that a complementary region of a lamination has infinitely many ideal vertices if and only if it projects to a complementary region of containing a puncture, and this happens if and only if it contains a horoball for some .
Combining all three of the facts above proves the proposition. ∎
Now we define to be those points that map by to and then define to be the ``restriction'' of to . Theorem 1.2 is a consequence of the Theorem 1.5 since is the restriction of to . Then Proposition 1.4 is immediate from Proposition 8.9 and the definitions. Theorem 1.1 then follows from Theorem 1.3.
We end with an alternate description of . For , consider the subset consisting of all ideal endpoints of complementary components of which have infinitely many such ideal endpoints. That is, is the set of ideal endpoints of complementary regions that project to complementary regions of that contain a puncture. The following is thus an immediate consequence of Theorem 1.5 and Proposition 8.9.
Corollary 8.10.
The set of points that map to is
References
- [ADP99] R. C. Alperin, Warren Dicks, and J. Porti. The boundary of the Gieseking tree in hyperbolic three-space. Topology Appl., 93(3):219–259, 1999.
- [Aou13] Tarik Aougab. Uniform hyperbolicity of the graphs of curves. Geom. Topol., 17(5):2855–2875, 2013.
- [BCM12] Jeffrey F. Brock, Richard D. Canary, and Yair N. Minsky. The classification of Kleinian surface groups, II: The ending lamination conjecture. Ann. of Math. (2), 176(1):1–149, 2012.
- [Beh04] Jason Alan Behrstock. Asymptotic geometry of the mapping class group and Teichmuller space. ProQuest LLC, Ann Arbor, MI, 2004. Thesis (Ph.D.)–State University of New York at Stony Brook.
- [BH99] Martin R. Bridson and André Haefliger. Metric spaces of non-positive curvature, volume 319 of Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]. Springer-Verlag, Berlin, 1999.
- [BKMM12] Jason Behrstock, Bruce Kleiner, Yair Minsky, and Lee Mosher. Geometry and rigidity of mapping class groups. Geom. Topol., 16(2):781–888, 2012.
- [Bon86] Francis Bonahon. Bouts des variétés hyperboliques de dimension . Ann. of Math. (2), 124(1):71–158, 1986.
- [Bow06] Brian H. Bowditch. Intersection numbers and the hyperbolicity of the curve complex. J. Reine Angew. Math., 598:105–129, 2006.
- [Bow07] Brian H. Bowditch. The Cannon-Thurston map for punctured-surface groups. Math. Z., 255(1):35–76, 2007.
- [Bow13] B. H. Bowditch. Stacks of hyperbolic spaces and ends of 3-manifolds. In Geometry and topology down under, volume 597 of Contemp. Math., pages 65–138. Amer. Math. Soc., Providence, RI, 2013.
- [Bow14] Brian H. Bowditch. Uniform hyperbolicity of the curve graphs. Pacific J. Math., 269(2):269–280, 2014.
- [BR13] O. Baker and T. R. Riley. Cannon-Thurston maps do not always exist. Forum Math. Sigma, 1:Paper No. e3, 11, 2013.
- [BR20] Owen Baker and Timothy Riley. Cannon-Thurston maps, subgroup distortion, and hyperbolic hydra. Groups Geom. Dyn., 14(1):255–282, 2020.
- [CB88] Andrew J. Casson and Steven A. Bleiler. Automorphisms of surfaces after Nielsen and Thurston, volume 9 of London Mathematical Society Student Texts. Cambridge University Press, Cambridge, 1988.
- [CEG06] R. D. Canary, D. B. A. Epstein, and P. L. Green. Notes on notes of Thurston [mr0903850]. In Fundamentals of hyperbolic geometry: selected expositions, volume 328 of London Math. Soc. Lecture Note Ser., pages 1–115. Cambridge Univ. Press, Cambridge, 2006. With a new foreword by Canary.
- [CLM12] Matt T. Clay, Christopher J. Leininger, and Johanna Mangahas. The geometry of right-angled Artin subgroups of mapping class groups. Groups Geom. Dyn., 6(2):249–278, 2012.
- [CRS14] Matt Clay, Kasra Rafi, and Saul Schleimer. Uniform hyperbolicity of the curve graph via surgery sequences. Algebr. Geom. Topol., 14(6):3325–3344, 2014.
- [CT07] James W. Cannon and William P. Thurston. Group invariant Peano curves. Geom. Topol., 11:1315–1355, 2007.
- [DKT16] Spencer Dowdall, Ilya Kapovich, and Samuel J. Taylor. Cannon-Thurston maps for hyperbolic free group extensions. Israel J. Math., 216(2):753–797, 2016.
- [DM16] Shubhabrata Das and Mahan Mj. Semiconjugacies between relatively hyperbolic boundaries. Groups Geom. Dyn., 10(2):733–752, 2016.
- [Fen92] Sérgio R. Fenley. Asymptotic properties of depth one foliations in hyperbolic -manifolds. J. Differential Geom., 36(2):269–313, 1992.
- [Fen16] Sérgio R. Fenley. Quasigeodesic pseudo-Anosov flows in hyperbolic 3-manifolds and connections with large scale geometry. Adv. Math., 303:192–278, 2016.
- [Fie20] Elizabeth Field. Trees, dendrites, and the cannon-thurston map. arXiv:1907.06271 [math.GR], 2020.
- [Flo80] William J. Floyd. Group completions and limit sets of Kleinian groups. Invent. Math., 57(3):205–218, 1980.
- [FLP12] Albert Fathi, François Laudenbach, and Valentin Poénaru. Thurston's work on surfaces, volume 48 of Mathematical Notes. Princeton University Press, Princeton, NJ, 2012. Translated from the 1979 French original by Djun M. Kim and Dan Margalit.
- [Fra15] Steven Frankel. Quasigeodesic flows and sphere-filling curves. Geom. Topol., 19(3):1249–1262, 2015.
- [Gué16] François Guéritaud. Veering triangulations and Cannon-Thurston maps. J. Topol., 9(3):957–983, 2016.
- [Ham06] Ursula Hamenstädt. Train tracks and the Gromov boundary of the complex of curves. In Spaces of Kleinian groups, volume 329 of London Math. Soc. Lecture Note Ser., pages 187–207. Cambridge Univ. Press, Cambridge, 2006.
- [Ham07] Ursula Hamenstädt. Geometry of the complex of curves and of Teichmüller space. In Handbook of Teichmüller theory. Vol. I, volume 11 of IRMA Lect. Math. Theor. Phys., pages 447–467. Eur. Math. Soc., Zürich, 2007.
- [Har86] John L. Harer. The virtual cohomological dimension of the mapping class group of an orientable surface. Invent. Math., 84(1):157–176, 1986.
- [HPW15] Sebastian Hensel, Piotr Przytycki, and Richard C. H. Webb. 1-slim triangles and uniform hyperbolicity for arc graphs and curve graphs. J. Eur. Math. Soc. (JEMS), 17(4):755–762, 2015.
- [HV17] Allen Hatcher and Karen Vogtmann. Tethers and homology stability for surfaces. Algebr. Geom. Topol., 17(3):1871–1916, 2017.
- [JKLO16] Woojin Jeon, Ilya Kapovich, Christopher Leininger, and Ken'ichi Ohshika. Conical limit points and the Cannon-Thurston map. Conform. Geom. Dyn., 20:58–80, 2016.
- [KB02] Ilya Kapovich and Nadia Benakli. Boundaries of hyperbolic groups. In Combinatorial and geometric group theory (New York, 2000/Hoboken, NJ, 2001), volume 296 of Contemp. Math., pages 39–93. Amer. Math. Soc., Providence, RI, 2002.
- [KL15] Ilya Kapovich and Martin Lustig. Cannon-Thurston fibers for iwip automorphisms of . J. Lond. Math. Soc. (2), 91(1):203–224, 2015.
- [Kla99a] Erica Klarreich. Semiconjugacies between Kleinian group actions on the Riemann sphere. Amer. J. Math., 121(5):1031–1078, 1999.
- [Kla99b] Erica Klarreich. The boundary at infinity of the curve complex and the relative teichmueller space. http://nasw.org/users/klarreich/research.htm, Preprint, 1999.
- [KLS09] Richard P. Kent, IV, Christopher J. Leininger, and Saul Schleimer. Trees and mapping class groups. J. Reine Angew. Math., 637:1–21, 2009.
- [Kra81] Irwin Kra. On the Nielsen-Thurston-Bers type of some self-maps of Riemann surfaces. Acta Math., 146(3-4):231–270, 1981.
- [LMS11] Christopher J. Leininger, Mahan Mj, and Saul Schleimer. The universal Cannon-Thurston map and the boundary of the curve complex. Comment. Math. Helv., 86(4):769–816, 2011.
- [Man10] Johanna Mangahas. Uniform uniform exponential growth of subgroups of the mapping class group. Geom. Funct. Anal., 19(5):1468–1480, 2010.
- [McM01] Curtis T. McMullen. Local connectivity, Kleinian groups and geodesics on the blowup of the torus. Invent. Math., 146(1):35–91, 2001.
- [Min94] Yair N. Minsky. On rigidity, limit sets, and end invariants of hyperbolic -manifolds. J. Amer. Math. Soc., 7(3):539–588, 1994.
- [Min10] Yair Minsky. The classification of Kleinian surface groups. I. Models and bounds. Ann. of Math. (2), 171(1):1–107, 2010.
- [Mit97] M. Mitra. Ending laminations for hyperbolic group extensions. Geom. Funct. Anal., 7(2):379–402, 1997.
- [Mit98a] Mahan Mitra. Cannon-Thurston maps for hyperbolic group extensions. Topology, 37(3):527–538, 1998.
- [Mit98b] Mahan Mitra. Cannon-Thurston maps for trees of hyperbolic metric spaces. J. Differential Geom., 48(1):135–164, 1998.
- [Mj14a] Mahan Mj. Cannon-Thurston maps for surface groups. Ann. of Math. (2), 179(1):1–80, 2014.
- [Mj14b] Mahan Mj. Ending laminations and Cannon-Thurston maps. Geom. Funct. Anal., 24(1):297–321, 2014. With an appendix by Shubhabrata Das and Mj.
- [Mj17] Mahan Mj. Cannon-Thurston maps for Kleinian groups. Forum Math. Pi, 5:e1, 49, 2017.
- [Mj18] Mahan Mj. Cannon-Thurston maps. In Proceedings of the International Congress of Mathematicians—Rio de Janeiro 2018. Vol. II. Invited lectures, pages 885–917. World Sci. Publ., Hackensack, NJ, 2018.
- [MM99] Howard A. Masur and Yair N. Minsky. Geometry of the complex of curves. I. Hyperbolicity. Invent. Math., 138(1):103–149, 1999.
- [MM00] H. A. Masur and Y. N. Minsky. Geometry of the complex of curves. II. Hierarchical structure. Geom. Funct. Anal., 10(4):902–974, 2000.
- [MO14] Yoshifumi Matsuda and Shin-ichi Oguni. On Cannon-Thurston maps for relatively hyperbolic groups. J. Group Theory, 17(1):41–47, 2014.
- [Mou18] Sarah C. Mousley. Nonexistence of boundary maps for some hierarchically hyperbolic spaces. Algebr. Geom. Topol., 18(1):409–439, 2018.
- [MP11] Mahan Mj and Abhijit Pal. Relative hyperbolicity, trees of spaces and Cannon-Thurston maps. Geom. Dedicata, 151:59–78, 2011.
- [MR18] Mahan Mj and Kasra Rafi. Algebraic ending laminations and quasiconvexity. Algebr. Geom. Topol., 18(4):1883–1916, 2018.
- [MS13] Howard Masur and Saul Schleimer. The geometry of the disk complex. J. Amer. Math. Soc., 26(1):1–62, 2013.
- [PO17] Witsarut Pho-On. Infinite unicorn paths and Gromov boundaries. Groups Geom. Dyn., 11(1):353–370, 2017.
- [Thu78] William P. Thurston. Geometry and topology of -manifolds, lecture notes. Princeton University, 1978.
- [Thu88] William P. Thurston. On the geometry and dynamics of diffeomorphisms of surfaces. Bull. Amer. Math. Soc. (N.S.), 19(2):417–431, 1988.
- [Vok] Kate M. Vokes. Hierarchical hyperbolicity of graphs of multicurves. Preprint, https://arxiv.org/abs/1711.03080.
- [Web15] Richard C. H. Webb. Uniform bounds for bounded geodesic image theorems. J. Reine Angew. Math., 709:219–228, 2015.