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

A Universal Cannon-Thurston map and the surviving curve complex.

Funda Gültepe Christopher J. Leininger  and  Witsarut Pho-on Department of Mathematics and Statistics
University of Toledo
Toledo, OH, 43606
funda.gultepe@utoledo.edu http://www.math.utoledo.edu/ fgultepe/ Department of Mathematics
Rice University
Houston, TX 77005
cjl12@rice.edu https://sites.google.com/view/chris-leiningers-webpage/home Department of Mathematics
Faculty of Science
Srinakharinwirot University
Bangkok 10110, Thailand
witsarut@g.swu.ac.th https://sites.google.com/g.swu.ac.th/witsarut/
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.

The first author was partially supported by a University of Toledo startup grant.
The second author was partially supported by NSF grant DMS-1510034, 1811518, and 2106419.

1. introduction

Given a closed hyperbolic 33–manifold MM that fibers over the circle with fiber a surface SS, Cannon and Thurston [CT07] proved that the lift to the universal covers 23\mathbb{H}^{2}\to\mathbb{H}^{3} of the inclusion SMS\to M extends to a continuous π1(S)\pi_{1}(S)-equivariant map of the compactifications. This is quite remarkable as the ideal boundary map 𝕊1𝕊2\mathbb{S}^{1}_{\infty}\to\mathbb{S}^{2}_{\infty} is a π1S\pi_{1}S–equivariant, sphere–filling Peano curve. A Cannon-Thurston map, 𝕊1𝕊2\mathbb{S}^{1}_{\infty}\to\mathbb{S}^{2}_{\infty}, for a type-preserving, properly discontinuous actions of the fundamental group π1S\pi_{1}S of hyperbolic surfaces (closed or punctured) acting on hyperbolic 33–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 SS is a hyperbolic surface with basepoint zSz\in S, and write S˙=S{z}\dot{S}=S\smallsetminus\{z\}. The curve complex of S˙\dot{S} is a δ\delta–hyperbolic space on which π1S=π1(S,z)\pi_{1}S=\pi_{1}(S,z) acts via the Birman exact sequence. In [LMS11], the second author, Mj, and Schleimer constructed a universal Cannon-Thurston map when SS is a closed surface of genus at least 22. Here we complete this picture, extending this to all surfaces SS with complexity ξ(S)2\xi(S)\geq 2.

Theorem 1.1 (Universal Cannon-Thurston Map).

Let SS be a connected, orientable surface with ξ(S)2\xi(S)\geq 2. Then there exists a subset 𝕊𝒜01𝕊1\mathbb{S}^{1}_{\mathcal{A}_{0}}\subset\mathbb{S}^{1}_{\infty} and a continuous, π1S\pi_{1}S–equivariant, finite-to-one surjective map Φ0:𝕊𝒜01𝒞(S˙)\partial\Phi_{0}\colon\mathbb{S}^{1}_{\mathcal{A}_{0}}\to\partial{\mathcal{C}}(\dot{S}). Moreover, if i:𝕊1𝕊2\partial i\colon\mathbb{S}^{1}_{\infty}\to\mathbb{S}^{2}_{\infty} is any Cannon-Thurston map for a proper, type-preserving, isometric action on 3\mathbb{H}^{3} without accidental parabolics, then there exists a map q:i(𝕊𝒜01)𝒞(S˙)q\colon\partial i(\mathbb{S}^{1}_{\mathcal{A}_{0}})\to\partial{\mathcal{C}}(\dot{S}) so that Φ0\partial\Phi_{0} factors as

𝕊𝒜01\textstyle{\mathbb{S}^{1}_{\mathcal{A}_{0}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}i\scriptstyle{\partial i\quad}Φ0\scriptstyle{\partial\Phi_{0}}i(𝕊𝒜01)\textstyle{\partial i(\mathbb{S}^{1}_{\mathcal{A}_{0}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}q\scriptstyle{q}𝒞(S˙).\textstyle{\partial{\mathcal{C}}(\dot{S}).}

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 𝕊𝒜01𝕊1\mathbb{S}^{1}_{\mathcal{A}_{0}}\subset\mathbb{S}^{1}_{\infty}. First, we elaborate on the universal property of the theorem (that is, the ``moreover" part).

Let p:=2Sp\colon\mathbb{H}=\mathbb{H}^{2}\to S denote the universal cover 111We will mostly be interested in real hyperbolic space in dimension 22, so will simply write =2\mathbb{H}=\mathbb{H}^{2}.. A proper, type-preserving, isometric action of π1S\pi_{1}S on a 3\mathbb{H}^{3} has quotient hyperbolic 33–manifold homeomorphic to S×S\times\mathbb{R}. 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 𝕊1𝕊2\mathbb{S}^{1}_{\infty}\to\mathbb{S}^{2}_{\infty} 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 𝕊2\mathbb{S}^{2}_{\infty}, respectively, where a pair of points x,y𝕊1x,y\in\mathbb{S}^{1}_{\infty} are identified if and only if xx and yy are ideal endpoints of a leaf or complementary region of the p1()p^{-1}(\mathcal{L}) for (one of) the ending lamination(s) \mathcal{L}; see [CT07, Min94, Bow07, Mj14b]. A more precise version of the universal property is thus given by the following. Here (S)\mathcal{EL}(S) is the space of ending laminations of SS, which are all possible ending laminations of ends of hyperbolic 33–manifolds as above; see Section 2.6 for definitions.

Theorem 1.2.

Given two distinct points x,y𝕊𝒜01x,y\in\mathbb{S}^{1}_{\mathcal{A}_{0}}, Φ0(x)=Φ0(y)\partial\Phi_{0}(x)=\partial\Phi_{0}(y) if and only if xx and yy are the ideal endpoints of a leaf or complementary region of p1()p^{-1}(\mathcal{L}) for some (S)\mathcal{L}\in\mathcal{EL}(S).

When SS has punctures, 𝒞(S˙)\partial{\mathcal{C}}(\dot{S}) is not the most natural ``receptacle" for a universal Cannon-Thurston map. Indeed, there is another hyperbolic space whose boundary naturally properly contains 𝒞(S˙)\partial{\mathcal{C}}(\dot{S}). The surviving curve complex of S˙\dot{S}, denoted 𝒞s(S˙){\mathcal{C}}^{s}(\dot{S}) is the subcomplex of 𝒞(S˙){\mathcal{C}}(\dot{S}) spanned by curves that ``survive" upon filling zz back in. In section 4, we prove that 𝒞s(S˙){\mathcal{C}}^{s}(\dot{S}) 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 Π:𝒞s(S˙)𝒞(S)\Pi\colon{\mathcal{C}}^{s}(\dot{S})\to{\mathcal{C}}(S) was studied by the second author with Kent and Schleimer in [KLS09] where it was shown that for any vertex v𝒞(S)v\in{\mathcal{C}}(S), the fiber Π1(v)\Pi^{-1}(v) is π1S\pi_{1}S–equivariantly isomorphic to the Bass-Serre tree dual to the splitting of π1S\pi_{1}S defined by the curve determined by vv; see also [Har86],[HV17]. As such, there is a π1S\pi_{1}S–equivariant map Φv:Π1(v)𝒞s(S˙)\Phi_{v}\colon\mathbb{H}\to\Pi^{-1}(v)\subset{\mathcal{C}}^{s}(\dot{S}); 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 v𝒞v\in{\mathcal{C}}, the map Φv:𝒞s(S˙)\Phi_{v}:\mathbb{H}\rightarrow{\mathcal{C}}^{s}(\dot{S}) has a continuous π1(S)\pi_{1}(S)–equivariant extension

Φ¯v:𝕊𝒜1𝒞¯s(S˙)\overline{\Phi}_{v}:\mathbb{H}\cup\mathbb{S}^{1}_{\mathcal{A}}\rightarrow\overline{\mathcal{C}}^{s}(\dot{S})

and the induced map

Φ=Φ¯v|𝕊𝒜1:𝕊𝒜1𝒞s(S˙)\partial\Phi=\overline{\Phi}_{v}|_{\mathbb{S}^{1}_{\mathcal{A}}}:\mathbb{S}^{1}_{\mathcal{A}}\rightarrow\partial{\mathcal{C}}^{s}(\dot{S})

is surjective and does not depend on vv. Moreover, Φ\partial\Phi is equivariant with respect to the action of the pure mapping class group PMod(S˙)\operatorname{PMod}(\dot{S}).

The subset 𝕊𝒜1𝕊1\mathbb{S}^{1}_{\mathcal{A}}\subset\mathbb{S}^{1}_{\infty} is defined analogously to the set 𝔸𝕊1\mathbb{A}\subset\mathbb{S}^{1}_{\infty} in [LMS11]. Specifically, x𝕊𝒜1x\in\mathbb{S}^{1}_{\mathcal{A}} if and only if any geodesic ray rr\subset\mathbb{H} starting at any point and limiting to xx at infinity has the property that every essential simple closed curve αS\alpha\subset S has nonempty intersection with p(r)p(r); see Section 8. It is straightforward to see that 𝕊𝒜1\mathbb{S}^{1}_{\mathcal{A}} is the largest set on which a Cannon-Thurston map can be defined to 𝒞s(S˙)\partial{\mathcal{C}}^{s}(\dot{S}).

As we explain below, 𝕊𝒜01𝕊𝒜1\mathbb{S}^{1}_{\mathcal{A}_{0}}\subsetneq\mathbb{S}^{1}_{\mathcal{A}} and a pair of points in 𝕊𝒜01\mathbb{S}^{1}_{\mathcal{A}_{0}} are identified by Φ0\partial\Phi_{0} if and only if they are identified by Φ\partial\Phi, and thus Φ\partial\Phi is also finite-to-one on 𝕊𝒜01\mathbb{S}^{1}_{\mathcal{A}_{0}}. It turns out that this precisely describes the difference between 𝕊𝒜1\mathbb{S}^{1}_{\mathcal{A}} and 𝕊𝒜01\mathbb{S}^{1}_{\mathcal{A}_{0}}. Let Z𝒞s(S˙)Z\subset\partial{\mathcal{C}}^{s}(\dot{S}) be the set of points xx for which Φ1(x)\partial\Phi^{-1}(x) is infinite.

Proposition 1.4.

We have 𝕊𝒜1𝕊𝒜01=Φ1(Z)\displaystyle\mathbb{S}^{1}_{\mathcal{A}}{\smallsetminus}\mathbb{S}^{1}_{\mathcal{A}_{0}}=\partial\Phi^{-1}(Z).

The analogue of Theorem 1.2 is also valid for Φ\Phi.

Theorem 1.5.

Given two distinct points x,y𝕊𝒜1x,y\in\mathbb{S}^{1}_{\mathcal{A}}, Φ(x)=Φ(y)\partial\Phi(x)=\partial\Phi(y) if and only if xx and yy are the ideal endpoints of a leaf or complementary region of p1()p^{-1}(\mathcal{L}) for some (S)\mathcal{L}\in\mathcal{EL}(S).

It is easy to see that for any ending lamination (S)\mathcal{L}\in\mathcal{EL}(S), the endpoints at infinity of any leaf of p1()p^{-1}(\mathcal{L}) (and hence also the non parabolic fixed points of complementary regions) are contained in 𝕊𝒜1\mathbb{S}^{1}_{\mathcal{A}}, though this a fairly small subset; for example, almost-every point x𝕊1x\in\mathbb{S}^{1}_{\infty} has the property that any geodesic ray limiting to xx has dense projection to SS. 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 𝕊𝒜1𝕊𝒜01\mathbb{S}^{1}_{\mathcal{A}}{\smallsetminus}\mathbb{S}^{1}_{\mathcal{A}_{0}}; see Corollary 8.10.

A important ingredient in the proofs of the above theorems is an identification of the Gromov boundary 𝒞s(S˙)\partial{\mathcal{C}}^{s}(\dot{S}), analogous to Klarreich's Theorem [Kla99b]; see Theorem 2.12. Specifically, we let s(S˙)\mathcal{E}\mathcal{L}^{s}(\dot{S}) denote the space of ending laminations on S˙\dot{S} together with ending laminations on all proper witnesses of S˙\dot{S}; see Section 2.3. We call s(S˙)\mathcal{EL}^{s}(\dot{S}) 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 PMod(S˙)\operatorname{PMod}(\dot{S})–equivariant homeomorphism 𝒞s(S˙)s(S˙)\partial{\mathcal{C}}^{s}(\dot{S})\to\mathcal{EL}^{s}(\dot{S}).

To describe the map Φ0\partial\Phi_{0} in Theorem 1.1 we consider the map Φ:𝕊𝒜1𝒞s(S˙)\partial\Phi\colon\mathbb{S}^{1}_{\mathcal{A}}\to\partial{\mathcal{C}}^{s}(\dot{S}) from Theorem 1.3, composed with the homeomorphism 𝒞s(S˙)s(S˙)\partial{\mathcal{C}}^{s}(\dot{S})\to\mathcal{EL}^{s}(\dot{S}) from Theorem 1.6. Since (S˙)\mathcal{EL}(\dot{S}) is a subset of s(S˙)\mathcal{EL}^{s}(\dot{S}), we can simply take 𝕊𝒜01𝕊𝒜1\mathbb{S}^{1}_{\mathcal{A}_{0}}\subset\mathbb{S}^{1}_{\mathcal{A}} to be the subset that maps onto (S˙)\mathcal{EL}(\dot{S}), and compose the restriction Φ\partial\Phi to this subset with the homeomorphism (S˙)𝒞(S˙)\mathcal{EL}(\dot{S})\to\partial{\mathcal{C}}(\dot{S}) from Klarreich's Theorem. The more geometric description of 𝕊𝒜01\mathbb{S}^{1}_{\mathcal{A}_{0}} is obtained by a more detailed analysis of the map Φ\partial\Phi 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

1HΓG11\rightarrow H\rightarrow\Gamma\rightarrow G\rightarrow 1

of infinite word hyperbolic groups, the Cannon–Thurston map exists and it is surjective. In this case the Cannon-Thurston map HΓ\partial H\rightarrow\partial\Gamma is defined between the Gromov boundary H\partial H of the fiber group HH and the Gromov boundary Γ\partial\Gamma of its extension Γ\Gamma. Mitra defined an algebraic ending lamination associated to points in the Gromov boundary of the base group GG in [Mit97], and recent work of Field [Fie20] proves that the quotient of H\partial H 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].

Finally we note that we have only discussed a few of the many results on the existence or structure of Cannon-Thurston maps in various settings. For more see e.g. [Mj14b, MR18, MP11, JKLO16, Gué16, Fen92, Fra15, Fen16, Mou18]).

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 𝒞s(S˙){\mathcal{C}^{s}(\dot{S})} and give an upper bound on the survival distance dsd^{s} in terms of projection distances into curve complexes of witnesses. In Section 4 we prove the hyperbolicity of 𝒞s(S˙){\mathcal{C}^{s}(\dot{S})}. Section 5 is devoted to the distance formula for 𝒞s(S˙){\mathcal{C}^{s}(\dot{S})}, a-la Masur-Minsky, and as a result we prove that survival paths are uniform quasi-geodesics in 𝒞s(S˙){\mathcal{C}^{s}(\dot{S})}. In Section 6 we explore the boundary of the survival curve complex 𝒞s(S˙){\mathcal{C}^{s}(\dot{S})} and prove that it is homeomorphic to the space of survival ending laminations on S˙\dot{S}, 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 Φ\partial\Phi in Section 8.1 and its surjectivity in Section 8.2. Finally, we Section 8.3 we prove the universal property of Φ\partial\Phi as well as constructing the map Φ0\partial\Phi_{0}.

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 SS is surface of genus g0g\geq 0 with n0n\neq 0 punctures, and complexity ξ(S)=3g3+n2\xi(S)=3g-3+n\geq 2. We fix a complete hyperbolic metric of finite area on SS and a locally isometric universal covering p:Sp\colon\mathbb{H}\to S. We also fix a point zSz\in S, and write S˙\dot{S} to denote either the punctured surface S{z}S{\smallsetminus}\{z\} or the surface with an additional marked point (S,z)(S,z), with the situation dictating the intended meaning when it makes a difference. We sometimes refer to the puncture produced by removing zz as the zz–puncture. We further choose z~p1(z)\tilde{z}\in p^{-1}(z)\subset\mathbb{H} and use this to identify π1S=π1(S,z)\pi_{1}S=\pi_{1}(S,z) with the covering group of p:Sp\colon\mathbb{H}\to S, acting by isometries.

2.1. Notation and conventions

Let x,y,C,K0x,y,C,K\geq 0 with K1K\geq 1. We write xK,Cyx\stackrel{{\scriptstyle K,C}}{{\preceq}}y to mean xKy+Cx\leq Ky+C. We also write

xK,CyxK,Cy and xK,Cy.x\stackrel{{\scriptstyle K,C}}{{\asymp}}y\qquad\Longleftrightarrow\qquad x\stackrel{{\scriptstyle K,C}}{{\preceq}}y\,\,\mbox{ and }x\,\,\stackrel{{\scriptstyle K,C}}{{\succeq}}y.

When the constants are clear from the context or independent of any varying quantities and unimportant, we also write xyx\preceq y as well as xyx\asymp y. In addition, we will use the shorthand notation {{x}}C{\{\{}x{\}\}}_{C} denote the cut-off function giving value xx if xCx\geq C and 0 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 11. The vertices of a connected simplicial complex will be denoted with a subscript 0, and the distance between vertices will be an integer computed as the minimal length of a path in the 11–skeleton. By a geodesic between a pair of vertices v,wv,w in a simplicial complex, we mean either an isometric embedding of an interval into the 11–skeleton with endpoints vv and ww 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 YY, 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 YY with 2ξ(Y)<2\leq\xi(Y)<\infty is the complex 𝒞(Y){\mathcal{C}}(Y) whose vertices are curves (up to isotopy) and whose kk–simplices are multi-curves with k+1k+1 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 δ\delta.

Theorem 2.1.

For any surface YY, 𝒞(Y){\mathcal{C}}(Y) is δ\delta–hyperbolic, for some δ>0\delta>0.

The surviving complex 𝒞s(S˙){\mathcal{C}}^{s}(\dot{S}) is defined to be the subcomplex of the curve complex 𝒞(S˙){\mathcal{C}}(\dot{S}), spanned by those curves that do not bound a twice-punctured disk, where one of the punctures is the zz–puncture. Given curves α,β𝒞0s(S˙)\alpha,\beta\in{\mathcal{C}}^{s}_{0}(\dot{S}), we write ds(α,β)d^{s}(\alpha,\beta) for the distance between α\alpha and β\beta (in the 11–skeleton).

2.3. Witnesses for 𝒞s(S˙){\mathcal{C}^{s}(\dot{S})} and subsurface projection to witnesses

A subsurface of S˙\dot{S} is either S˙\dot{S} itself or a component YS˙Y\subset\dot{S} of the complement of a small, open, regular neighborhood of a (representative of a) multi-curve AA; we assume YY is not a pair of pants (a sphere with three boundary components/punctures). The boundary of YY, denoted Y\partial Y, is the sub-multi-curve of AA consisting of those components that are isotopic into YY. 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 𝒞s(S˙){\mathcal{C}^{s}(\dot{S})} is a subsurface WS˙W\subset\dot{S} such that for every curve α\alpha in S˙\dot{S}, no representative of the isotopy class of α\alpha can be made disjoint from WW.

Remark 2.3.

Witnesses were introduced in a more general setting by Masur and Schleimer in [MS13] where they were called holes.

Clearly, S˙\dot{S} is a witness. Note that if β\beta is the boundary of a twice-punctured disk DD, one of which is the zz–puncture, and the complementary component WSW\subset S with W=β\partial W=\beta is a witness. To see this, we observe that any curve α\alpha in 𝒞s(S˙){\mathcal{C}^{s}(\dot{S})} that can be isotoped disjoint from WW must be contained in DD, but the only such curve in S˙\dot{S} is β\beta. It is clear that these two types of subsurfaces account for all witnesses. We let Ω(S˙)\Omega(\dot{S}) denote the set of witnesses and Ω0(S˙)=Ω(S˙){S˙}\Omega_{0}(\dot{S})=\Omega(\dot{S})\smallsetminus\{\dot{S}\} the set of proper witnesses. We note that any proper witness WW is determined by its boundary curve, W\partial W: if WS˙W\neq\dot{S}, then WW is the closure of the component of S˙W\dot{S}{\smallsetminus}\partial W not containing the zz–puncture.

An important tool in what follows is the subsurface projection of curves in 𝒞s(S˙){\mathcal{C}^{s}(\dot{S})} to witnesses; see [MM00].

Definition 2.4.

(Projection to witnesses) Let WS˙W\subseteq\dot{S} be a witness for 𝒞s(S˙){\mathcal{C}^{s}(\dot{S})} and α𝒞0s(S˙)\alpha\in{\mathcal{C}}^{s}_{0}(\dot{S}) a curve. We define the projection of α\alpha to WW, πW(α)\pi_{W}(\alpha) as follows. If W=S˙W=\dot{S}, then πW(α)=α\pi_{W}(\alpha)=\alpha. If WS˙W\neq\dot{S}, then πW(α)\pi_{W}(\alpha) is the set of curves

πW(α)=(𝒩(α0W)).\pi_{W}(\alpha)=\bigcup\partial(\mathcal{N}(\alpha_{0}\cup\partial W)).

where (1) we have taken representatives of α\alpha and WW so that α\alpha and W\partial W intersect transversely and minimally, (2) the union is over all complementary arcs α0\alpha_{0} of αW\alpha{\smallsetminus}\partial W that meet WW, (3) 𝒩(α0W)\mathcal{N}(\alpha_{0}\cup\partial W) is a small regular neighborhood of of the union, and (4) we have discarded any components of (𝒩(α0W))\partial(\mathcal{N}(\alpha_{0}\cup\partial W)) that are not essential curves in WW. The projection πW(α)\pi_{W}(\alpha) is always a subset of 𝒞(W){\mathcal{C}}(W) with diameter at most 22; see [MM00]. We note that πW(α)\pi_{W}(\alpha) is never empty by definition of a witness.

Given α,β𝒞0s(S˙)\alpha,\beta\in{\mathcal{C}}^{s}_{0}(\dot{S}) and a witness WW, we define the distance between α\alpha and β\beta in WW by

dW(α,β)=diam{πW(α)πW(β)}.d_{W}(\alpha,\beta)=\operatorname{diam}\{\pi_{W}(\alpha)\cup\pi_{W}(\beta)\}.

Note that if W=S˙W=\dot{S}, then dS˙(α,β)d_{\dot{S}}(\alpha,\beta) is simply the usual distance between α\alpha and β\beta in 𝒞(S˙){\mathcal{C}}(\dot{S}). According to [MM00, Lemma 2.3], projections satisfy a 22–Lipschitz projection bound.

Proposition 2.5.

For any two distinct curves α,β𝒞s(S˙)\alpha,\beta\in{\mathcal{C}^{s}(\dot{S})}, dW(α,β)2ds(α,β)d_{W}(\alpha,\beta)\leq 2d^{s}(\alpha,\beta). In fact, for any path v0,,vnv_{0},\ldots,v_{n} in 𝒞(S˙){\mathcal{C}}(\dot{S}) connecting α\alpha to β\beta, such that πW(vj)\pi_{W}(v_{j})\neq\emptyset for all jj, dW(α,β)2nd_{W}(\alpha,\beta)\leq 2n.

We should mention that in [MM00] Masur and Minsky consider the map from 𝒞(S˙){\mathcal{C}}(\dot{S}) and proves the second statement. Since 𝒞s(S˙){\mathcal{C}}^{s}(\dot{S}) is a subcomplex of 𝒞(S˙){\mathcal{C}}(\dot{S}) 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 𝒢\mathcal{G} be a geodesic in 𝒞(Y){\mathcal{C}}(Y) for some surface YY, all of whose vertices intersect a subsurface WYW\subset Y. Then, there exists a number M>0M>0 such that,

diamW(πW(𝒢))<M\operatorname{diam}_{W}(\pi_{W}(\mathcal{G}))<M

where πW(𝒢)\pi_{W}(\mathcal{G}) denotes the image of the geodesic 𝒢\mathcal{G} in WW.

We assume (as we may) that M8M\geq 8, as this makes some of our estimates cleaner. In fact, there is a uniform MM that is independent of YY in this theorem, given by Webb [Web15].

2.4. Construction of the survival map

Consider the forgetful map

Π:𝒞s(S˙)𝒞(S)\Pi:{{\mathcal{C}^{s}(\dot{S})}}\rightarrow{\mathcal{C}}(S)

induced from the inclusion S˙S\dot{S}\rightarrow S. By definition of 𝒞s(S˙){\mathcal{C}^{s}(\dot{S})}, Π\Pi is well defined since every curve in 𝒞s(S˙){\mathcal{C}^{s}(\dot{S})} determines a curve in 𝒞(S){\mathcal{C}}(S). Each point v𝒞(S)v\in{\mathcal{C}}(S) determines a weighted multi-curve: vv is contained in the interior of a unique simplex, which corresponds to a multi-curve on SS, and the barycentric coordinates determine weights on the components of the multi-curve. According to [KLS09], the fiber of the map Π\Pi is naturally identified with the Bass-Serre tree associated to the corresponding weighted multi-curve: Π1(v)=Tv\Pi^{-1}(v)=T_{v}.

An important tool in our analysis is the survival map

Φ:𝒞(S)×𝒞s(S˙).\Phi:{\mathcal{C}}(S)\times\mathbb{H}\rightarrow{\mathcal{C}^{s}(\dot{S})}.

The construction of the analogous map when SS 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 Φ\Phi, we note that for every v𝒞(S)v\in{\mathcal{C}}(S), the restriction of Φ\Phi to {v}×\mathbb{H}\cong\{v\}\times\mathbb{H} will be denoted Φv:𝒞s(S˙)\Phi_{v}\colon\mathbb{H}\to{\mathcal{C}}^{s}(\dot{S}), and this is simpler to describe: Φv\Phi_{v} is a π1S\pi_{1}S–equivariantly factors as TvΠ1(v)\mathbb{H}\to T_{v}\to\Pi^{-1}(v), where the action of π1S\pi_{1}S on \mathbb{H} comes from our reference hyperbolic structure on SS, the associated covering map p:Sp\colon\mathbb{H}\to S, and choice of basepoint z~p1(z)\widetilde{z}\in p^{-1}(z).

To describe Φ\Phi in general, it is convenient to construct a more natural map from which Φ\Phi is defined as the descent to a quotient. Specifically, we will define a map

Φ~:𝒞(S)×Diff0(S)𝒞s(S˙)\widetilde{\Phi}\colon{\mathcal{C}}(S)\times\operatorname{Diff}_{0}(S)\to{\mathcal{C}^{s}(\dot{S})}

where Diff0(S)\operatorname{Diff}_{0}(S) is the component of the group of diffeomorphisms that of SS containing the identity (all diffeomorphisms of SS are assumed to extend to to diffeomorphisms of the closed surface obtained by filling in the punctures). To define Φ~\widetilde{\Phi}, first for each curve α𝒞0(S)\alpha\in{\mathcal{C}}_{0}(S), we let α\alpha denote the geodesic representative in our fixed hyperbolic metric on SS, and choose once and for all ϵ(α)>0\epsilon(\alpha)>0 so that for any two vertices α,α\alpha,\alpha^{\prime}, i(α,α)i(\alpha,\alpha^{\prime}) is equal to the number of components of Nϵ(α)(α)Nϵ(α)(α)N_{\epsilon(\alpha)}(\alpha)\cap N_{\epsilon(\alpha^{\prime})}(\alpha^{\prime}). If f(z)f(z) is disjoint from the interior of Nϵ(α)(α)N_{\epsilon(\alpha)}(\alpha), then Φ~(α,f)=f1(α)\widetilde{\Phi}(\alpha,f)=f^{-1}(\alpha), viewed as a curve on S˙\dot{S}. If f(z)f(z) is contained in the interior of Nϵ(α)(α)N_{\epsilon(\alpha)}(\alpha), then we let α±\alpha_{\pm} denote the two boundary components of this neighborhood, and define Φ~(α,f)\widetilde{\Phi}(\alpha,f) to be a point of the edge between the curves f1(α)f^{-1}(\alpha_{-}) and f1(α+)f^{-1}(\alpha_{+}) determined by the relative distance to α+\alpha_{+} and α\alpha_{-}. For a point s𝒞(S)s\in{\mathcal{C}}(S) inside a simplex Δ\Delta of dimension greater than 0, we use the neighborhoods as well as the barycentric coordinates of ss inside Δ\Delta to define Φ~(s,f)Π1(s)Π1(Δ)\widetilde{\Phi}(s,f)\in\Pi^{-1}(s)\subset\Pi^{-1}(\Delta); see [LMS11, Section 2.2] for details.

Next we note that the isotopy from the identity to ff lifts to an isotopy from the identity to a canonical lift f~\widetilde{f} of ff. The map Φ\Phi is then defined from our choice z~p1(z)\widetilde{z}\in p^{-1}(z) and the canonical lift by the equation

Φ(α,f~(z~))=Φ~(α,f).\Phi(\alpha,\widetilde{f}(\widetilde{z}))=\widetilde{\Phi}(\alpha,f).

Alternatively, the we have the evaluation map ev:Diff0(S)S{\rm{ev}}\colon\operatorname{Diff}_{0}(S)\to S, ev(f)=f(z){\rm{ev}}(f)=f(z), which lifts to a map ev~:Diff0(S)\widetilde{{\rm{ev}}}\colon\operatorname{Diff}_{0}(S)\to\mathbb{H} (given by ev~(f)=f~(z~)\widetilde{{\rm{ev}}}(f)=\widetilde{f}(\widetilde{z}), where again f~\widetilde{f} is the canonical lift), and then Φ\Phi is defined as the descent by id𝒞(S)×ev~{\rm{id}}_{{\mathcal{C}}(S)}\times\widetilde{{\rm{ev}}}:

𝒞(S)×Diff0(S)\textstyle{{\mathcal{C}}(S)\times\operatorname{Diff}_{0}(S)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Φ~\scriptstyle{\widetilde{\Phi}}id𝒞(S)×ev~\scriptstyle{{\rm{id}}_{{\mathcal{C}}(S)}\times\widetilde{{\rm{ev}}}}𝒞(S)×\textstyle{{\mathcal{C}}(S)\times\mathbb{H}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Φ\scriptstyle{\Phi}𝒞s(S˙).\textstyle{{\mathcal{C}}^{s}(\dot{S}).}

Note that every w~\widetilde{w}\in\mathbb{H} is w~=f~(z~)\widetilde{w}=\widetilde{f}(\widetilde{z}) for some fDiff0(S)f\in\operatorname{Diff}_{0}(S) (indeed, ev~\widetilde{{\rm{ev}}} defines a locally trivial fiber bundle). As is shown in [LMS11], Φ(α,w~)\Phi(\alpha,\widetilde{w}) is well-defined independent of the choice of such a diffeomorphism ff with f~(z~)=w~\widetilde{f}(\widetilde{z})=\widetilde{w} since any two differ by an isotopy fixing zz, and Φ\Phi is π1S\pi_{1}S–equivariant (where the points z~\widetilde{z} is used to identify the fundamental group with the group of covering transformations). It is straightforward to see that Φ(α,)\Phi(\alpha,\cdot) is constant on components of p1(Nϵ(α)(α))\mathbb{H}{\smallsetminus}p^{-1}(N_{\epsilon(\alpha)}(\alpha)): two points w~,w~\widetilde{w},\widetilde{w}^{\prime} in such a component are given by w~=f~(z~)\widetilde{w}=\widetilde{f}(\widetilde{z}) and w~=f~(z~)\widetilde{w}^{\prime}=\widetilde{f}^{\prime}(\widetilde{z}) where ff and ff^{\prime} are isotopic by an isotopy ftf_{t}, so that ft(z)f_{t}(z) remains outside Nϵ(α)(α)N_{\epsilon(\alpha)}(\alpha) for all tt.

2.5. Cusps and witnesses

The following lemma relates Φ\Phi to the proper witnesses. Let 𝒫\mathcal{P}\subset\partial\mathbb{H} denote the set of parabolic fixed points. Assume that for each x𝒫x\in\mathcal{P}, we choose a horoball HxH_{x} invariant by the parabolic subgroup Stabπ1S(x){\rm{Stab}}_{\pi_{1}S}(x), the stabilizer of xx in π1S\pi_{1}S. We further assume, as we may, that (1) the union of the horoballs is π1S\pi_{1}S–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 Nϵ(α)(α)N_{\epsilon(\alpha)}(\alpha) for all curves α\alpha. Recall that any proper witness is determined by its boundary.

Lemma 2.7.

There is a π1S\pi_{1}S–equivariant bijection 𝒲:𝒫Ω0(S˙)\mathcal{W}\colon\mathcal{P}\to\Omega_{0}(\dot{S}) determined by

(1) 𝒲(x)=f1(p(Hx)),\partial\mathcal{W}(x)=f^{-1}(\partial p(H_{x})),

for any fDiff0(S)f\in\operatorname{Diff}_{0}(S) with f~(z~)\widetilde{f}(\widetilde{z}) in the interior of the horoball HxH_{x}. Moreover, Φ(𝒞(S)×Hx)=𝒞(𝒲(x))\Phi({\mathcal{C}}(S)\times H_{x})={\mathcal{C}}(\mathcal{W}(x)), we have Φ(Π(u)×Hx))=u\Phi(\Pi(u)\times H_{x}))=u for all u𝒞(𝒲(x))u\in{\mathcal{C}}(\mathcal{W}(x)), and Stabπ1S(x)Stab_{\pi_{1}S}(x), acts trivially on 𝒞(𝒲(x)){\mathcal{C}}(\mathcal{W}(x)).

From the lemma (and as illustrated in the proof) Φ|C(S)×Hx\Phi|_{C(S)\times H_{x}} defines an isomorphism 𝒞(S)𝒞(𝒲(x)){\mathcal{C}}(S)\to{\mathcal{C}}(\mathcal{W}(x)) inverting the isomorphism Π|𝒞(𝒲(x)):𝒞(𝒲(x))𝒞(S)\Pi|_{{\mathcal{C}}(\mathcal{W}(x))}\colon{\mathcal{C}}(\mathcal{W}(x))\to{\mathcal{C}}(S).

Proof.

For any fDiff0(S)f\in\operatorname{Diff}_{0}(S) with w~=f~(z~)Hx\widetilde{w}=\widetilde{f}(\widetilde{z})\in H_{x} and any curve v𝒞0(S)v\in{\mathcal{C}}_{0}(S), we have Φ(v,w~)=Φ~(v,f)=f1(v)\Phi(v,\widetilde{w})=\widetilde{\Phi}(v,f)=f^{-1}(v). On the other hand, f1(p(Hx))f^{-1}(\partial p(H_{x})) is the boundary of a twice punctured disk containing the zz puncture, and hence f1(p(Hx))f^{-1}(\partial p(H_{x})) is the boundary of a witness we denote 𝒲(x)\mathcal{W}(x). Since vv and p(Hx)\partial p(H_{x}) are disjoint,

Φ(v,w~)𝒞(𝒲(x))𝒞s(S˙).\Phi(v,\widetilde{w})\in{\mathcal{C}}(\mathcal{W}(x))\subset{\mathcal{C}^{s}(\dot{S})}.

The same proof that Φ(v,w~)\Phi(v,\widetilde{w}) is well-defined (independent of the choice of fDiff0(S)f\in\operatorname{Diff}_{0}(S) with f~(z~)=w~\widetilde{f}(\widetilde{z})=\widetilde{w}), shows f1(p(Hx))f^{-1}(\partial p(H_{x})) is independent of such a choice of ff (up to isotopy). Therefore, 𝒲\mathcal{W} is well defined by (1). Since v𝒞0(S)v\in{\mathcal{C}}_{0}(S) was arbitrary and Φ(v,)\Phi(v,\cdot) is constant on components of the complement of p1(Nϵ(v)(v))p^{-1}(N_{\epsilon(v)}(v)), we have

Φ(𝒞(S)×Hx)𝒞(𝒲(x)).\Phi({\mathcal{C}}(S)\times H_{x})\subset{\mathcal{C}}(\mathcal{W}(x)).

Given u𝒞(𝒲(x))u\in{\mathcal{C}}(\mathcal{W}(x)), we view uu as a curve disjoint from f1(p(Hx))f^{-1}(\partial p(H_{x})) and hence f(u)f(u) is disjoint from p(Hx)p(H_{x}). There is an isotopy of f(u)f(u) to vv fixing p(Hx)p(H_{x}) (since this is just a neighborhood of the cusp) and hence an isotopy of uu to f1(v)f^{-1}(v) disjoint from f1(p(Hx))f^{-1}(\partial p(H_{x})). This implies Φ({v}×Hx)=u\Phi(\{v\}\times H_{x})=u, proving that Φ(𝒞(S)×Hx)=𝒞(𝒲(x))\Phi({\mathcal{C}}(S)\times H_{x})={\mathcal{C}}(\mathcal{W}(x)), as well as the formula Φ(Π(u)×Hx)=u\Phi(\Pi(u)\times H_{x})=u for all u𝒞(𝒲(x))u\in{\mathcal{C}}(\mathcal{W}(x)).

Next observe that for any proper witness WW, the subcomplex 𝒞(W)𝒞s(S˙){\mathcal{C}}(W)\subset{\mathcal{C}}^{s}(\dot{S}) uniquely determines WW. Therefore, the property that Φ(𝒞(S)×Hx)=𝒞(𝒲(x))\Phi({\mathcal{C}}(S)\times H_{x})={\mathcal{C}}(\mathcal{W}(x)), together with the π1S\pi_{1}S–equivariance of Φ\Phi implies that 𝒲\mathcal{W} is π1S\pi_{1}S–equivariant. All that remains is to show that 𝒲\mathcal{W} is a bijection. Let C1,,CnC_{1},\ldots,C_{n} be the pairwise disjoint horoball cusp neighborhoods of the punctures obtained by projecting the horoballs HxH_{x} for all x𝒫x\in\mathcal{P}.

For any proper witness WW, there is a diffeomorphism f:SSf\colon S\to S, isotopic to the identity by an isotopy ftf_{t} which is the identity on WW for all tt, and so that f(z)Cif(z)\in C_{i}, for some ii. Note that there is an arc connecting zz to the ithi^{th} puncture which is disjoint from both W\partial W and Ci\partial C_{i}. It follows that W\partial W and Ci\partial C_{i} are isotopic, and thus by further isotopy (no longer the identity on WW) we may assume that f(W)=Cif(\partial W)=\partial C_{i}. Therefore, f1(Ci)=Wf^{-1}(\partial C_{i})=\partial W. Observe that the canonical lift f~\tilde{f} has f~(z~)Hx\tilde{f}(\tilde{z})\in H_{x} for some x𝒫x\in\mathcal{P} with p(Hx)=Cip(H_{x})=C_{i}. Therefore, f1(p(Hx))=Wf^{-1}(\partial p(H_{x}))=W, and so 𝒲(x)=W\mathcal{W}(x)=W, so 𝒲\mathcal{W} is surjective.

To see that 𝒲\mathcal{W} is injective, suppose x,y𝒫x,y\in\mathcal{P} are such that 𝒲(x)=𝒲(y)\mathcal{W}(x)=\mathcal{W}(y). The two punctures surrounded by 𝒲(x)\partial\mathcal{W}(x) and by 𝒲(y)\partial\mathcal{W}(y) are therefore the same, hence there exists an element γπ1S\gamma\in\pi_{1}S so that γx=y\gamma\cdot x=y. By π1S\pi_{1}S–equivariance, we must have

γ𝒲(x)=𝒲(γx)=𝒲(y)=𝒲(x).\gamma\cdot\partial\mathcal{W}(x)=\partial\mathcal{W}(\gamma\cdot x)=\partial\mathcal{W}(y)=\partial\mathcal{W}(x).

Choose a representative loop for γ\gamma with minimal self-intersection and denote this γ0\gamma_{0}. If γ0\gamma_{0} is simple closed, then the mapping class associated to γ\gamma is the product of Dehn twists (with opposite signs) in the boundary curves of a regular neighborhood of γ0\gamma_{0}. Otherwise, γ0\gamma_{0} fills a subsurface YS˙Y\subset\dot{S} and is pseudo-Anosov on this subsurface by a result of Kra [Kra81] (see also [KLS09]). It follows that γ𝒲(x)=𝒲(x)\gamma\cdot\partial\mathcal{W}(x)=\partial\mathcal{W}(x) if and only if γ0\gamma_{0} is disjoint from 𝒲(x)\partial\mathcal{W}(x), which happens if and only if f(γ0)p(Hx)f(\gamma_{0})\subset p(H_{x}) (up to isotopy relative to f(z)f(z)). In the action of π1S\pi_{1}S on \mathbb{H}, the element γ\gamma sends f~(z~)\tilde{f}(\tilde{z}) to γf~(z~)\gamma\cdot\tilde{f}(\tilde{z}), and these are the initial and terminal endpoints of the f~\tilde{f}–image of the lift of γ0\gamma_{0} with initial point z~\tilde{z}. On the other hand, f~(z~)Hx\tilde{f}(\tilde{z})\in H_{x}, and hence so is γf~(z~)\gamma\cdot\tilde{f}(\tilde{z}), which means that γ\gamma is fixes xx. Therefore, y=γx=xy=\gamma\cdot x=x, and thus 𝒲\mathcal{W} 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 YY we mean a compact subset of the interior of YY 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 α\alpha is a lamination with exactly one leaf. For a fixed complete, finite area, hyperbolic metric on YY, all geodesic laminations are all contained in a compact subset of the interior of YY. 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 Y\mathcal{L}\subset Y is the image in YY 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 YY as a complementary region (of the preimage of a lamination).

We write 𝒢(Y)\mathcal{G}\mathcal{L}(Y) for the set of laminations on the surface YY, dropping the reference to YY when it is clear from the context. The set of essential simple closed curves, up to isotopy (i.e. the vertex set of 𝒞(Y){\mathcal{C}}(Y)) is thus naturally a subset of 𝒢(Y)\mathcal{G}\mathcal{L}(Y). 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 YY. 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 𝒢\mathcal{G}\mathcal{L} 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

dH(,)=inf{ϵ>0𝒩ϵ() and 𝒩ϵ()}.d_{H}(\mathcal{L},\mathcal{L}^{\prime})=\inf\{\epsilon>0\mid\mathcal{L}\subset\mathcal{N}_{\epsilon}(\mathcal{L}^{\prime})\mbox{ and }\mathcal{L}^{\prime}\subset\mathcal{N}_{\epsilon}(\mathcal{L})\}.

If a sequence of {i}\{\mathcal{L}_{i}\} converges to \mathcal{L} in this topology, we write iH\mathcal{L}_{i}\xrightarrow{\text{H}}\mathcal{L}. The following provides a useful characterization of convergence in this topology; see [CEG06].

Lemma 2.8.

We have iH\mathcal{L}_{i}\xrightarrow{\text{H}}\mathcal{L} if and only if

  1. (1)

    for all xx\in\mathcal{L} there is a sequence of points xiix_{i}\in\mathcal{L}_{i} so that xixx_{i}\to x, and

  2. (2)

    for every subsequence {ik}k=1\{\mathcal{L}_{i_{k}}\}_{k=1}^{\infty}, if xikikx_{i_{k}}\in\mathcal{L}_{i_{k}}, and xikxx_{i_{k}}\to x, then xx\in\mathcal{L}.

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 𝒢\mathcal{G}\mathcal{L} 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 {i}\{\mathcal{L}_{i}\} converges to \mathcal{L} in the coarse Hausdorff topology, then we write iCH\mathcal{L}_{i}\xrightarrow{\text{CH}}\mathcal{L}. The following describes convergence in this topology; see [CEG06].

Lemma 2.9.

We have iCH\mathcal{L}_{i}\xrightarrow{\text{CH}}\mathcal{L} 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 iCH\mathcal{L}_{i}\xrightarrow{\text{CH}}\mathcal{L} if and only if every Hausdorff convergent subsequence converges to a lamination \mathcal{L}^{\prime} containing \mathcal{L}.

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 =(Y)\mathcal{EL}=\mathcal{EL}(Y) denote the space of ending laminations on YY, 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 33–space without accidental parabolics as discussed in the introduction.

A measured lamination is a lamination \mathcal{L} together with an invariant transverse measure μ\mu; 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 (,μ)(\mathcal{L},\mu) is the sublamination |μ||\mu|\subseteq\mathcal{L} with the property that a transverse arc has positive measure if and only if the intersection with |μ||\mu| is nonempty, and is a union of minimal components and simple closed geodesics. We often assume that (,μ)(\mathcal{L},\mu) has full support, meaning =|μ|\mathcal{L}=|\mu|. In this case, we sometimes write μ\mu instead of (,μ)(\mathcal{L},\mu).

The space =(Y)\mathcal{ML}=\mathcal{ML}(Y) of measured laminations on YY 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, (,μ)(\mathcal{L},\mu), we have (|μ|,μ)(|\mu|,\mu) is an element of \mathcal{ML}, and so every measured lamination determines a unique point of \mathcal{ML}. We let \mathcal{FL}\subset\mathcal{ML} denote the subspace of measured laminations whose support is an ending lamination (i.e. it is in \mathcal{EL}). We write 𝒫\mathcal{PML} and 𝒫\mathcal{PFL} for the respective projectivizations of \mathcal{ML} and \mathcal{FL}, 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 𝒫𝒢\mathcal{PML}\to\mathcal{G}\mathcal{L}, given by μ|μ|\mu\mapsto|\mu|, is continuous with respect to the coarse Hausdorff topology on 𝒢\mathcal{G}\mathcal{L}.

For the surface S˙\dot{S}, we consider the subspace

s(S˙):=WΩ(S˙)(W)𝒢(S˙),\mathcal{E}\mathcal{L}^{s}(\dot{S}):=\bigsqcup_{W\in\Omega(\dot{S})}\mathcal{E}\mathcal{L}(W)\subset{\mathcal{G}\mathcal{L}}(\dot{S}),

which is the union of ending laminations of all witnesses of S˙\dot{S}. Similarly, we will write s(S˙)(S˙)\mathcal{FL}^{s}(\dot{S})\subset\mathcal{ML}(\dot{S}) for those measured laminations supported on laminations in s(S˙)\mathcal{E}\mathcal{L}^{s}(\dot{S}), and 𝒫s(S˙)𝒫(S˙)\mathcal{PFL}^{s}(\dot{S})\subset\mathcal{PML}(\dot{S}) for its projectivization.

2.7. Gromov Boundary of a hyperbolic space

A δ\delta–hyperbolic space 𝒳\mathcal{X} can be equipped with a boundary at infinity, 𝒳\partial\mathcal{X} as follows. Given x,y𝒳x,y\in\mathcal{X} and a basepoint o𝒳o\in\mathcal{X}, the Gromov product of xx and yy based at oo is given by

x,yo=12(d(x,o)+d(y,o)d(x,y)).\langle x,y\rangle_{o}=\frac{1}{2}\left(d(x,o)+d(y,o)-d(x,y)\right).

Up to a bounded error (depending only on δ\delta), x,yo\langle x,y\rangle_{o} is the distance from oo to a geodesic connecting xx and yy. The quantity x,yo\langle x,y\rangle_{o} is estimated by the distance from the basepoint oo to a quasi-geodesic between xx and yy. 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 x,y,z𝒳x,y,z\in\mathcal{X},

x,yomin{x,zo,y,zo}\langle x,y\rangle_{o}\succeq\min\{\langle x,z\rangle_{o},\langle y,z\rangle_{o}\}

where the constants in the coarse lower bound depend only on the hyperbolicity constant.

A sequence {xn}𝒳\{x_{n}\}\subset\mathcal{X} is said to converge to infinity if limm,nxm,xno=\displaystyle{\lim_{m,n\rightarrow\infty}\langle x_{m},x_{n}\rangle_{o}=\infty}. Two sequences {xn}\{x_{n}\} and {yn}\{y_{n}\} are equivalent if limm,nym,xn=\displaystyle{\lim_{m,n\rightarrow\infty}\langle y_{m},x_{n}\rangle=\infty}. The points in 𝒳\partial\mathcal{X} are equivalence classes of sequences converging to infinity, and if {xk}x𝒳\{x_{k}\}\in x\in\partial\mathcal{X}, then we say {xk}\{x_{k}\} converges to xx and write xkxx_{k}\to x in 𝒳¯=𝒳𝒳\overline{\mathcal{X}}=\mathcal{X}\cup\partial\mathcal{X}. The topology on the boundary is such that a sequence {xn}nX\{x^{n}\}_{n}\subset\partial X converges to a point xXx\in\partial X if there exist sequences {xkn}k\{x_{k}^{n}\}_{k} representing xnx^{n} for all nn, and {xm}m\{x_{m}\}_{m} representing xx so that

limnlim infk,mxkn,xmo=.\lim_{n\to\infty}\liminf_{k,m\to\infty}\langle x_{k}^{n},x_{m}\rangle_{o}=\infty.

For details see, e.g. [BH99, KB02].

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 \mathcal{FL}\subset\mathcal{ML} 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 \mathcal{EL} with the coarse Hausdorff topology (which for \mathcal{EL} 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 YY equipped with a complete hyperbolic metric of finite area (possibly having geodesic boundary), there is a homeomorphism Y:𝒞(Y)(Y)\mathcal{F}_{Y}\colon\partial{\mathcal{C}}(Y)\to\mathcal{EL}(Y) so that αnx\alpha_{n}\rightarrow x if and only if αnCHY(x)\alpha_{n}\xrightarrow{\text{CH}}\mathcal{F}_{Y}(x).

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 {αn}𝒞0s(S˙)\{\alpha_{n}\}\subset{\mathcal{C}}_{0}^{s}(\dot{S}) and (W)\mathcal{L}\in\mathcal{EL}(W) for some witness WW, then αnCH\alpha_{n}\stackrel{{\scriptstyle CH}}{{\to}}\mathcal{L} if and only if πW(αn)CH\pi_{W}(\alpha_{n})\stackrel{{\scriptstyle CH}}{{\to}}\mathcal{L}.

Note that for each nn, πW(αn)\pi_{W}(\alpha_{n}) is a union of curves, which are not necessarily disjoint. In particular, πW(αn)\pi_{W}(\alpha_{n}) 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 𝒞(W){\mathcal{C}}(W), it has diameter at most 22, and hence if an,anπW(αn)a_{n},a_{n}^{\prime}\subset\pi_{W}(\alpha_{n}) are any two curves, for each nn, and (W)\mathcal{L}\in\mathcal{EL}(W), then ana_{n} and ana_{n}^{\prime} either both coarse Hausdorff converge to \mathcal{L} or neither does (by Theorem 2.12). Consequently, it makes sense to say that πW(αn)\pi_{W}(\alpha_{n}) coarse Hausdorff converges to a lamination in (W)\mathcal{EL}(W).

Proof.

For the rest of this proof we fix a complete hyperbolic metric on S˙\dot{S} and realize WS˙W\subset\dot{S} as an embedded subsurface with geodesic boundary. Let us first assume πW(αn)CH(W)\pi_{W}(\alpha_{n})\stackrel{{\scriptstyle CH}}{{\to}}\mathcal{L}\in\mathcal{EL}(W). After passing to an arbitrary convergent subsequence, we may assume αnH\alpha_{n}\stackrel{{\scriptstyle H}}{{\to}}{\mathcal{L}^{\prime}}. It suffices to show that \mathcal{L}\subset\mathcal{L}^{\prime}.

Let n1,,nrαnW\ell_{n}^{1},\ldots,\ell_{n}^{r}\subset\alpha_{n}\cap W be the decomposition into isotopy classes of arcs of intersection: that is, each nj\ell_{n}^{j} is a union of all arcs of intersection of αn\alpha_{n} with WW so that any two arcs of αnW\alpha_{n}\cap W are isotopic if and only if they are contained in the same set nj\ell_{n}^{j} (we may have to pass to a further subsequence so that each intersection αnW\alpha_{n}\cap W consists of the same number rr of isotopy classes, which we do). For each nj\ell_{n}^{j}, let αnjπW(αn)\alpha_{n}^{j}\subset\pi_{W}(\alpha_{n}) be the geodesic multi-curve produced from the isotopy class nj\ell_{n}^{j} by surgery in the definition of projection. Note that αnj\alpha_{n}^{j} and nj\ell_{n}^{j} have no transverse intersections. Pass to a further subsequence so that αnjHj\alpha_{n}^{j}\stackrel{{\scriptstyle H}}{{\to}}\mathcal{L}^{j} and njHj\ell_{n}^{j}\stackrel{{\scriptstyle H}}{{\to}}\mathcal{L}_{j}^{\prime}; here, each nj\ell_{n}^{j} is a compact subset of WW so Hausdorff convergence to a closed set still makes sense, though j\mathcal{L}_{j}^{\prime} are not necessarily geodesic laminations. By Corollary 2.10 (and the discussion in the paragraph preceding this proof), j\mathcal{L}\subset\mathcal{L}^{j}, for each jj. Appealing to Lemma 2.8, it easily follows that W=1r\mathcal{L}^{\prime}\cap W=\mathcal{L}_{1}^{\prime}\cup\cdots\cup\mathcal{L}_{r}^{\prime}. Since anja_{n}^{j} has no transverse intersections with nj\ell_{n}^{j}, j\mathcal{L}_{j}^{\prime} has no transverse intersections with j\mathcal{L}^{j}, for each jj. Therefore, \mathcal{L} has no transverse intersections with W\mathcal{L}^{\prime}\cap W, and since W\mathcal{L}\subset W, \mathcal{L}^{\prime} has no transverse intersections with \mathcal{L}. Since (W)\mathcal{L}\in\mathcal{EL}(W), it follows that \mathcal{L}\subset\mathcal{L}^{\prime}, as required.

Now in the opposite direction we assume that αnCH(W)\alpha_{n}\stackrel{{\scriptstyle CH}}{{\to}}\mathcal{L}\in\mathcal{EL}(W). Let n1,,nrαnW\ell_{n}^{1},\ldots,\ell_{n}^{r}\subset\alpha_{n}\cap W and αn1,,αnrπW(αn)\alpha_{n}^{1},\ldots,\alpha_{n}^{r}\subset\pi_{W}(\alpha_{n}) be as above, so that for each jj (after passing to a subsequence) we have

njHj and αnjHj.\ell_{n}^{j}\stackrel{{\scriptstyle H}}{{\to}}\mathcal{L}^{j}\quad\mbox{ and }\quad\alpha_{n}^{j}\stackrel{{\scriptstyle H}}{{\to}}\mathcal{L}_{j}^{\prime}.

Similar to the above, 1r\mathcal{L}\subset\mathcal{L}^{1}\cup\cdots\cup\mathcal{L}^{r} and since nj\ell_{n}^{j} has no transverse intersections with αnj\alpha_{n}^{j}, j\mathcal{L}_{j}^{\prime} has no transverse intersections with \mathcal{L}. Since \mathcal{L} is an ending lamination, j\mathcal{L}\subset\mathcal{L}_{j}^{\prime}. Since the convergent subsequence was arbitrary, it follows that πW(αn)CH\pi_{W}(\alpha_{n})\stackrel{{\scriptstyle CH}}{{\to}}\mathcal{L}. ∎

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 𝒞s(S˙){\mathcal{C}}^{s}(\dot{S}), 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 WS˙W\subseteq\dot{S} and x,y𝒞(W)x,y\in{\mathcal{C}}(W), let [x,y]W𝒞(W)[x,y]_{W}\subset{\mathcal{C}}(W) denote a geodesic between xx and yy.

The following definition is reminiscent of hierarchy paths from [MM00], though our situation is considerably simpler.

Definition 3.1.

Given x,y𝒞s(S˙)x,y\in{\mathcal{C}}^{s}(\dot{S}), let [x,y]S˙[x,y]_{\dot{S}} be any 𝒞(S˙){\mathcal{C}}(\dot{S})–geodesic. If WS˙W\subsetneq\dot{S} is a proper witness such that W\partial W is a vertex of [x,y]S˙[x,y]_{\dot{S}}, we say that WW is a witness for [x,y]S˙[x,y]_{\dot{S}}. Note that if WW is a witness for [x,y]S˙[x,y]_{\dot{S}}, then the immediate predecessor and successor x,yx^{\prime},y^{\prime} to W\partial W in [x,y]S˙[x,y]_{\dot{S}} are necessarily contained in 𝒞(W){\mathcal{C}}(W) (hence also in 𝒞s(S˙){\mathcal{C}}^{s}(\dot{S})) and we let [x,y]W𝒞(W)[x^{\prime},y^{\prime}]_{W}\subset{\mathcal{C}}(W) be a geodesic (which we also view as a path in 𝒞s(S˙){\mathcal{C}}^{s}(\dot{S})). Replacing every consecutive triple x,W,y[x,y]S˙x^{\prime},\partial W,y^{\prime}\subset[x,y]_{\dot{S}} with the path [x,y]W[x^{\prime},y^{\prime}]_{W} produces a path from xx to yy in 𝒞s(S˙){\mathcal{C}}^{s}(\dot{S}) which we call a survival path from xx to yy, and denote it σ(x,y)\sigma(x,y). We call [x,y]S˙[x,y]_{\dot{S}} the main geodesic of σ(x,y)\sigma(x,y) and, if WW is witness for [x,y]S˙[x,y]_{\dot{S}}, we call the corresponding 𝒞(W){\mathcal{C}}(W)–geodesic [x,y]W[x^{\prime},y^{\prime}]_{W} the witness geodesic of σ(x,y)\sigma(x,y) for WW, and also say that WW is a witness for σ(x,y)\sigma(x,y).

An immediate corollary of Theorem 2.6, we have

Corollary 3.2.

For any x,y𝒞s(S˙)x,y\in{\mathcal{C}}^{s}(\dot{S}) and proper witness WW, if dW(x,y)>Md_{W}(x,y)>M, then WW is a witness for [x,y]S˙[x,y]_{\dot{S}}, for any geodesics [x,y]S˙[x,y]_{\dot{S}} between xx and yy.

Proof.

Since dW(x,y)>Md_{W}(x,y)>M, it follows by Theorem 2.6 that some vertex of [x,y]S˙[x,y]_{\dot{S}} has empty projection to WW. But the only multi-curve in 𝒞(S˙){\mathcal{C}}(\dot{S}) with empty projection to WW is W\partial W, hence W\partial W is a vertex of [x,y]S˙[x,y]_{\dot{S}}. ∎

No two consecutive vertices of [x,y]S˙[x,y]_{\dot{S}} can be boundaries of witness (since any two such boundaries nontrivially intersect). Therefore, the next lemma follows.

Lemma 3.3.

For any x,y𝒞s(S˙)x,y\in{\mathcal{C}^{s}(\dot{S})} and geodesics [x,y]S˙[x,y]_{\dot{S}}, there are at most dS˙(x,y)2\tfrac{d_{\dot{S}}(x,y)}{2} witnesses for [x,y]S˙[x,y]_{\dot{S}}. ∎

The following lemma estimates the lengths of witness geodesics on a survival path.

Lemma 3.4.

Given a survival path σ(x,y)\sigma(x,y) and a witness WW for σ(x,y)\sigma(x,y), the initial and terminal vertices xx^{\prime} and yy^{\prime} of the witness geodesic segment [x,y]W[x^{\prime},y^{\prime}]_{W} satisfy

dW(x,x),dW(y,y)<M.d_{W}(x,x^{\prime}),d_{W}(y,y^{\prime})<M.

Consequently, dW(x,y)d_{W}(x^{\prime},y^{\prime}) of satisfies

dW(x,y)2M,0dW(x,y).d_{W}(x^{\prime},y^{\prime})\stackrel{{\scriptstyle\mbox{\tiny$2M\!,\!0$}}}{{\asymp}}d_{W}(x,y).
Proof.

By Theorem 2.6 applied to the subsegments of [x,y]S˙[x,y]_{\dot{S}} from xx to xx^{\prime} and yy^{\prime} to yy 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 x,y𝒞s(S˙)x,y\in{\mathcal{C}}^{s}(\dot{S}) and k>Mk>M, we have

ds(x,y)2k2+2kWΩ(S˙){{dW(x,y)}}k.d^{s}(x,y)\leq 2k^{2}+2k\sum_{W\in\Omega(\dot{S})}{\{\{}d_{W}(x,y){\}\}}_{k}.

Recall that Ω(S˙)\Omega(\dot{S}) denotes the set of all witnesses for 𝒞s(S˙){\mathcal{C}}^{s}(\dot{S}) and that {{x}}k{\{\{}x{\}\}}_{k} is the cut-off function giving value xx if xkx\geq k and 0 otherwise.

Proof.

Since σ(x,y)\sigma(x,y) is a path from xx to yy, it suffices to prove that the length of σ(x,y)\sigma(x,y) is bounded above by the right-hand side. For each witness WW of x,yx,y whose boundary appears in [x,y]S˙[x,y]_{\dot{S}}, we have replaced the length two segment {x,W,y}\{x^{\prime},\partial W,y^{\prime}\} with [x,y]W[x^{\prime},y^{\prime}]_{W}, which has length dW(x,y)d_{W}(x^{\prime},y^{\prime}). By Lemma 3.4 we have

dW(x,y)2M+dW(x,y).d_{W}(x^{\prime},y^{\prime})\leq 2M+d_{W}(x,y).

If dW(x,y)k>Md_{W}(x,y)\geq k>M, this implies the length dW(x,y)d_{W}(x^{\prime},y^{\prime}), of [x,y]W[x^{\prime},y^{\prime}]_{W} is less than 3dW(x,y)3d_{W}(x,y). Otherwise, the length is less than 3k3k. Let W1,,WnW_{1},\ldots,W_{n} denote the witnesses for x,yx,y whose boundaries appear in [x,y]S˙[x,y]_{\dot{S}}. By Lemma 3.3, n12dS˙(x,y)n\leq\tfrac{1}{2}d_{\dot{S}}(x,y), half the length of [x,y]S˙[x,y]_{\dot{S}}. Further note that by Corollary 3.2, if dW(x,y)k>Md_{W}(x,y)\geq k>M, then WW is one of the witnesses WjW_{j}, for some jj.

Combining all of these (and fact that k>M>2k>M>2) we obtain the following bound on the length of σ(x,y)\sigma(x,y), and hence ds(x,y)d^{s}(x,y):

ds(x,y)dS˙(x,y)+j=1n3dWj(x,y)dS˙(x,y)+3j=1n({{dWj(x,y)}}k+k)=dS˙(x,y)+3nk+3j=1n{{dWj(x,y)}}k(3k2+1)dS˙(x,y)+3j=1n{{dWj(x,y)}}k2k({{dS˙(x,y)}}k+k)+3j=1n{{dWj(x,y)}}k2k2+2kWΩ(S˙){{dW(x,y)}}k\begin{array}[]{rclcl}d^{s}(x,y)&\leq&\displaystyle{d_{\dot{S}}(x,y)+\sum_{j=1}^{n}3d_{W_{j}}(x,y)}\leq\displaystyle{d_{\dot{S}}(x,y)+3\sum_{j=1}^{n}({\{\{}d_{W_{j}}(x,y){\}\}}_{k}+k)}\\ &=&\displaystyle{d_{\dot{S}}(x,y)+3nk+3\sum_{j=1}^{n}{\{\{}d_{W_{j}}(x,y){\}\}}_{k}}\leq\displaystyle{(\tfrac{3k}{2}+1)d_{\dot{S}}(x,y)+3\sum_{j=1}^{n}{\{\{}d_{W_{j}}(x,y){\}\}}_{k}}\\ &\leq&\displaystyle{2k({\{\{}d_{\dot{S}}(x,y){\}\}}_{k}+k)+3\sum_{j=1}^{n}{\{\{}d_{W_{j}}(x,y){\}\}}_{k}}\leq\displaystyle{2k^{2}+2k\sum_{W\in\Omega(\dot{S})}{\{\{}d_{W}(x,y){\}\}}_{k}}\\ \end{array}

Lemma 3.6.

Given x,y𝒞s(S˙)x,y\in{\mathcal{C}^{s}(\dot{S})}, if WW is not a witness for [x,y]S˙[x,y]_{\dot{S}}, then

diamW(σ(x,y))M+4.\operatorname{diam}_{W}(\sigma(x,y))\leq M+4.
Proof.

Since WW is not a witness for [x,y]S˙[x,y]_{\dot{S}}, every z[x,y]S˙z\in[x,y]_{\dot{S}} has non-empty projection to WW. Therefore, diamW([x,y])S˙M\operatorname{diam}_{W}([x,y])_{\dot{S}}\leq M by Theorem 2.6. If w𝒞(W)w^{\prime}\in{\mathcal{C}}(W^{\prime}) is on a witness geodesic segment of σ(x,y)\sigma(x,y), then dS˙(w,W)=1d_{\dot{S}}(w^{\prime},\partial W^{\prime})=1 so dW(w,W)2d_{W}(w^{\prime},\partial W^{\prime})\leq 2 by Proposition 2.5. Since W[x,y]S˙\partial W^{\prime}\in[x,y]_{\dot{S}}, the lemma follows by the triangle inequality. ∎

Lemma 3.7.

Suppose σ(x,y)\sigma(x,y) is a survival path and x,yσ(x,y)x^{\prime},y^{\prime}\in\sigma(x,y) with xx<yyx\leq x^{\prime}<y^{\prime}\leq y, with respect to the ordering from σ(x,y)\sigma(x,y). Then if x,yx^{\prime},y^{\prime} lie on the main geodesic [x,y]S˙[x,y]_{\dot{S}}, then the subpath of σ(x,y)\sigma(x,y) from xx^{\prime} to yy^{\prime} is a survival path.

If x𝒞(W)x^{\prime}\in{\mathcal{C}}(W) and/or y𝒞(W)y^{\prime}\in{\mathcal{C}}(W^{\prime}) for proper witnesses W,WW,W^{\prime} for x,yx,y, respectively, then the same conclusion holds, provided the subsegments of 𝒞(W){\mathcal{C}}(W) and/or 𝒞(W){\mathcal{C}}(W^{\prime}) in σ(x,y)\sigma(x,y) between xx^{\prime} and yy^{\prime} has length at least 2M2M.

Proof.

When x,yx^{\prime},y^{\prime} are on the main geodesic, this is straightforward, since in this case, the subsegment of the main geodesic between xx^{\prime} and yy^{\prime} serves as the main geodesic for a survival path between xx^{\prime} and yy^{\prime}.

There are several cases for the second statement. The proofs are all similar, so we just describe one case where, say, x[x′′,y′′]W𝒞(W)x^{\prime}\in[x^{\prime\prime},y^{\prime\prime}]_{W}\subset{\mathcal{C}}(W) with xx′′xy′′yyx\leq x^{\prime\prime}\leq x^{\prime}\leq y^{\prime\prime}\leq y^{\prime}\leq y, and yy^{\prime} is in the main geodesic. The assumption in this case means that in 𝒞(W){\mathcal{C}}(W), the distance between xx^{\prime} and y′′y^{\prime\prime} is at least 2M2M. Lemma 3.4 implies that dW(y′′,y)<Md_{W}(y^{\prime\prime},y)<M, and so by the triangle inequality, dW(x,y)>Md_{W}(x^{\prime},y)>M. Therefore, by Theorem 2.6 any geodesic from xx^{\prime} to yy must pass through W\partial W. In particular, the path that starts at xx^{\prime}, travels to W\partial W, then continues along the subsegment of [x,y]S˙[x,y]_{\dot{S}} from W\partial W to yy^{\prime}, is a geodesic in 𝒞(S˙){\mathcal{C}}(\dot{S}). We can easily build a survival path from xx^{\prime} to yy^{\prime} using this geodesic that is a subsegment of σ(x,y)\sigma(x,y), as required. The other cases are similar. ∎

3.1. Infinite survival paths

Masur-Minsky proved that for any surface ZZ and any two points in 𝒞¯(Z)=𝒞(Z)𝒞(Z)\bar{\mathcal{C}}(Z)={\mathcal{C}}(Z)\cup\partial{\mathcal{C}}(Z), where 𝒞(Z)\partial{\mathcal{C}}(Z) is the Gromov boundary, there is a geodesic ``connecting" these points; see [MM00]. Given x,y𝒞¯(Z)x,y\in\bar{\mathcal{C}}(Z), we let [x,y]Z[x,y]_{Z} denote such a geodesic.

The construction of survival paths above can be carried out for geodesic lines and rays in 𝒞(S˙){\mathcal{C}}(\dot{S}), replacing any length two path x,W,yx^{\prime},\partial W,y^{\prime} with a 𝒞(W){\mathcal{C}}(W) geodesic from xx^{\prime} to yy^{\prime} to produce a survival ray or survival line, respectively. More generally, to a geodesic segment or ray of 𝒞(S˙){\mathcal{C}}(\dot{S}) 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 WW and WW^{\prime} and points z,zz,z^{\prime} in the Gromov boundaries of 𝒞(W){\mathcal{C}}(W) and 𝒞(W){\mathcal{C}}(W^{\prime}), respectively, we can construct a survival line starts and ends with geodesic rays in 𝒞(W){\mathcal{C}}(W) and 𝒞(W){\mathcal{C}}(W^{\prime}), limiting to zz and zz^{\prime}, 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

z,zWΩ(S˙)𝒞¯(W),z,z^{\prime}\in\bigcup_{W\in\Omega(\dot{S})}\bar{\mathcal{C}}(W),

and we denote such by σ(z,z)\sigma(z,z^{\prime}), as in the finite case. From this discussion, we have the following.

Lemma 3.8.

For any distinct pair of elements

z,zWΩ(S˙)𝒞¯(W)z,z^{\prime}\in\bigcup_{W\in\Omega(\dot{S})}\bar{\mathcal{C}}(W)

there exists a (possibly infinite) survival path σ(z,z)\sigma(z,z^{\prime}) ``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 𝒞s(S˙){\mathcal{C}^{s}(\dot{S})} 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 zz, 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.

The condition for hyperbolicity we use is the following; see [MS13, Bow14].

Proposition 4.3.

Given ϵ>0\epsilon>0, there exists δ>0\delta>0 with the following property. Suppose that GG is a connected graph and for each x,yV(G)x,y\in V(G) there is an associated connected subgraph ς(x,y)G\varsigma(x,y)\subseteq G including x,yx,y. Suppose that,

  1. (1)

    For all x,y,zV(G)x,y,z\in V(G),

    ς(x,y)𝒩ϵ(ς(x,z)ς(z,y))\varsigma(x,y)\subseteq\mathcal{N}_{\epsilon}(\varsigma(x,z)\cup\varsigma(z,y))
  2. (2)

    For any x,yV(G)x,y\in V(G) with d(x,y)1d(x,y)\leq 1, the diameter of ς(x,y)\varsigma(x,y) in GG is at most ϵ\epsilon.

Then, GG is δ\delta–hyperbolic.

We will apply Proposition 4.3 to the graph 𝒞s(S˙){\mathcal{C}}^{s}(\dot{S}), and for vertices x,y𝒞s(S˙)x,y\in{\mathcal{C}}^{s}(\dot{S}), the required subcomplex is a (choice of some) survival path σ(x,y)\sigma(x,y). Note that if x,yx,y are distance one apart, then σ(x,y)=[x,y]\sigma(x,y)=[x,y], which has diameter 11. Therefore, as long as ϵ1\epsilon\geq 1, condition (2) in Theorem 4.3 will be satisfied. We therefore focus on condition (1), and express this briefly by saying that x,y,zx,y,z span an ϵ\epsilon–slim survival triangle. The next lemma verifies condition (1) in a special case.

Lemma 4.4.

Given R>4R>4, there exists ϵ>0\epsilon>0 with the following property. If x,y,z𝒞s(S˙)x,y,z\in{\mathcal{C}}^{s}(\dot{S}) are any three points such that dW(u,v)Rd_{W}(u,v)\leq R for all proper witness WS˙W\subsetneq\dot{S} and every u,v{x,y,z}u,v\in\{x,y,z\}, then x,y,zx,y,z span an ϵ\epsilon–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 R+2MR+2M; we will use this fact throughout the proof without further mention. We also observe that by Theorem 2.6, for any wσ(x,y)[x,y]S˙w\in\sigma(x,y)\cap[x,y]_{\dot{S}} and any proper witness WS˙W\subsetneq\dot{S}, at least one of dW(x,w)d_{W}(x,w) or dW(y,w)d_{W}(y,w) is at most MM.

Next suppose ww is on a subsegment [x,y]Wσ(x,y)[x^{\prime},y^{\prime}]_{W}\subset\sigma(x,y) for some proper witness WW of σ(x,y)\sigma(x,y). Observe that ww is within R+2M2=R2+M\tfrac{R+2M}{2}=\tfrac{R}{2}+M from either xx^{\prime} or yy^{\prime} and so by Theorem 2.6 and the triangle inequality, one of dW(w,x)d_{W}(w,x) or dW(w,y)d_{W}(w,y) is at most R2+2M\tfrac{R}{2}+2M. If WW^{\prime} is any other proper witness, we claim that dW(x,w)d_{W^{\prime}}(x,w) or dW(y,w)d_{W^{\prime}}(y,w) is at most M+2R2+2MM+2\leq\frac{R}{2}+2M. To see this, note that either W\partial W^{\prime} lies in [x,W]S˙[x,y]S˙[x,\partial W]_{\dot{S}}\subset[x,y]_{\dot{S}}, in [W,y]S˙[x,y]S˙[\partial W,y]_{\dot{S}}\subset[x,y]_{\dot{S}}, or neither. In the first two cases, dW(W,y)Md_{W^{\prime}}(\partial W,y)\leq M or dW(W,x)Md_{W^{\prime}}(\partial W,x)\leq M, respectively, by Theorem 2.6, while in the third case both of these inequalities hold. Therefore, since ww and W\partial W are disjoint, dW(W,w)2d_{W^{\prime}}(\partial W,w)\leq 2, and hence dW(x,w)d_{W^{\prime}}(x,w) or dW(y,w)d_{W^{\prime}}(y,w) is at most M+2R2+2MM+2\leq\frac{R}{2}+2M.

Now let wσ(x,y)w\in\sigma(x,y) be any vertex and w0[x,y]S˙σ(x,y)w_{0}\in[x,y]_{\dot{S}}\cap\sigma(x,y) the nearest vertex along σ(x,y)\sigma(x,y), and observe that dS˙(w,w0)2d_{\dot{S}}(w,w_{0})\leq 2. Since 𝒞(S˙){\mathcal{C}}(\dot{S}) is δ\delta–hyperbolic (for some δ>0\delta>0), there is a vertex w0[x,z]S˙[y,z]S˙w_{0}^{\prime}\in[x,z]_{\dot{S}}\cup[y,z]_{\dot{S}} with dS˙(w0,w0)δd_{\dot{S}}(w_{0},w_{0}^{\prime})\leq\delta. Without loss of generality, we assume w0[x,z]S˙w_{0}^{\prime}\in[x,z]_{\dot{S}}. Choose wσ(x,z)w^{\prime}\in\sigma(x,z) to be w=w0w^{\prime}=w_{0}^{\prime} if w0σ(x,z)w_{0}^{\prime}\in\sigma(x,z) or one of the adjacent vertices of [x,z]S˙[x,z]_{\dot{S}} if w0w_{0}^{\prime} is the boundary of a witness. Then dS˙(w0,w)1d_{\dot{S}}(w_{0}^{\prime},w^{\prime})\leq 1, so

dS˙(w,w)δ+3.d_{\dot{S}}(w,w^{\prime})\leq\delta+3.

Now suppose WS˙W\subsetneq\dot{S} is a proper witness. Then at least one of dW(w,x)d_{W}(w,x) or dW(w,y)d_{W}(w,y) is at most R2+2M\tfrac{R}{2}+2M as is at least one of dW(w,x)d_{W}(w^{\prime},x) or dW(w,z)d_{W}(w^{\prime},z). If dW(w,x),dW(w,x)R2+2Md_{W}(w,x),d_{W}(w^{\prime},x)\leq\frac{R}{2}+2M, then applying the triangle inequality, we see that

dW(w,w)R+4M.d_{W}(w,w^{\prime})\leq R+4M.

If instead, dW(x,w)R2+2Md_{W}(x,w)\leq\frac{R}{2}+2M and dW(w,z)R2+2Md_{W}(w^{\prime},z)\leq\frac{R}{2}+2M, then the triangle inequality implies

dW(w,w)dW(w,x)+dW(x,z)+dW(z,w)2R+4M.d_{W}(w,w^{\prime})\leq d_{W}(w,x)+d_{W}(x,z)+d_{W}(z,w^{\prime})\leq 2R+4M.

The other two possibilities are similar, and hence dW(w,w)2R+4Md_{W}(w,w^{\prime})\leq 2R+4M.

Applying Corollary 3.5 with k=Mk=M, recalling that by Lemma 3.3 there are at most dS˙(w,w)2δ+32\tfrac{d_{\dot{S}}(w,w^{\prime})}{2}\leq\frac{\delta+3}{2} proper witnesses for any geodesic [w,w]S˙[w,w^{\prime}]_{\dot{S}}, we have

ds(w,w)2M2+2MWΩ(S˙){{dW(w,w)}}M2M2+2M(δ+3+δ+32(2R+4M)).d^{s}(w,w^{\prime})\leq 2M^{2}+2M\sum_{W\in\Omega(\dot{S})}{\{\{}d_{W}(w,w^{\prime}){\}\}}_{M}\leq 2M^{2}+2M\left(\delta+3+\tfrac{\delta+3}{2}(2R+4M)\right).

Setting ϵ\epsilon equal to the right-hand side (which really depends only on RR, since MM and δ\delta are independent of anything), completes the proof. ∎

A standard argument subdividing an nn–gon into triangles proves the following.

Corollary 4.5.

Given R>0R>0 let ϵ>0\epsilon>0 be as in Lemma 4.4. If n3n\geq 3 and x1,,xn𝒞s(S˙)x_{1},\ldots,x_{n}\in{\mathcal{C}}^{s}(\dot{S}) are such that dW(xi,xj)Rd_{W}(x_{i},x_{j})\leq R for all 1i,jn1\leq i,j\leq n, then for all wσ(xi,xi+1)w\in\sigma(x_{i},x_{i+1}), there exists jij\neq i and wσ(xj,xj+1)w^{\prime}\in\sigma(x_{j},x_{j+1}) (with all indices taken modulo nn) such that ds(w,w)n2ϵd^{s}(w,w^{\prime})\leq\lceil\tfrac{n}{2}\rceil\epsilon.

For the remainder of the proof (and elsewhere in the paper) it is useful to make the following definition.

Definition 4.6.

Given x,y,z𝒞s(S˙)x,y,z\in{\mathcal{C}}^{s}(\dot{S}) and R>0R>0, consider the proper witnesses with projection at least RR:

ΩR(x,y)={WΩ0(S˙)dW(x,y)>R},\Omega_{R}(x,y)=\{W\in\Omega_{0}(\dot{S})\mid d_{W}(x,y)>R\},

and set

ΩR(x,y,z)=ΩR(x,y)ΩR(x,z)ΩR(y,z).\Omega_{R}(x,y,z)=\Omega_{R}(x,y)\cup\Omega_{R}(x,z)\cup\Omega_{R}(y,z).

In words, ΩR(x,y)\Omega_{R}(x,y) is the set of all proper witness for which xx and yy have distance greater than RR.

Lemma 4.7.

For any three points x,y,z𝒞s(S˙)x,y,z\in{\mathcal{C}}^{s}(\dot{S}) and R2MR\geq 2M, there is at most one WΩR(x,y,z)W\in\Omega_{R}(x,y,z) such that

WΩR/2(x,y)ΩR/2(x,z)ΩR/2(y,z).W\in\Omega_{R/2}(x,y)\cap\Omega_{R/2}(x,z)\cap\Omega_{R/2}(y,z).
Proof.

Suppose there exist two distinct

W,WΩR/2(x,y)ΩR/2(x,z)ΩR/2(y,z).W,W^{\prime}\in\Omega_{R/2}(x,y)\cap\Omega_{R/2}(x,z)\cap\Omega_{R/2}(y,z).

Then by Theorem 2.6, W,W\partial W,\partial W^{\prime} are (distinct) vertices in any 𝒞(S˙){\mathcal{C}}(\dot{S})–geodesic between any two vertices in {x,y,z}\{x,y,z\}. Choose geodesics [x,W]S˙[x,\partial W]_{\dot{S}}, [y,W]S˙[y,\partial W]_{\dot{S}}, and [z,W]S˙[z,\partial W]_{\dot{S}}, and note that concatenating any two of these (with appropriate orientations) produces a geodesic between a pair of vertices in {x,y,z}\{x,y,z\}. Since WW\partial W^{\prime}\neq\partial W must also lie on all 𝒞(S˙){\mathcal{C}}(\dot{S})–geodesics between these three vertices, it must lie on at least one of the geodesic segments to W\partial W; without loss of generality, suppose W[x,W]S˙\partial W^{\prime}\in[x,\partial W]_{\dot{S}}. If W\partial W^{\prime} is not a vertex of either [y,W]S˙[y,\partial W]_{\dot{S}} or [z,W]S˙[z,\partial W]_{\dot{S}}, then our geodesic from yy to zz does not contain W\partial W^{\prime}, a contradiction. Without loss of generality, we may assume W[y,W]S˙\partial W^{\prime}\in[y,\partial W]_{\dot{S}}. But then the geodesic subsegment between xx and W\partial W^{\prime} in [x,W]S˙[x,\partial W]_{\dot{S}} together with the geodesic subsegment between W\partial W^{\prime} and yy in [y,W]S˙[y,\partial W]_{\dot{S}} is also a geodesics (as above) and does not pass through W\partial W, a contradiction. ∎

Proof of Theorem 4.1.

Let x,y,z𝒞s(S˙)x,y,z\in{\mathcal{C}}^{s}(\dot{S}). By the triangle inequality, if WΩ2M(x,y)W\in\Omega_{2M}(x,y), then at least one of dW(x,z)d_{W}(x,z) or dW(y,z)d_{W}(y,z) is greater than MM. By Lemma 4.7, there is at most one WW such that both are greater than MM. If such WW exists, denote it W0W_{0} and write D0={W0}D_{0}=\{W_{0}\}; otherwise, write D0=D_{0}=\emptyset. Defining

Dx={WΩ2M(x,y,z)D0dW(x,y)>M,dW(x,z)>M}D_{x}=\{W\in\Omega_{2M}(x,y,z){\smallsetminus}D_{0}\mid d_{W}(x,y)>M,d_{W}(x,z)>M\}

(and defining DyD_{y}, DzD_{z} similarly), we can express Ω2M(x,y,z)\Omega_{2M}(x,y,z) as a disjoint union

Ω2M(x,y,z)=DxDyDzD0.\Omega_{2M}(x,y,z)=D_{x}\sqcup D_{y}\sqcup D_{z}\sqcup D_{0}.

By Theorem 2.6, the 𝒞(S˙){\mathcal{C}}(\dot{S})–geodesics [x,y]S˙[x,y]_{\dot{S}} and [x,z]S˙[x,z]_{\dot{S}} contain W\partial W for all WDxW\in D_{x}, and we write

Dx={Wx1,Wx2,,Wxmx}D_{x}=\{W_{x}^{1},W_{x}^{2},\ldots,W_{x}^{m_{x}}\}

so that x1=Wx1,x2=Wx2,,xmx=Wxmxx_{1}=\partial W_{x}^{1},x_{2}=\partial W_{x}^{2},\ldots,x_{m_{x}}=\partial W_{x}^{m_{x}} appear in this order along [x,y]S˙[x,y]_{\dot{S}} and [x,z]S˙[x,z]_{\dot{S}}. Similarly write

Dy={Wy1,,Wymy} and Dz={Wz1,,Wzmz}.D_{y}=\{W_{y}^{1},\ldots,W_{y}^{m_{y}}\}\mbox{ and }D_{z}=\{W_{z}^{1},\ldots,W_{z}^{m_{z}}\}.

The 𝒞(S˙){\mathcal{C}}(\dot{S})–geodesic triangle between xx, yy, and zz must appear as in the examples illustrated in Figure 1.

xxx1x_{1}x2x_{2}x3x_{3}y2y_{2}y1y_{1}yyz1z_{1}zzxxx1x_{1}x2x_{2}W0\partial W_{0}y1y_{1}yyz1z_{1}zz
Figure 1. Geodesic triangles in 𝒞(S˙){\mathcal{C}}(\dot{S}): Here xj=Wxjx_{j}=\partial W_{x}^{j}, yj=Wyjy_{j}=\partial W_{y}^{j}, and zj=Wzjz_{j}=\partial W_{z}^{j}, and {x1,x2,x3}[x,y]S˙[x,z]S˙\{x_{1},x_{2},x_{3}\}\subset[x,y]_{\dot{S}}\cap[x,z]_{\dot{S}}, {y1,y2}[x,y]S˙[y,z]S˙\{y_{1},y_{2}\}\subset[x,y]_{\dot{S}}\cap[y,z]_{\dot{S}}, and {z1}[x,z]S˙[y,z]S˙\{z_{1}\}\subset[x,z]_{\dot{S}}\cap[y,z]_{\dot{S}}. The left triangle has D0=D_{0}=\emptyset, while the triangle on the right has D0={W0}D_{0}=\{W_{0}\}, hence W0[x,y]S˙[x,z]S˙[y,z]S˙\partial W_{0}\in[x,y]_{\dot{S}}\cap[x,z]_{\dot{S}}\cap[y,z]_{\dot{S}}. Note: there may be more vertices in common to pairs of geodesics than the vertices xj,yj,zjx_{j},y_{j},z_{j}. Furthermore, there may be various degenerations, e.g. Dx=D0=D_{x}=D_{0}=\emptyset, in which case the three bigons in the upper left-hand portion of the left figure disappears and x3x_{3} becomes xx.

We now subdivide each of the survival paths σ(x,y)\sigma(x,y), σ(x,z)\sigma(x,z), and σ(y,z)\sigma(y,z) into subsegments as follows. In this subdivision, σ(x,y)\sigma(x,y) is a concatenation of witness geodesics for each witness WW in DxDyD0D_{x}\cup D_{y}\cup D_{0} and complementary subsegments connecting consecutive such witness geodesics. The complementary segments are themselves survival paths obtained as concatenations of 𝒞(S˙){\mathcal{C}}(\dot{S})–geodesic segments and witness geodesic segments for witnesses for which dW(x,y)2Md_{W}(x,y)\leq 2M. The paths σ(x,z)\sigma(x,z) and σ(y,z)\sigma(y,z) 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 Ω2M(x,y,z)\Omega_{2M}(x,y,z)) have length at most 4M4M.

Let wσ(x,y)w\in\sigma(x,y) be any point. We must show that there is some wσ(x,z)σ(y,z)w^{\prime}\in\sigma(x,z)\cup\sigma(y,z) so that ds(w,w)d^{s}(w,w^{\prime}) is uniformly bounded. There are two cases (which actually divide up further into several sub-cases), depending on whether or not ww lies on a witness geodesics for a witness WΩ2M(x,y,z)W\in\Omega_{2M}(x,y,z).

Suppose first that ww lies on a witness geodesic [x,y]Wσ(x,y)[x^{\prime},y^{\prime}]_{W}\subset\sigma(x,y) for WDxW\in D_{x}. By definition of DxD_{x}, WΩM(x,y)ΩM(x,z)W\in\Omega_{M}(x,y)\cap\Omega_{M}(x,z), and so there is also a witness geodesic [x′′,z′′]Wσ(x,z)[x^{\prime\prime},z^{\prime\prime}]_{W}\subset\sigma(x,z). Since there are S˙\dot{S}–geodesics [x,x]S˙,[x,x′′]S˙,[y,y]S˙,[z′′,z]S˙[x,x^{\prime}]_{\dot{S}},[x,x^{\prime\prime}]_{\dot{S}},[y^{\prime},y]_{\dot{S}},[z^{\prime\prime},z]_{\dot{S}} so that every vertex has a nonempty projection to 𝒞(W){\mathcal{C}}(W), and since dW(y,z)<Md_{W}(y,z)<M (again, by definition of DxD_{x}), Theorem 2.6 and the triangle inequality imply

(2) dW(x,x′′)dW(x,x)+dW(x,x′′)2M and dW(y,z′′)dW(y,y)+dW(y,z)+dW(z,z′′)3M.\begin{array}[]{l}d_{W}(x^{\prime},x^{\prime\prime})\leq d_{W}(x^{\prime},x)+d_{W}(x,x^{\prime\prime})\leq 2M\mbox{ and }\\ d_{W}(y^{\prime},z^{\prime\prime})\leq d_{W}(y^{\prime},y)+d_{W}(y,z)+d_{W}(z,z^{\prime\prime})\leq 3M.\end{array}

So [x,y]W[x^{\prime},y^{\prime}]_{W} and [x′′,z′′]W[x^{\prime\prime},z^{\prime\prime}]_{W} are 𝒞(W){\mathcal{C}}(W)–geodesics whose starting and ending points are within distance 3M3M of each other. Since 𝒞(W){\mathcal{C}}(W) is δ\delta–hyperbolic for some δ>0\delta>0, it follows that there is some w[x′′,z′′]Wσ(x,z)w^{\prime}\in[x^{\prime\prime},z^{\prime\prime}]_{W}\subset\sigma(x,z) so that dW(w,w′′)2δ+3Md_{W}(w,w^{\prime\prime})\leq 2\delta+3M. Since 𝒞(W){\mathcal{C}}(W) is a subgraph of 𝒞s(S˙){\mathcal{C}}^{s}(\dot{S}), ds(w,w)2δ+3Md^{s}(w,w^{\prime})\leq 2\delta+3M. We can similarly find the required ww^{\prime} if ww is in a witness geodesic segment for a witness WDyW\in D_{y}.

Next suppose ww lies in the witness geodesic [x,y]W0σ(x,y)[x^{\prime},y^{\prime}]_{W_{0}}\subset\sigma(x,y), for W0D0W_{0}\in D_{0} (if D0D_{0}\neq\emptyset). The argument in this sub-case is similar to the previous one, as we now describe. Let [x′′,z′′]W0σ(x,z)[x^{\prime\prime},z^{\prime\prime}]_{W_{0}}\subset\sigma(x,z) and [y′′,z]W0σ(y,z)[y^{\prime\prime},z^{\prime}]_{W_{0}}\subset\sigma(y,z) be the W0W_{0}–geodesic segments. Arguing as in the proof of (2), we see that the endpoints of these three geodesic segments in 𝒞(W){\mathcal{C}}(W) satisfy

dW0(x,x′′),dW0(y,y′′),dW0(z,z′′)2M.d_{W_{0}}(x^{\prime},x^{\prime\prime}),d_{W_{0}}(y^{\prime},y^{\prime\prime}),d_{W_{0}}(z^{\prime},z^{\prime\prime})\leq 2M.

Since 𝒞(W){\mathcal{C}}(W) is δ\delta–hyperbolic, we can again easily deduce that for some

w[x′′,z′′]W0[y′′,z]W0σ(x,z)σ(y,z),w^{\prime}\in[x^{\prime\prime},z^{\prime\prime}]_{W_{0}}\cup[y^{\prime\prime},z^{\prime}]_{W_{0}}\subset\sigma(x,z)\cup\sigma(y,z),

we have ds(w,w)dW0(w,w)3δ+2Md^{s}(w,w^{\prime})\leq d_{W_{0}}(w,w^{\prime})\leq 3\delta+2M.

Finally, we assume wσ(x,y)w\in\sigma(x,y) lies in a complementary subsegment σ(x,y)σ(x,y)\sigma(x^{\prime},y^{\prime})\subset\sigma(x,y) of one of the Ω2M(x,y,z)\Omega_{2M}(x,y,z)–witness subsegments of σ(x,y)\sigma(x,y) as described above. Note that x,y[x,y]S˙σ(x,y)x^{\prime},y^{\prime}\in[x,y]_{\dot{S}}\cap\sigma(x,y) 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 D0=D_{0}=\emptyset). Thus, depending on which complementary subsegment we are looking at, we claim that one of the following must hold:

  1. (1)

    there exists σ(x′′,z′′)σ(x,z)\sigma(x^{\prime\prime},z^{\prime\prime})\subset\sigma(x,z) so that ds(x,x′′),ds(y,z′′)3Md^{s}(x^{\prime},x^{\prime\prime}),d^{s}(y^{\prime},z^{\prime\prime})\leq 3M,

  2. (2)

    there exists σ(y′′,z′′)σ(y,z)\sigma(y^{\prime\prime},z^{\prime\prime})\subset\sigma(y,z) so that ds(y,y′′),ds(x,z′′)3Md^{s}(y^{\prime},y^{\prime\prime}),d^{s}(x^{\prime},z^{\prime\prime})\leq 3M, or

  3. (3)

    there exists σ(x′′,z′′)σ(x,z)\sigma(x^{\prime\prime},z^{\prime\prime})\subset\sigma(x,z) and σ(y′′,z)σ(y,z)\sigma(y^{\prime\prime},z^{\prime})\subset\sigma(y,z) so that
    ds(x,x′′),ds(y,y′′),ds(z,z′′)3Md^{s}(x^{\prime},x^{\prime\prime}),d^{s}(y^{\prime},y^{\prime\prime}),d^{s}(z^{\prime},z^{\prime\prime})\leq 3M.

The proofs of these statements are very similar to the proof in the case that wDxw\in D_{x} or DyD_{y}. If σ(x,y)\sigma(x^{\prime},y^{\prime}) is a complementary segment which is part of a bigon and xx^{\prime} is in 𝒞(W){\mathcal{C}}(W) for some WDxW\in D_{x} (or x=x′′x=x^{\prime\prime}), then we are in case (1) and we take the corresponding complementary segment σ(x′′,z′′)σ(x,z)\sigma(x^{\prime\prime},z^{\prime\prime})\subset\sigma(x,z) of the bigon with x′′𝒞(W)x^{\prime\prime}\in{\mathcal{C}}(W) (or x′′=xx^{\prime\prime}=x). It follows that all vertices of [x,y]S˙[x^{\prime},y]_{\dot{S}}, [y,z]S˙[y,z]_{\dot{S}}, and [z,x′′]S˙[z,x^{\prime\prime}]_{\dot{S}} have non-empty projections to WW, so by Theorem 2.6 and the triangle inequality we have

ds(x,x′′)dW(x,x′′)dW(x,y)+dW(y,z)+dW(z,x′′)3M.d^{s}(x^{\prime},x^{\prime\prime})\leq d_{W}(x^{\prime},x^{\prime\prime})\leq d_{W}(x^{\prime},y)+d_{W}(y,z)+d_{W}(z,x^{\prime\prime})\leq 3M.

On the other hand, y,z′′𝒞(W)y^{\prime},z^{\prime\prime}\in{\mathcal{C}}(W^{\prime}) for some WDxD0W^{\prime}\in D_{x}\cup D_{0} and similarly

ds(y,z′′)dW(y,z′′)dW(y,x)+dW(x,z′′)2M<3M,d^{s}(y^{\prime},z^{\prime\prime})\leq d_{W^{\prime}}(y^{\prime},z^{\prime\prime})\leq d_{W^{\prime}}(y^{\prime},x)+d_{W^{\prime}}(x,z^{\prime\prime})\leq 2M<3M,

and so the conclusion of (1) holds. If y𝒞(W)y^{\prime}\in{\mathcal{C}}(W) for some WDyW\in D_{y}, then a symmetric argument proves (2) holds. The only other possibility is that D0=D_{0}=\emptyset, x𝒞(W)x^{\prime}\in{\mathcal{C}}(W), and y𝒞(W)y^{\prime}\in{\mathcal{C}}(W^{\prime}), where WDxW\in D_{x} and WDyW^{\prime}\in D_{y}, so that σ(x,y)\sigma(x^{\prime},y^{\prime}) 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 R>0R>0 to the projections to all proper witnesses of the vertices of this quadrilateral (respectively, hexagon). Let ϵ>0\epsilon>0 be the constant from Lemma 4.4 for this RR. By Corollary 4.5, there is some ww^{\prime} on one of the other sides of this quadrilateral/hexagon so that ds(w,w)3ϵd^{s}(w,w^{\prime})\leq 3\epsilon. It may be that ww^{\prime} is in σ(x,z)\sigma(x,z) or σ(y,z)\sigma(y,z), or that it lies in one of the witness segments. As described above, these segments have length at most 3M3M, and so in this latter case, we can find w′′σ(x,z)σ(y,z)w^{\prime\prime}\in\sigma(x,z)\cup\sigma(y,z) with ds(w,w′′)3ϵ+3Md^{s}(w,w^{\prime\prime})\leq 3\epsilon+3M.

Combining all the above, we see that there is always some wσ(x,z)σ(y,z)w^{\prime}\in\sigma(x,z)\cup\sigma(y,z) with ds(w,w)d^{s}(w,w^{\prime}) bounded above by

max{3ϵ+3M,2δ+3M,4δ+2M}.\max\{3\epsilon+3M,2\delta+3M,4\delta+2M\}.

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 kmax{M,24}k\geq\max\{M,24\}, there exists K1K\geq 1, C0C\geq 0 so that

ds(x,y)K,CWΩ(S˙){{dW(x,y)}}k,d^{s}(x,y)\stackrel{{\scriptstyle K,C}}{{\asymp}}\sum_{W\in\Omega(\dot{S})}{\{\{}d_{W}(x,y){\}\}}_{k},

for all x,y𝒞s(S˙)x,y\in{\mathcal{C}^{s}(\dot{S})}.

Recall that here xK,Cyx\stackrel{{\scriptstyle K,C}}{{\asymp}}y is shorthand for the condition 1K(xC)yKx+C\frac{1}{K}(x-C)\leq y\leq Kx+C and that {{x}}k=x{\{\{}x{\}\}}_{k}=x if xkx\geq k and 0, otherwise. Note that we have already proved an upper bound on ds(x,y)d^{s}(x,y) of the required form in Corollary 3.5 and thus we need only prove the lower bound.

Remark 5.2.

As with Theorem 4.1, another approach to this theorem would be to follow Masur-Schleimer [MS13] or Vokes [Vok]. As with Theorem 4.1 we give a proof using survival paths, which is straightforward and elementary.

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 WW and WW^{\prime} are witnesses for 𝒞s(S˙){\mathcal{C}^{s}(\dot{S})} and u𝒞(S˙)u\in{\mathcal{C}}(\dot{S}) with nonempty projection to both WW and WW^{\prime}. Then

dW(u,W)10dW(u,W)4d_{W}(u,\partial W^{\prime})\geq 10\Rightarrow d_{W^{\prime}}(u,\partial W)\leq 4

We will also need the following application which we use to provide an ordering on the witnesses for a pair x,y𝒞s(S˙)x,y\in{\mathcal{C}^{s}(\dot{S})} 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 k14k\geq 14 and W,WW,W^{\prime} are witnesses in the set Ωk(x,y)\Omega_{k}(x,y). Then the following are equivalent:

(1) dW(y,W)10d_{W^{\prime}}(y,\partial W)\geq 10 (2) dW(y,W)4d_{W}(y,\partial W^{\prime})\leq 4
(3) dW(x,W)10d_{W}(x,\partial W^{\prime})\geq 10 (4) dW(x,W)4d_{W^{\prime}}(x,\partial W)\leq 4
Proof.

By Lemma 5.3 we have (1)(2)(1)\Rightarrow(2) and (3)(4)(3)\Rightarrow(4). To prove (2)(3)(2)\Rightarrow(3) we use triangle inequality:

dW(x,W)dW(x,y)dW(y,W)k410d_{W}(x,\partial W^{\prime})\geq d_{W}(x,y)-d_{W}(y,\partial W^{\prime})\geq k-4\geq 10

since k14k\geq 14. The proof of (4)(1)(4)\Rightarrow(1) is identical to the proof that (2)(3)(2)\Rightarrow(3). ∎

Definition 5.5.

For any k14k\geq 14, we define a relation << on Ωk(x,y)\Omega_{k}(x,y), declaring W<WW<W^{\prime} for W,WΩk(x,y)W,W^{\prime}\in\Omega_{k}(x,y), if any of the equivalent statements of the Proposition 5.4 is satisfied.

Lemma 5.6.

For any k20k\geq 20, the relation << is a total order on Ωk(x,y)\Omega_{k}(x,y).

Proof.

We first prove that any two element W,WΩk(x,y)W,W^{\prime}\in\Omega_{k}(x,y) are ordered. If not, then that means Proposition 5.4 (3) fails to hold as stated, or with yy replacing xx, and thus we have dW(y,W)<10d_{W}(y,\partial W^{\prime})<10 and dW(x,W)<10d_{W}(x,\partial W^{\prime})<10. Hence,

dW(x,y)dW(x,W)+dW(y,W)<20kd_{W}(x,y)\leq d_{W}(x,\partial W^{\prime})+d_{W}(y,\partial W^{\prime})<20\leq k

which contradicts the assumption that WΩk(x,y)W\in\Omega_{k}(x,y).

The relation is clearly anti-symmetric, so it remains to prove that it is transitive. To that end, let W<W<W′′W<W^{\prime}<W^{\prime\prime} in Ωk(x,y)\Omega_{k}(x,y), and we assume WW′′W\not<W^{\prime\prime}, hence W′′<WW^{\prime\prime}<W. Since W<WW<W^{\prime} and W′′<WW^{\prime\prime}<W, we have dW(y,W)4d_{W}(y,\partial W^{\prime})\leq 4 and dW(x,W′′)4d_{W}(x,\partial W^{\prime\prime})\leq 4. So by the triangle inequality

dW(W,W′′)dW(x,y)dW(y,W)dW(x,W′′)k8>10.d_{W}(\partial W^{\prime},\partial W^{\prime\prime})\geq d_{W}(x,y)-d_{W}(y,\partial W^{\prime})-d_{W}(x,\partial W^{\prime\prime})\geq k-8>10.

Then by Lemma 5.3, we have

dW(W,W′′)4.d_{W^{\prime}}(\partial W,\partial W^{\prime\prime})\leq 4.

So, appealing to the fact that W<WW<W^{\prime} and W<W′′W^{\prime}<W^{\prime\prime} and Proposition 5.4 the triangle inequality implies

20kdW(x,y)dW(x,W)+dW(W,W′′)+dW(W′′,y)12,20\leq k\leq d_{W^{\prime}}(x,y)\leq d_{W^{\prime}}(x,\partial W)+d_{W^{\prime}}(\partial W,\partial W^{\prime\prime})+d_{W^{\prime}}(\partial W^{\prime\prime},y)\leq 12,

a contradiction. ∎

The next lemma is also useful in the proof of Theorem 5.1.

Lemma 5.7.

Let x,y,u𝒞s(S˙)x,y,u\in{\mathcal{C}^{s}(\dot{S})}, W,WΩk(x,y)W,W^{\prime}\in\Omega_{k}(x,y) with k20k\geq 20, and W<WW<W^{\prime}. Then,

dW(u,y)14dW(u,x)8d_{W}(u,y)\geq 14\Rightarrow d_{W^{\prime}}(u,x)\leq 8
Proof.

From our assumptions, the definition of the order on Ωk(x,y)\Omega_{k}(x,y), and the triangle inequality we have

dW(u,W)dW(u,y)dW(y,W)144=10.d_{W}(u,\partial W^{\prime})\geq d_{W}(u,y)-d_{W}(y,\partial W^{\prime})\geq 14-4=10.

By Lemma 5.3, we have dW(u,W)4d_{W^{\prime}}(u,\partial W)\leq 4. Thus, by the definition of the order on Ωk(x,y)\Omega_{k}(x,y) and the triangle inequality, we have

dW(u,x)dW(u,W)+dW(W,x)4+4=8.d_{W^{\prime}}(u,x)\leq d_{W^{\prime}}(u,\partial W)+d_{W^{\prime}}(\partial W,x)\leq 4+4=8.

We are now ready to prove the lower bound in Theorem 5.1, which we record in the following proposition.

Proposition 5.8.

Fix k24k\geq 24. Given x,y𝒞s(S˙)x,y\in{\mathcal{C}^{s}(\dot{S})} we have

ds(x,y)196WΩ(S˙){{dW(x,y)}}kd^{s}(x,y)\geq\frac{1}{96}\sum_{W\in\Omega(\dot{S})}{\{\{}d_{W}(x,y){\}\}}_{k}
Proof.

Let [x,y][x,y] be a geodesic between x,yx,y in 𝒞s(S˙){\mathcal{C}^{s}(\dot{S})}, and denote its vertices

x=x0,x1,,xn1,xn=y.x=x_{0},x_{1},\ldots,x_{n-1},x_{n}=y.

So, n=ds(x,y)n=d^{s}(x,y) is the length of [x,y][x,y]. Let m=|Ωk(x,y)|m=|\Omega_{k}(x,y)|, suppose m>0m>0, and write

Ωk(x,y)={W1<W2<<Wm}.\Omega_{k}(x,y)=\{W_{1}<W_{2}<\cdots<W_{m}\}.

For each 1j<m1\leq j<m let 0ijn0\leq i_{j}\leq n be such that dWj(xij,y)14d_{W_{j}}(x_{i_{j}},y)\geq 14 and dWj(x,y)13d_{W_{j}}(x_{\ell},y)\leq 13 for all >ij\ell>i_{j}. That is, xijx_{i_{j}} is the last vertex z[x,y]z\in[x,y] for which dWj(z,y)14d_{W_{j}}(z,y)\geq 14. Then, if j>jj^{\prime}>j, so Wj<WjW_{j}<W_{j^{\prime}}, Lemma 5.7 implies dWj(xij,x)8d_{W_{j^{\prime}}}(x_{i_{j}},x)\leq 8 and so

dWj(xij,y)dWj(x,y)dWj(xij,x)k8248=16.d_{W_{j^{\prime}}}(x_{i_{j}},y)\geq d_{W_{j^{\prime}}}(x,y)-d_{W_{j^{\prime}}}(x_{i_{j}},x)\geq k-8\geq 24-8=16.

Since the projection πWj\pi_{W_{j^{\prime}}} is 22–Lipschitz (see Proposition 2.5) and xijx_{i_{j}} and xij+1x_{i_{j}+1} are distance 11 in 𝒞s(S˙){\mathcal{C}^{s}(\dot{S})}, we have

dWj(xij+1,y)dWj(xij,y)dWj(xij,xij+1)16214.d_{W_{j^{\prime}}}(x_{i_{j}+1},y)\geq d_{W_{j^{\prime}}}(x_{i_{j}},y)-d_{W_{j^{\prime}}}(x_{i_{j}},x_{i_{j}+1})\geq 16-2\geq 14.

Therefore, i1<i2<<im1i_{1}<i_{2}<\cdots<i_{m-1}. Set i0=0i_{0}=0 and im=ni_{m}=n.

Given 1j<m1\leq j<m, dWj(xij+1,y)13d_{W_{j}}(x_{i_{j}+1},y)\leq 13 and again appealing to Proposition 2.5, we have

dWj(xij,y)dWj(xij,xij+1)+dWj(xij+1,y)2+1315.d_{W_{j}}(x_{i_{j}},y)\leq d_{W_{j}}(x_{i_{j}},x_{i_{j}+1})+d_{W_{j}}(x_{i_{j}+1},y)\leq 2+13\leq 15.

Observe this inequality is trivially true for j=mj=m since y=xn=ximy=x_{n}=x_{i_{m}} and so the left hand side is at most 22 in this case. Another application of Lemma 5.7 implies dWj(x,xij1)8d_{W_{j}}(x,x_{i_{j-1}})\leq 8 for all 1jm1\leq j\leq m (the case j=1j=1 is similarly trivial). Therefore

(3) dWj(xij1,xij)dWj(x,y)dWj(x,xij1)dWj(xij,y)dWj(x,y)23,d_{W_{j}}(x_{i_{j-1}},x_{i_{j}})\geq d_{W_{j}}(x,y)-d_{W_{j}}(x,x_{i_{j-1}})-d_{W_{j}}(x_{i_{j}},y)\geq d_{W_{j}}(x,y)-23,

for all 1jm1\leq j\leq m.

Appealing one more time to Proposition 2.5, together with Inequality (3), we have

ds(x,y)=n=j=1mijij112j=1mdWj(xij1,xij)12j=1m(dWj(x,y)23)d^{s}(x,y)=n=\sum_{j=1}^{m}i_{j}-i_{j-1}\geq\frac{1}{2}\sum_{j=1}^{m}d_{W_{j}}(x_{i_{j-1}},x_{i_{j}})\geq\frac{1}{2}\sum_{j=1}^{m}(d_{W_{j}}(x,y)-23)

Next, observe that since dWj(x,y)k24d_{W_{j}}(x,y)\geq k\geq 24 we have

dWj(x,y)23124dWj(x,y).d_{W_{j}}(x,y)-23\geq\frac{1}{24}d_{W_{j}}(x,y).

Since 𝒞s(S˙)𝒞(S˙){\mathcal{C}}^{s}(\dot{S})\subset{\mathcal{C}}(\dot{S}) is a subcomplex, we have ds(x,y)dS˙(x,y)d^{s}(x,y)\geq d_{\dot{S}}(x,y) and so

2ds(x,y)dS˙(x,y)+148WΩk(x,y)dW(x,y)148WΩ(S˙){{dW(x,y)}}k.2d^{s}(x,y)\geq d_{\dot{S}}(x,y)+\frac{1}{48}\sum_{W\in\Omega_{k}(x,y)}d_{W}(x,y)\geq\frac{1}{48}\sum_{W\in\Omega(\dot{S})}{\{\{}d_{W}(x,y){\}\}}_{k}.

Proof of Theorem 5.1.

Given kmax{M,24}k\geq\max\{M,24\}, let K=max{2k,96}K=\max\{2k,96\} and C=2k2C=2k^{2}. The theorem then follows from Corollary 3.5 and Proposition 5.8. ∎

As a consequence of the Theorem 5.1 we have the following two facts.

Corollary 5.9.

Given a witness WSW\subset S, the inclusion map 𝒞(W)𝒞s(S˙){\mathcal{C}}(W)\hookrightarrow{\mathcal{C}^{s}(\dot{S})} is a quasi-isometric embedding.

Corollary 5.10.

Survival paths are uniform quasi-geodesics in 𝒞s(S˙){\mathcal{C}^{s}(\dot{S})}.

Moreover, we have

Lemma 5.11.

Survival paths can be reparametrized to be uniform quasi-geodesics in 𝒞(S˙)\mathcal{C}(\dot{S}).

Proof.

Let σ(x,y)\sigma(x,y) be a survival path with main geodesic [x,y]S˙[x,y]_{\dot{S}}. For every proper witness WS˙W\subsetneq\dot{S}, if there is a WW–witness geodesic segment in σ(x,y)\sigma(x,y), we reparametrize along this segment so that it is traversed along an interval of length 22. Since such WW–witness geodesic segments replaced geodesic subsegments of [x,y]S˙[x,y]_{\dot{S}} of length 22, and since they lie in the 11–neighborhood of W\partial W, this clearly defines the required reparametrization. ∎

Corollary 5.12.

Any infinite survival path is a uniform quasi-geodesic.

Proof.

This is immediate from Corollary 5.10 and Proposition 3.9. ∎

6. Boundary of the surviving curve complex 𝒞s(S˙){\mathcal{C}^{s}(\dot{S})}

Recall that we denote the disjoint union of ending lamination spaces of all witnesses by

s(S˙):=WΩ(S˙)(W).\mathcal{E}\mathcal{L}^{s}(\dot{S}):=\bigsqcup_{W\in\Omega(\dot{S})}\mathcal{E}\mathcal{L}(W).

We call this the space of surviving ending laminations of S˙\dot{S}, 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 :𝒞s(S˙)s(S˙)\mathcal{F}\colon\partial{\mathcal{C}}^{s}(\dot{S})\to\mathcal{EL}^{s}(\dot{S}) such that for any sequence {αn}𝒞s(S˙)\{\alpha_{n}\}\subset{\mathcal{C}}^{s}(\dot{S}), αnx\alpha_{n}\to x in 𝒞¯s(S˙)\bar{\mathcal{C}}^{s}(\dot{S}) if and only if αnCH(x)\alpha_{n}\xrightarrow{\text{CH}}\mathcal{F}(x).

We denote the Gromov product of α,β𝒞s(S˙)\alpha,\beta\in{\mathcal{C}^{s}(\dot{S})} based at o𝒞s(S˙)o\in{\mathcal{C}^{s}(\dot{S})} by α,βos\langle\alpha,\beta\rangle_{o}^{s}, and recall that the Gromov boundary 𝒞s(S˙)\partial{\mathcal{C}^{s}(\dot{S})} of 𝒞s(S˙){\mathcal{C}^{s}(\dot{S})} is defined to be the set of equivalence classes of sequences {αn}\{\alpha_{n}\} which converge at infinity with respect to ,os\langle\,\,,\,\,\rangle_{o}^{s}. 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 WS˙W\subsetneq\dot{S}, Corollary 5.9 implies that 𝒞(W)\partial{\mathcal{C}}(W) embeds into 𝒞s(S˙)\partial{\mathcal{C}^{s}(\dot{S})}. Likewise, Corollary 5.12, combined together with the fact that dS˙dsd_{\dot{S}}\leq d^{s} implies that 𝒞(S˙)\partial{\mathcal{C}}(\dot{S}) also embeds 𝒞s(S˙)\partial{\mathcal{C}^{s}(\dot{S})}. Using these embeddings, we view 𝒞(W)\partial{\mathcal{C}}(W) as a subspace of 𝒞s(S˙)\partial{\mathcal{C}^{s}(\dot{S})}, for all witnesses WS˙W\subseteq\dot{S}. The next proposition says that the subspaces are all disjoint.

Proposition 6.2.

For any two witnesses WWW\neq W^{\prime} for S˙\dot{S}, 𝒞(W)𝒞(W)=\partial{\mathcal{C}}(W)\cap\partial{\mathcal{C}}(W^{\prime})=\emptyset.

Proof.

Let x𝒞(W)x\in\partial{\mathcal{C}}(W) and x𝒞(W)x^{\prime}\in\partial{\mathcal{C}}(W^{\prime}). Then by Lemma 3.8, there is a bi-infinite survival path σ(x,x)\sigma(x,x^{\prime}) and by Corollary 5.12 this survival path is a quasi geodesic. Hence xxx\neq x^{\prime}. ∎

We now have a natural inclusion of the disjoint union of Gromov boundaries

WΩ(S˙)𝒞(W)𝒞s(S˙).\bigsqcup_{W\in\Omega(\dot{S})}\partial{\mathcal{C}}(W)\subset\partial{\mathcal{C}^{s}(\dot{S})}.

In fact, this disjoint union accounts for the entire Gromov boundary.

Lemma 6.3.

We have

WΩ(S˙)𝒞(W)=𝒞s(S˙).\bigsqcup_{W\in\Omega(\dot{S})}\partial{\mathcal{C}}(W)=\partial{\mathcal{C}^{s}(\dot{S})}.
Proof.

Let x𝒞s(S˙)x\in\partial{\mathcal{C}^{s}(\dot{S})} and αnx𝒞s(S˙)\alpha_{n}\rightarrow x\in\partial{\mathcal{C}^{s}(\dot{S})}, and we assume without loss of generality that {αn}\{\alpha_{n}\} is a quasi-geodesic in 𝒞s(S˙){\mathcal{C}}^{s}(\dot{S}) and that the first vertex is the basepoint α0=o\alpha_{0}=o. If dS˙(αn,o)d_{\dot{S}}(\alpha_{n},o)\rightarrow\infty as nn\rightarrow\infty, then given R>0R>0, let N>0N>0 be such that dS˙(αn,o)Rd_{\dot{S}}(\alpha_{n},o)\geq R for all nNn\geq N. For any mnNm\geq n\geq N, the subsegment of the quasi-geodesics, {αn,αn+1,,αm}\{\alpha_{n},\alpha_{n+1},\ldots,\alpha_{m}\}, is some uniformly bounded distance DD to σ(αn,αm)\sigma(\alpha_{n},\alpha_{m}) in 𝒞s(S˙){\mathcal{C}}^{s}(\dot{S}), by hyperbolicity and Corollary 5.10. Therefore, the distance in 𝒞(S˙){\mathcal{C}}(\dot{S}) from any point of σ(αn,αm)\sigma(\alpha_{n},\alpha_{m}) to oo is at least RDR-D. So the distance from any point of [αn,αm]S˙[\alpha_{n},\alpha_{m}]_{\dot{S}} to oo is at least RD1R-D-1. Letting RR\to\infty, it follows that αn,αmo\langle\alpha_{n},\alpha_{m}\rangle_{o}\rightarrow\infty in 𝒞(S˙)\mathcal{C}(\dot{S}). Consequently, {αn}\{\alpha_{n}\} converges to a point in 𝒞(S˙)\partial{\mathcal{C}}(\dot{S}), so x𝒞(S˙)x\in\partial{\mathcal{C}}(\dot{S}). For the rest of the proof, we may assume that dS˙(αn,o)d_{\dot{S}}(\alpha_{n},o) is bounded by some constant 0<R<0<R<\infty for all nn.

By the distance formula 5.1,

ds(α0,αn)K,CWΩ(S˙){{dW(α0,αn)}}k,d^{s}(\alpha_{0},\alpha_{n})\stackrel{{\scriptstyle K,C}}{{\asymp}}\sum_{W\in\Omega(\dot{S})}{\{\{}d_{W}(\alpha_{0},\alpha_{n}){\}\}}_{k},

and since ds(α0,αn)d^{s}(\alpha_{0},\alpha_{n})\rightarrow\infty, we can find a witness WnW_{n} for σ(α0,αn)\sigma(\alpha_{0},\alpha_{n}) for each nn\in\mathbb{N}, so that we have dWn(α0,αn)d_{W_{n}}(\alpha_{0},\alpha_{n})\rightarrow\infty as nn\to\infty.

We would like to show that there is a unique witness WW such that dW(α0,αn)d_{W}(\alpha_{0},\alpha_{n})\rightarrow\infty. To do that, let h>0h>0 be the maximal Hausdorff distance in 𝒞s(S˙){\mathcal{C}}^{s}(\dot{S}) between σ(α0,αn)\sigma(\alpha_{0},\alpha_{n}) and {αk}k=0n\{\alpha_{k}\}_{k=0}^{n}, for all n0n\geq 0 (which is finite by hyperbolicity of 𝒞s(S˙){\mathcal{C}}^{s}(\dot{S}) and Corollary 5.10).

Claim 6.4.

Given nn\in\mathbb{N}, if dW(α0,αn)M+1+2(h+1)d_{W}(\alpha_{0},\alpha_{n})\geq M+1+2(h+1) for some witness WS˙W\subsetneq\dot{S}, then WW is a witness for σ(α0,αm)\sigma(\alpha_{0},\alpha_{m}), for all mnm\geq n.

Proof.

Let αnσ(α0,αm)\alpha_{n}^{\prime}\in\sigma(\alpha_{0},\alpha_{m}) be such that ds(αn,αn)hd^{s}(\alpha_{n},\alpha_{n}^{\prime})\leq h. If WW is not a witness for σ(α0,αm)\sigma(\alpha_{0},\alpha_{m}), then every vertex of the main geodesic [α0,αm]S˙[\alpha_{0},\alpha_{m}]_{\dot{S}} of σ(α0,αm)\sigma(\alpha_{0},\alpha_{m}) has nonempty project to WW. Furthermore, the geodesic in 𝒞s(S˙){\mathcal{C}}^{s}(\dot{S}) from αn\alpha_{n} to αn\alpha_{n}^{\prime} of length at most hh can be extended to a path in 𝒞(S˙{\mathcal{C}}(\dot{S}) to [α0,αm]S˙[\alpha_{0},\alpha_{m}]_{\dot{S}} of length at most h+1h+1 such that every vertex has nonempty project to WW. By Proposition 2.5 and the triangle inequality we have,

|dW(α0,αn)dW(α0,αn)|2(h+1).|d_{W}(\alpha_{0},\alpha_{n})-d_{W}(\alpha_{0},\alpha_{n}^{\prime})|\leq 2(h+1).

If αn[α0,αm]S˙\alpha_{n}^{\prime}\in[\alpha_{0},\alpha_{m}]_{\dot{S}}, then since every vertex of this geodesic has nonempty projection to WW, it follows that dW(α0,αn)<Md_{W}(\alpha_{0},\alpha_{n}^{\prime})<M. If αn[α0,αm]S˙\alpha_{n}^{\prime}\not\in[\alpha_{0},\alpha_{m}]_{\dot{S}}, then there is a witness WW^{\prime} for [α0,αm]S˙[\alpha_{0},\alpha_{m}]_{\dot{S}} such that dS˙(αn,W)=1d_{\dot{S}}(\alpha^{\prime}_{n},\partial W^{\prime})=1, and as a result dW(α0,αn)<M+1d_{W}(\alpha_{0},\alpha_{n}^{\prime})<M+1. In any case,

dW(α0,αn)dW(α0,αn)+2(h+1)<M+1+2(h+1),d_{W}(\alpha_{0},\alpha_{n})\leq d_{W}(\alpha_{0},\alpha_{n}^{\prime})+2(h+1)<M+1+2(h+1),

which is a contradiction. This proves the claim. ∎

Since dWn(α0,αn)d_{W_{n}}(\alpha_{0},\alpha_{n})\to\infty, there exists n0>0n_{0}>0 such that dWn0(α0,αn0)2(h+1)+M+1d_{W_{n_{0}}}(\alpha_{0},\alpha_{n_{0}})\geq 2(h+1)+M+1, and hence for all mn0m\geq n_{0}, Wn0W_{n_{0}} is a witness for σ(α0,αm)\sigma(\alpha_{0},\alpha_{m}). Let Ω([α0,αn]S˙)\Omega([\alpha_{0},\alpha_{n}]_{\dot{S}}) be the set of proper witnesses for [α0,αn]S˙[\alpha_{0},\alpha_{n}]_{\dot{S}}, and set

Ωn=m=nΩ([α0,αm]S˙).\Omega_{n}=\bigcap_{m=n}^{\infty}\Omega([\alpha_{0},\alpha_{m}]_{\dot{S}}).

Note that ΩnΩn+1\Omega_{n}\subset\Omega_{n+1} for all nn and that Ωn\Omega_{n} is nonempty for all nn0n\geq n_{0}. Since each Ωn\Omega_{n} contains no more than R/2R/2 elements by Corollary 5.10, the (nested) union Ω\Omega_{\infty} is given by Ω=ΩN\Omega_{\infty}=\Omega_{N} for some Nn0N\geq n_{0}. The boundaries of the witnesses in Ω\Omega_{\infty} lie on the geodesic [α0,αm]S˙[\alpha_{0},\alpha_{m}]_{\dot{S}} for all mNm\geq N, and we let WΩW_{\infty}\in\Omega_{\infty} be the one furthest from α0\alpha_{0}. Without loss of generality, we may assume that [α0,αm]S˙[\alpha_{0},\alpha_{m}]_{\dot{S}} and [α0,αm]S˙[\alpha_{0},\alpha_{m^{\prime}}]_{\dot{S}} all agree on [α0,W][\alpha_{0},\partial W_{\infty}], for all m,mNm,m^{\prime}\geq N.

For any mNm\geq N and any witness WW of [α0,αN]S˙[\alpha_{0},\alpha_{N}]_{\dot{S}} with W\partial W further from W\partial W_{\infty}, note that dW(α0,αm)<M+1+2(h+1)d_{W}(\alpha_{0},\alpha_{m})<M+1+2(h+1): otherwise, by Claim 6.4 WW would be a witness for [α0,αm]S˙[\alpha_{0},\alpha_{m^{\prime}}]_{\dot{S}} for all mmm^{\prime}\geq m and so WΩW\in\Omega_{\infty} with W\partial W further from α0\alpha_{0} than W\partial W_{\infty}, a contradiction to our choice of WW_{\infty}.

For any nNn\geq N, let βn\beta_{n} be the last vertex of σ(α0,αn)\sigma(\alpha_{0},\alpha_{n}) in 𝒞(W){\mathcal{C}}(W_{\infty}). By the previous paragraph together with Theorem 2.6 and the bound

dS˙(βn,αn)dS˙(α0,αn)R/2,d_{\dot{S}}(\beta_{n},\alpha_{n})\leq d_{\dot{S}}(\alpha_{0},\alpha_{n})\leq R/2,

we see that the subpath of σ(α0,αn)\sigma(\alpha_{0},\alpha_{n}) from βn\beta_{n} to αn\alpha_{n} has length bounded above by some constant C>0C>0, independent of nn. In particular, ds(αn,βn)Cd^{s}(\alpha_{n},\beta_{n})\leq C. Therefore, αn\alpha_{n} and βn\beta_{n} converge to the same point xx on the Gromov boundary of 𝒞s(S˙){\mathcal{C}}^{s}(\dot{S}). Since βn𝒞(W)\beta_{n}\in{\mathcal{C}}(W_{\infty}), which is quasi-isometrically embedded in 𝒞s(S˙){\mathcal{C}}^{s}(\dot{S}), it follows that x𝒞(W)x\in\partial{\mathcal{C}}(W_{\infty}), as required. ∎

The next Lemma provides a convenient tool for deciding when a sequence in 𝒞s(S˙){\mathcal{C}}^{s}(\dot{S}) converges to a point in 𝒞(W)\partial{\mathcal{C}}(W), for some proper witness WW.

Lemma 6.5.

Given {αn}𝒞s(S˙)\{\alpha_{n}\}\subset{\mathcal{C}^{s}(\dot{S})} and x𝒞(W)x\in\partial\mathcal{C}(W) for some witness WW, then αnx\alpha_{n}{\rightarrow}x if and only if πW(αn)x\pi_{W}(\alpha_{n})\rightarrow x.

Proof.

Throughout, we assume o=α0o=\alpha_{0}, the basepoint, which without loss of generality we assume lies in WW, and let {βn}𝒞(W)\{\beta_{n}\}\subset{\mathcal{C}}(W) be any sequence converging to xx, so that for the Gromov product in 𝒞(W){\mathcal{C}}(W) we have βn,βmoW\langle\beta_{n},\beta_{m}\rangle_{o}^{W}\to\infty as n,mn,m\to\infty.

Since σ(αn,βn)\sigma(\alpha_{n},\beta_{n}) is a uniform quasi-geodesic by Corollary 5.10 it follows that

ds(o,σ(αn,βn))αn,βmos,d^{s}(o,\sigma(\alpha_{n},\beta_{n}))\asymp\langle\alpha_{n},\beta_{m}\rangle^{s}_{o},

with uniform constants (where the distance on the left is the minimal distance from oo to the survival path).

Let δn,m\delta_{n,m} be the first point of intersection of σ(αn,βm)\sigma(\alpha_{n},\beta_{m}) with 𝒞(W){\mathcal{C}}(W) (starting from αn\alpha_{n}). By Lemma 3.4, dW(δn,m,αn)<Md_{W}(\delta_{n,m},\alpha_{n})<M. Consequently, because δn,m𝒞(W)\delta_{n,m}\in{\mathcal{C}}(W) and πW(αn)𝒞(W)\pi_{W}(\alpha_{n})\subset{\mathcal{C}}(W), this means

ds(δn,m,πW(αn))dW(δn,m,πW(αn))=dW(δn,m,αn)<M.d^{s}(\delta_{n,m},\pi_{W}(\alpha_{n}))\leq d_{W}(\delta_{n,m},\pi_{W}(\alpha_{n}))=d_{W}(\delta_{n,m},\alpha_{n})<M.

Therefore, by hyperbolicity, the 𝒞s(S˙{\mathcal{C}}^{s}(\dot{S})–geodesic from (any curve in) πW(αn)\pi_{W}(\alpha_{n}) to βn\beta_{n} lies in a uniformly bounded neighborhood of σ(αn,βm)\sigma(\alpha_{n},\beta_{m}), and so

πW(αn),βmosds(o,[πW(αn),βm])ds(o,σ(αn,βm))αn,βmos.\langle\pi_{W}(\alpha_{n}),\beta_{m}\rangle^{s}_{o}\asymp d^{s}(o,[\pi_{W}(\alpha_{n}),\beta_{m}])\succeq d^{s}(o,\sigma(\alpha_{n},\beta_{m}))\asymp\langle\alpha_{n},\beta_{m}\rangle^{s}_{o}.

If αnx\alpha_{n}\to x, then the right-hand side of the above coarse inequality tends to infinty, and hence so does the left-hand side. This implies πW(αn)x\pi_{W}(\alpha_{n})\to x.

Next suppose that πW(αn)x𝒞(W)\pi_{W}(\alpha_{n})\to x\in\partial{\mathcal{C}}(W). As above, we have

αn,πW(αn)osds(o,[αn,πW(αn)]),\langle\alpha_{n},\pi_{W}(\alpha_{n})\rangle_{o}^{s}\asymp d^{s}(o,[\alpha_{n},\pi_{W}(\alpha_{n})]),

and so it suffices to show that the right-hand side tends to infinity as nn\to\infty. Since πW(αn)x𝒞(W)\pi_{W}(\alpha_{n})\to x\in\partial{\mathcal{C}}(W), we have dW(o,αn)d_{W}(o,\alpha_{n})\to\infty, and setting δn\delta_{n} to be the first point of σ(αn,o)\sigma(\alpha_{n},o) in 𝒞(W){\mathcal{C}}(W), Lemma 3.4 implies that ds(δn,πW(αn))<Md^{s}(\delta_{n},\pi_{W}(\alpha_{n}))<M. Therefore, σ(αn,o)\sigma(\alpha_{n},o) passes within dsd^{s}–distance MM of πW(αn)\pi_{W}(\alpha_{n}) on its way to oo. Since σ(αn,o)\sigma(\alpha_{n},o) is a uniform quasi-geodesic by Corollary 5.10, it follows that [αn,πW(αn)][\alpha_{n},\pi_{W}(\alpha_{n})] is uniformly Hausdorff close to the initial segment Jnσ(αn,o)J_{n}\subset\sigma(\alpha_{n},o) from αn\alpha_{n} to δn\delta_{n}. Since the closest point of JnJ_{n} to oo is, coarsely, the point δn\delta_{n}, which is uniformly close to πW(αn)\pi_{W}(\alpha_{n}), we have

αn,πW(αn)os\displaystyle\langle\alpha_{n},\pi_{W}(\alpha_{n})\rangle_{o}^{s} \displaystyle\asymp ds(o,[αn,πW(αn)])\displaystyle d^{s}(o,[\alpha_{n},\pi_{W}(\alpha_{n})])
\displaystyle\asymp ds(o,Jn)ds(o,δn)ds(o,πW(αn))dW(o,αn).\displaystyle d^{s}(o,J_{n})\,\,\asymp\,\,d^{s}(o,\delta_{n})\,\,\asymp\,\,d^{s}(o,\pi_{W}(\alpha_{n}))\,\,\asymp\,\,d_{W}(o,\alpha_{n})\to\infty.

Therefore, αn\alpha_{n} and πW(αn)\pi_{W}(\alpha_{n}) converge together to x𝒞(W)x\in\partial{\mathcal{C}}(W). This completes the proof. ∎

Proof of Theorem 6.1.

By Lemma 6.3, for any x𝒞s(S˙)x\in\partial{\mathcal{C}^{s}(\dot{S})}, there exists a witness WS˙W\subseteq\dot{S} so x𝒞(W)x\in\partial{\mathcal{C}}(W). Let (x)=W(x)\mathcal{F}(x)=\mathcal{F}_{W}(x), where W:𝒞(W)(W)\mathcal{F}_{W}\colon\partial{\mathcal{C}}(W)\to\mathcal{EL}(W) is the homeomorphism given by Theorem 2.12. This defines a bijection :𝒞s(S˙)s(S˙)\mathcal{F}\colon\partial{\mathcal{C}^{s}(\dot{S})}\to\mathcal{EL}^{s}(\dot{S}).

We let x𝒞s(S˙)x\in\partial{\mathcal{C}^{s}(\dot{S})} with αnx\alpha_{n}\to x in 𝒞¯s(S˙)\bar{\mathcal{C}}^{s}(\dot{S}), and prove that αn\alpha_{n} coarse Hausdorff converges to (x)\mathcal{F}(x). Let WS˙W\subseteq\dot{S} be the witness with x𝒞(W)x\in\partial{\mathcal{C}}(W). According to Proposition 6.5, πW(αn)x\pi_{W}(\alpha_{n})\to x in 𝒞¯(W)\bar{\mathcal{C}}(W). By Theorem 2.12, πW(αn)CHW(x)=(x)\pi_{W}(\alpha_{n})\stackrel{{\scriptstyle CH}}{{\to}}\mathcal{F}_{W}(x)=\mathcal{F}(x), and by Lemma 2.13, αnCH(x)\alpha_{n}\stackrel{{\scriptstyle CH}}{{\to}}\mathcal{F}(x), as required.

To prove the other implication, we suppose that αnCH\alpha_{n}\stackrel{{\scriptstyle CH}}{{\to}}\mathcal{L}, for some s(S˙)\mathcal{L}\in\mathcal{EL}^{s}(\dot{S}), and prove that αnx\alpha_{n}\to x in 𝒞¯s(S˙)\bar{\mathcal{C}}^{s}(\dot{S}) where (x)=\mathcal{F}(x)=\mathcal{L}. Let WSW\subseteq S be the witness with (W)\mathcal{L}\in\mathcal{EL}(W). By Lemma 2.13, πW(αn)CH\pi_{W}(\alpha_{n})\stackrel{{\scriptstyle CH}}{{\to}}\mathcal{L}. By Theorem 2.12, πW(αn)x\pi_{W}(\alpha_{n})\to x in 𝒞¯(W)\bar{\mathcal{C}}(W) where W(x)=\mathcal{F}_{W}(x)=\mathcal{L}. By Proposition 6.5, αnx\alpha_{n}\to x in 𝒞¯s(S˙)\bar{\mathcal{C}}^{s}(\dot{S}) and (x)=W(x)=\mathcal{F}(x)=\mathcal{F}_{W}(x)=\mathcal{L}.

All that remains is to show that \mathcal{F} 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 \mathcal{F} is continuous. Let {xn}𝒞s(S˙)\{x^{n}\}\subset\partial{\mathcal{C}^{s}(\dot{S})} with xnxx^{n}\to x as nn\to\infty. Pass to any Hausdorff convergence subsequence so that (xn)H\mathcal{F}(x^{n})\xrightarrow{\text{H}}\mathcal{L} for some lamination \mathcal{L}. If we can show that (x)\mathcal{F}(x)\subseteq\mathcal{L}, then this will show that the original sequence coarse Hausdorff converges to (x)\mathcal{F}(x), and thus \mathcal{F} will be continuous.

For each nn, let {αkn}k=1𝒞s(S˙)\{\alpha_{k}^{n}\}_{k=1}^{\infty}\subset{\mathcal{C}}^{s}(\dot{S}) be a sequence with αknxn\alpha_{k}^{n}\rightarrow x^{n} as kk\rightarrow\infty. Since xnxx^{n}\to x, we may pass to subsequences so that for any sequence {kn}\{k_{n}\}, we have αknnx\alpha_{k_{n}}^{n}\to x as nn\to\infty. From the first part of the argument, αknCH(xn)\alpha_{k}^{n}\xrightarrow{\text{CH}}\mathcal{F}(x^{n}) as kk\to\infty, for all nn. For each nn, pass to a subsequence so that αknHn\alpha_{k}^{n}\xrightarrow{\text{H}}\mathcal{L}_{n}, thus (xn)n\mathcal{F}(x^{n})\subseteq\mathcal{L}_{n}. By passing to yet a further subsequence for each nn, we may assume dH(αkn,n)<1nd_{H}(\alpha_{k}^{n},\mathcal{L}_{n})<\frac{1}{n} for all kk; in particular, this holds for k=1k=1. Now pass to a subsequence of {n}\{\mathcal{L}_{n}\} so that nHo\mathcal{L}_{n}\stackrel{{\scriptstyle H}}{{\to}}\mathcal{L}_{o} for some lamination o\mathcal{L}_{o}, it follows that α1nHo\alpha_{1}^{n}\stackrel{{\scriptstyle H}}{{\to}}\mathcal{L}_{o}, as nn\to\infty. Since α1nx\alpha_{1}^{n}\to x (from the above, setting kn=1k_{n}=1 for all nn), this implies that (x)o\mathcal{F}(x)\subseteq\mathcal{L}_{o}.

Since (xn)n\mathcal{F}(x^{n})\subseteq\mathcal{L}_{n} we have o\mathcal{L}\subseteq\mathcal{L}_{o}. If (x)(S˙)\mathcal{F}(x)\in\mathcal{EL}(\dot{S}) then it is the unique minimal sublamination of o\mathcal{L}_{o}, and since o\mathcal{L}\subseteq\mathcal{L}_{o}, we have (x)\mathcal{F}(x)\subseteq\mathcal{L}. If (x)(W)\mathcal{F}(x)\in\mathcal{EL}(W) for some proper witness, then either (x)\mathcal{F}(x) is the unique minimal sublamination of o\mathcal{L}_{o}, or o\mathcal{L}_{o} contains (x)W\mathcal{F}(x)\cup\partial W. Since W\partial W does not intersect the interior of WW, whereas o\mathcal{L}\subset\mathcal{L}_{o} is a sublamination that does nontrivially intersects the interior of WW, it follows that (x)\mathcal{F}(x)\subseteq\mathcal{L}. Therefore we have (x)\mathcal{F}(x)\subseteq\mathcal{L} in both cases, and so \mathcal{F} is continuous.

To prove continuity of 𝒢=1\mathcal{G}=\mathcal{F}^{-1}, suppose nCH\mathcal{L}_{n}\xrightarrow{\text{CH}}\mathcal{L}, and we must show 𝒢(n)𝒢()\mathcal{G}(\mathcal{L}_{n})\to\mathcal{G}(\mathcal{L}). We first pick a sequence of curves αkn\alpha^{n}_{k} such that αkn𝒢(n)\alpha^{n}_{k}\rightarrow\mathcal{G}(\mathcal{L}_{n}) in 𝒞¯s(S˙)\bar{\mathcal{C}}^{s}(\dot{S}). Then, αknCHn\alpha^{n}_{k}\xrightarrow{\text{CH}}\mathcal{L}_{n} as kk\rightarrow\infty, by the first part of the proof, and after passing to subsequences as necessary, we may assume: (i) αknHn\alpha_{k}^{n}\stackrel{{\scriptstyle H}}{{\to}}\mathcal{L}_{n}^{\prime} as kk\to\infty, and hence nn\mathcal{L}_{n}\subseteq\mathcal{L}_{n}^{\prime} for all nn; (ii) dH(αkn,n)<1nd_{H}(\alpha_{k}^{n},\mathcal{L}_{n}^{\prime})<\tfrac{1}{n} for all kk; and (iii) αkn,αnomin{k,}+n\langle\alpha_{k}^{n},\alpha_{\ell}^{n}\rangle_{o}\geq\min\{k,\ell\}+n, for all k,,nk,\ell,n.

Now pass to any Hausdorff convergent subsequence nH\mathcal{L}_{n}^{\prime}\stackrel{{\scriptstyle H}}{{\to}}\mathcal{L}^{\prime}. It suffices to show that for this subsequence 𝒢(n)𝒢()\mathcal{G}(\mathcal{L}_{n})\to\mathcal{G}(\mathcal{L}). Observe that we also have \mathcal{L}\subseteq\mathcal{L}^{\prime} and by (ii) above we also have αknn\alpha_{k_{n}}^{n}\to\mathcal{L}^{\prime} as nn\to\infty, for any sequence {kn}\{k_{n}\}. Thus, for example, we can conclude that α1nCH\alpha_{1}^{n}\stackrel{{\scriptstyle CH}}{{\to}}\mathcal{L}, and so by the first part of the proof we have α1n𝒢()\alpha_{1}^{n}\to\mathcal{G}(\mathcal{L}).

As equivalence classes of sequences, we thus have 𝒢(n)=[{αkn}]\mathcal{G}(\mathcal{L}_{n})=[\{\alpha_{k}^{n}\}] and 𝒢()=[{α1m}]\mathcal{G}(\mathcal{L})=[\{\alpha_{1}^{m}\}]. We further observe that by hyperbolicity and the conditions above, for all k,n,mk,n,m we have

α1m,αknomin{α1m,α1no,α1n,αkno}min{α1m,α1no,1+n}.\langle\alpha_{1}^{m},\alpha_{k}^{n}\rangle_{o}\succeq\min\{\langle\alpha_{1}^{m},\alpha_{1}^{n}\rangle_{o},\langle\alpha_{1}^{n},\alpha_{k}^{n}\rangle_{o}\}\geq\min\{\langle\alpha_{1}^{m},\alpha_{1}^{n}\rangle_{o},1+n\}.

Therefore,

supmlim infk,nα1m,αknosupmlim infnα1m,α1no=,\sup_{m}\liminf_{k,n\to\infty}\langle\alpha_{1}^{m},\alpha_{k}^{n}\rangle_{o}\succeq\sup_{m}\liminf_{n\to\infty}\langle\alpha_{1}^{m},\alpha_{1}^{n}\rangle_{o}=\infty,

from which it follows that 𝒢(n)𝒢()\mathcal{G}(\mathcal{L}_{n})\to\mathcal{G}(\mathcal{L}), as required. This completes the proof. ∎

Proof of Theorem 1.6.

Let :𝒞s(S˙)s(S˙)\mathcal{F}\colon\partial{\mathcal{C}}^{s}(\dot{S})\to\mathcal{EL}^{s}(\dot{S}) be the homeomorphism from Theorem 6.1. It suffices to show that \mathcal{F} is PMod(S˙)\operatorname{PMod}(\dot{S})–equivariant. For this, let fPMod(S˙)f\in\operatorname{PMod}(\dot{S}) be any mapping class and x𝒞s(S˙)x\in\partial{\mathcal{C}}^{s}(\dot{S}) any boundary point. If {αn}𝒞s(S˙)\{\alpha_{n}\}\subset{\mathcal{C}}^{s}(\dot{S}) is any sequence with αnx\alpha_{n}\to x in 𝒞¯s(S˙)\bar{\mathcal{C}}^{s}(\dot{S}), then fαnfxf\cdot\alpha_{n}\to f\cdot x since ff acts by isometries on 𝒞s(S˙){\mathcal{C}}^{s}(\dot{S}). Applying Theorem 6.1 to the sequence {fαn}\{f\cdot\alpha_{n}\} we see that fαnCH(fx)f\cdot\alpha_{n}\stackrel{{\scriptstyle CH}}{{\to}}\mathcal{F}(f\cdot x). On the other hand we also have fαnCHf(x)f\cdot\alpha_{n}\stackrel{{\scriptstyle CH}}{{\to}}f\cdot\mathcal{F}(x), since ff acts by homeomorphisms on the space of laminations with the coarse Hausdorff topology. Therefore, f(x)=(fx)f\cdot\mathcal{F}(x)=\mathcal{F}(f\cdot x), 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 𝒞(S)𝒫(S){\mathcal{C}}(S)\to\mathcal{PML}(S) given by sending a point in the interior of the simplex {v0,,vk}\{v_{0},\ldots,v_{k}\} with barycentric coordinates (s0,,sk)(s_{0},\ldots,s_{k}) to the projective class, [s0v0++skvk][s_{0}v_{0}+\cdots+s_{k}v_{k}]; here we are viewing s0v0+skvks_{0}v_{0}+\cdots s_{k}v_{k} as a measured geodesic lamination with support v0vkv_{0}\cup\ldots\cup v_{k} and with the transverse counting measure scaled by sis_{i} on the ithi^{th} component, for each ii. We denote the image by 𝒫𝒞(S)\mathcal{PML}_{{\mathcal{C}}}(S), which by construction admits a bijective map 𝒫𝒞(S)𝒞(S)\mathcal{PML}_{\mathcal{C}}(S)\to{\mathcal{C}}(S) (inverse to the inclusion above).

By Theorem 2.12, 𝒞(S)(S)\partial\mathcal{C}(S)\cong\mathcal{EL}(S), and so it is natural to define

𝒫𝒞¯(S)=𝒫𝒞(S)𝒫(S),\mathcal{PML}_{\bar{\mathcal{C}}}(S)=\mathcal{PML}_{\mathcal{C}}(S)\cup\mathcal{PFL}(S),

and we extend the bijection 𝒫𝒞(S)𝒞(S)\mathcal{PML}_{\mathcal{C}}(S)\to{\mathcal{C}}(S) to a surjective map

𝒫𝒞¯(S)𝒞¯(S)\mathcal{PML}_{\overline{\mathcal{C}}}(S)\rightarrow\overline{\mathcal{C}}(S)

By Proposition 2.11 and Theorem 2.12, this is continuous at every point of 𝒫(S)\mathcal{PFL}(S).

Similar to the survival map Φ~\tilde{\Phi} defined in Section 2.4, we can define a map

Ψ~:𝒫(S)×Diff0(S)𝒫(S˙).\widetilde{\Psi}\colon\mathcal{PML}(S)\times\operatorname{Diff}_{0}(S)\to\mathcal{PML}(\dot{S}).

This is defined by exactly the same procedure as in Section 2.4 of [LMS11], which goes roughly as follows: If μ\mu is a measured lamination with no closed leaves in its support |μ||\mu|, and if f(z)|μ|f(z)\not\in|\mu|, then Ψ~(μ,f)=f1(μ)\widetilde{\Psi}(\mu,f)=f^{-1}(\mu). When |μ||\mu| contains closed leaves we replace those with the foliated annular neighborhoods of such curves defined in Section 2.4). When the f(z)f(z) lies on a leaf of |μ||\mu| (or the modified |μ||\mu| when there are closed leaves) we ``split |μ||\mu| apart at f(z)f(z)", then take the f1f^{-1}–image. The same proof as that given in [LMS11, Proposition 2.9] shows that Ψ~\widetilde{\Psi} is continuous.

As in Section 2.4 (and in [LMS11]) via the lifted evaluation map ev~:Diff0(S)\widetilde{{\rm{ev}}}\colon\operatorname{Diff}_{0}(S)\to\mathbb{H}, given by ev~(f)=f~(z~)\widetilde{{\rm{ev}}}(f)=\tilde{f}(\tilde{z}) (for f~\tilde{f} the canonical lift), the map Ψ~\widetilde{\Psi} descends to a continuous, π1S\pi_{1}S–equivariant map Ψ\Psi making the following diagram commute:

𝒫(S)×Diff0(S)\mathcal{PML}(S)\times\operatorname{Diff}_{0}(S)𝒫(S)×\mathcal{PML}(S)\times\mathbb{H}𝒫(S˙)\mathcal{PML}(\dot{S}).Ψ~\widetilde{\Psi}Ψ\Psiid𝒫(S)×ev~{\rm{id}}_{\mathcal{PML}(S)}\times\widetilde{{\rm{ev}}}

By construction, the restriction Ψ𝒞=Ψ|𝒫𝒞(S)×\Psi_{\mathcal{C}}=\Psi|_{\mathcal{PML}_{\mathcal{C}}(S)\times\mathbb{H}} and Φ\Phi agree after composing with the bijection between 𝒫𝒞(S)\mathcal{PML}_{\mathcal{C}}(S) and 𝒞(S){\mathcal{C}}(S) in the first factor. Since Φ\Phi maps 𝒞(S)×{\mathcal{C}}(S)\times\mathbb{H} onto 𝒞s(S˙){\mathcal{C}}^{s}(\dot{S}), if we define 𝒫𝒞s(S˙)\mathcal{PML}_{{\mathcal{C}}^{s}}(\dot{S}) to be the image of 𝒞s(S˙){\mathcal{C}}^{s}(\dot{S}) via the map 𝒞s(S˙)𝒫(S˙){\mathcal{C}}^{s}(\dot{S})\to\mathcal{PML}(\dot{S}) defined similarly to the one above, then the following diagram of π1S\pi_{1}S–equivariant maps commutes, with the vertical arrows being bijections

(4) 𝒫𝒞(S)×\textstyle{\mathcal{PML}_{\mathcal{C}}(S)\times\mathbb{H}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ψ𝒞\scriptstyle{\quad\Psi_{\mathcal{C}}}𝒫𝒞s(S˙)\textstyle{\mathcal{PML}_{{\mathcal{C}}^{s}}(\dot{S})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒞(S)×\textstyle{{\mathcal{C}}(S)\times\mathbb{H}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Φ\scriptstyle{\quad\Phi}𝒞s(S˙)\textstyle{{\mathcal{C}}^{s}(\dot{S})}

Similar to 𝒫𝒞¯(S)=𝒫𝒞(S)𝒫(S)\mathcal{PML}_{\bar{\mathcal{C}}}(S)=\mathcal{PML}_{\mathcal{C}}(S)\cup\mathcal{PFL}(S) above, we define

𝒫𝒞¯s(S˙)=𝒫𝒞s(S˙)𝒫s(S˙),\mathcal{PML}_{\bar{\mathcal{C}}^{s}}(\dot{S})=\mathcal{PML}_{{\mathcal{C}}^{s}}(\dot{S})\cup\mathcal{PFL}^{s}(\dot{S}),

where, recall, 𝒫s(S˙)\mathcal{PFL}^{s}(\dot{S}) is the space of measured laminations on S˙\dot{S} whose support is contained in s(S˙)\mathcal{EL}^{s}(\dot{S}). Then Ψ𝒞\Psi_{\mathcal{C}} extends to a map

Ψ𝒞¯:𝒫𝒞¯(S)×𝒫𝒞¯s(S˙).\Psi_{\bar{\mathcal{C}}}\colon\mathcal{PML}_{\bar{\mathcal{C}}}(S)\times\mathbb{H}\to\mathcal{PML}_{\bar{\mathcal{C}}^{s}}(\dot{S}).

The fact that Ψ([μ],w)\Psi([\mu],w) is in 𝒫s(S˙)\mathcal{PFL}^{s}(\dot{S}) for any ww\in\mathbb{H} and [μ]𝒫(S)[\mu]\in\mathcal{PFL}(S) is straightforward from the definition (c.f. [LMS11, Proposition 2.12]): for generic ww, Ψ([μ],w)\Psi([\mu],w) is obtained from [μ][\mu] by adding the zz–puncture in one of the complementary components of |μ||\mu| and adjusting by a homeomorphism. With this, it follows that the map Φ\Phi extends to a map Φ^\hat{\Phi} making the following diagram, extending (4), commute.

𝒫𝒞¯(S)×\textstyle{\mathcal{PML}_{\bar{\mathcal{C}}}(S)\times\mathbb{H}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ψ𝒞¯\scriptstyle{\quad\Psi_{\bar{\mathcal{C}}}}𝒫𝒞¯s(S˙)\textstyle{\mathcal{PML}_{\bar{\mathcal{C}}^{s}}(\dot{S})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒞¯(S)×\textstyle{\bar{\mathcal{C}}(S)\times\mathbb{H}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Φ^\scriptstyle{\quad\hat{\Phi}}𝒞¯s(S˙)\textstyle{\bar{\mathcal{C}}^{s}(\dot{S})}

We will call the map Φ^:𝒞¯(S)×𝒞s¯(S˙)\hat{\Phi}:\bar{{\mathcal{C}}}(S)\times\mathbb{H}\rightarrow{\overline{\mathcal{C}^{s}}(\dot{S})} the extended survival map. Vertical maps in the diagram are natural maps which take projective measured laminations to their supports and they send 𝒫(S)×\mathcal{PFL}(S)\times\mathbb{H} onto (S)×\mathcal{EL}(S)\times\mathbb{H} and 𝒫s(S˙)\mathcal{PFL}^{s}(\dot{S}) onto s(S˙)\mathcal{EL}^{s}(\dot{S}).

Lemma 7.1.

The extended survivial map Φ^\hat{\Phi} is π1S\pi_{1}S–equivariant and is continuous at every point of 𝒞(S)\partial\mathcal{C}(S).

Proof.

To prove the continuity statement, we use the homeomorphism \mathcal{F} from Theorem 6.1 to identity 𝒞s(S˙)\partial{\mathcal{C}}^{s}(\dot{S}) with s(S˙)\mathcal{EL}^{s}(\dot{S}). Now suppose {n}𝒞¯(S)\{\mathcal{L}_{n}\}\in\overline{\mathcal{C}}(S), 𝒞(S)\mathcal{L}\in\partial\mathcal{C}(S) , n\mathcal{L}_{n}\rightarrow\mathcal{L} and {xn}\{x_{n}\} be a sequence in \mathbb{H} such that xnxx_{n}\rightarrow x. Passing to a subsequence, there is a measure μn\mu_{n} on n\mathcal{L}_{n} and a measure μ\mu on \mathcal{L} such that μnμ\mu_{n}\rightarrow\mu in (S)\mathcal{ML}(S). Since Ψ\Psi is continuous on 𝒫(S)×\mathcal{PFL}(S)\times\mathbb{H}

Ψ([μn],xn)Ψ([μ],x).\Psi([\mu_{n}],x_{n})\rightarrow\Psi([\mu],x).

By Proposition 2.11 this implies,

|Ψ(μn,xn)|CH|Ψ(μ,x)|.|\Psi(\mu_{n},x_{n})|\stackrel{{\scriptstyle CH}}{{\longrightarrow}}|\Psi(\mu,x)|.

On the other hand, Ψ((S)×)s(S˙)\Psi(\mathcal{FL}(S)\times\mathbb{H})\subset\mathcal{FL}^{s}(\dot{S}), and by Theorem 6.1 this means

Φ^(n,xn)Φ^(,x)\hat{\Phi}(\mathcal{L}_{n},x_{n})\to\hat{\Phi}(\mathcal{L},x)

in 𝒞¯s(S˙)\bar{\mathcal{C}}^{s}(\dot{S}), since |Ψ^(μn,xn)|=Φ^(n,xn)|\hat{\Psi}(\mu_{n},x_{n})|=\hat{\Phi}(\mathcal{L}_{n},x_{n}) and |Ψ^(μ,x)|=Φ^(,x)|\hat{\Psi}(\mu,x)|=\hat{\Phi}(\mathcal{L},x).

The π1S\pi_{1}S–equivariance follows from that of Φ\Phi on 𝒞s(S˙){\mathcal{C}}^{s}(\dot{S}) 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 (1,x1),(2,x2)(S)×(\mathcal{L}_{1},x_{1}),(\mathcal{L}_{2},x_{2})\in\mathcal{EL}(S)\times\mathbb{H}. Then Φ^(1,x1)=Φ^(2,x2)\hat{\Phi}(\mathcal{L}_{1},x_{1})=\hat{\Phi}(\mathcal{L}_{2},x_{2}) if and only if 1=2\mathcal{L}_{1}=\mathcal{L}_{2} and x1,x2x_{1},x_{2} are on the same leaf of, or in the same complementary region of, p1(1)p^{-1}(\mathcal{L}_{1})\subset\mathbb{H}.

Suppose that (S)\mathcal{L}\in\mathcal{EL}(S), x𝒫x\in\mathcal{P}\subset\partial\mathbb{H} is a parabolic fixed point, HxH_{x}\subset\mathbb{H} is the horoball based at xx as in Section 2.5, and UU\subset\mathbb{H} is the complementary region of p1()p^{-1}(\mathcal{L}) containing HxH_{x}. Given yUy\in U, choose any fDiff(S)f\in\operatorname{Diff}(S) so that f~(z~)=y\tilde{f}(\tilde{z})=y, so that Φ^(,y)=f1()\hat{\Phi}(\mathcal{L},y)=f^{-1}(\mathcal{L}). Observe that p(U)p(U) is a complementary region of \mathcal{L} containing a puncture (corresponding to xx), and hence Φ^(,y)\hat{\Phi}(\mathcal{L},y) is a lamination with two punctures in the complementary component f1(p(U))f^{-1}(p(U)) (one of which is the zz-puncture). Therefore, Φ^(,y)\hat{\Phi}(\mathcal{L},y) is an ending lamination in a proper witness. More precisely, by Lemma 7.2, we may assume yHxy\in H_{x} without changing the image Φ^(,y)\hat{\Phi}(\mathcal{L},y), and then as in the proof of Lemma 2.7, f1(p(Hx))f^{-1}(\partial p(H_{x})) is the boundary of the witness 𝒲(x)\mathcal{W}(x) which is disjoint from Φ^(,x)\hat{\Phi}(\mathcal{L},x). Thus, Φ^(,y)(𝒲(x))\hat{\Phi}(\mathcal{L},y)\in\mathcal{EL}(\mathcal{W}(x)).

In fact, every ending lamination on a proper witness arises as such an image as the next lemma shows.

Lemma 7.3.

Suppose 0(W)\mathcal{L}_{0}\in\mathcal{EL}(W) is an ending lamination in a proper witness WS˙W\subsetneq\dot{S}. Then there exists (S)\mathcal{L}\in\mathcal{EL}(S), x𝒫x\in\mathcal{P}, a complementary region UU of p1()p^{-1}(\mathcal{L}) containing HxH_{x}, and yHxy\in H_{x} so that 𝒲(x)=W\mathcal{W}(x)=W and Φ^(,y)=0\hat{\Phi}(\mathcal{L},y)=\mathcal{L}_{0}.

Proof.

Note that the inclusion of WS˙W\subset\dot{S} is homotopic through embeddings to a diffeomorphism, after filling in zz (since after filling in zz, W\partial W is peripheral). Consequently, after filling in zz, 0\mathcal{L}_{0} is isotopic to a geodesic ending lamination \mathcal{L} on SS. Let f:SSf\colon S\to S be a diffeomorphism isotopic to the identity with f(0)=f(\mathcal{L}_{0})=\mathcal{L}. Then 0=f1()=Ψ~([μ],f)\mathcal{L}_{0}=f^{-1}(\mathcal{L})=\widetilde{\Psi}([\mu],f) where [μ][\mu] is the projective class of any transverse measure on \mathcal{L}.

Next, observe that f(z)f(z) lies in a complementary region VV of \mathcal{L} which is a punctured polygon (since W\partial W is a simple closed curve disjoint from 0\mathcal{L}_{0} bounding a twice punctured disk including the zz-puncture). Let UU\subset\mathbb{H} be the complementary region of p1()p^{-1}(\mathcal{L}) that projects to VV. Then UU is an infinite sided polygon invariant by a parabolic subgroup fixing some x𝒫x\in\mathcal{P}. Now let f~:\tilde{f}\colon\mathbb{H}\to\mathbb{H} be the canonical lift as in Section 2.4 and let y=f~(z~)y^{\prime}=\tilde{f}(\tilde{z}), so that by definition Ψ~([μ],f)=Ψ([μ],y)=Φ^(,y)\widetilde{\Psi}([\mu],f)=\Psi([\mu],y^{\prime})=\hat{\Phi}(\mathcal{L},y^{\prime}). By Lemma 7.2, for any yHxUy\in H_{x}\subset U, it follows that Φ^(,y)=Φ^(,y)=0\hat{\Phi}(\mathcal{L},y)=\hat{\Phi}(\mathcal{L},y^{\prime})=\mathcal{L}_{0}. From the remarks preceeding this lemma, it follows that 0(𝒲(x))\mathcal{L}_{0}\in\mathcal{EL}(\mathcal{W}(x)). Since (W)(W)=\mathcal{EL}(W)\cap\mathcal{EL}(W^{\prime})=\emptyset, unless W=WW=W^{\prime}, it follows that 𝒲(x)=W\mathcal{W}(x)=W, completing the proof. ∎

8. Universal Cannon–Thurston maps

In this section we will prove the following.


Theorem 1.3 For any vertex v𝒞v\in{\mathcal{C}}, the map Φv:𝒞s(S˙)\Phi_{v}:\mathbb{H}\rightarrow{\mathcal{C}}^{s}(\dot{S}) has a continuous π1(S)\pi_{1}(S)–equivariant extension

Φ¯v:𝕊𝒜1𝒞¯s(S˙)\overline{\Phi}_{v}:\mathbb{H}\cup\mathbb{S}^{1}_{\mathcal{A}}\rightarrow\overline{\mathcal{C}}^{s}(\dot{S})

and the induced map

Φ=Φ¯v|𝕊𝒜1:𝕊𝒜1𝒞s(S˙)\partial\Phi=\overline{\Phi}_{v}|_{\mathbb{S}^{1}_{\mathcal{A}}}:\mathbb{S}^{1}_{\mathcal{A}}\rightarrow\partial{\mathcal{C}}^{s}(\dot{S})

is surjective and does not depend on vv. Moreover, Φ\partial\Phi is equivariant with respect to the action of the pure mapping class group PMod(S˙)\operatorname{PMod}(\dot{S}).


Before proceeding, we describe the subset 𝕊𝒜12\mathbb{S}^{1}_{\mathcal{A}}\subset\partial\mathbb{H}^{2} in Theorem 1.3.

Definition 8.1.

Let YS˙Y\subseteq\dot{S} be a subsurface. A point xx\in\partial\mathbb{H} fills YY if,

  • The image of every geodesic ending in xx projected to S˙\dot{S} intersects every curve which projects to YY,

  • There is a geodesic ray rr\subset\mathbb{H} ending at xx with p(r)Yp(r)\subset Y.

Now let 𝕊𝒜1\mathbb{S}^{1}_{\mathcal{A}}\subset\partial\mathbb{H} be the set of points that fill S˙\dot{S}.

We note that when x𝕊𝒜1x\not\in\mathbb{S}^{1}_{\mathcal{A}}, there is a ray rr ending at xx so that rr is contained in a proper subsurface YS˙Y\subsetneq\dot{S}. The boundary of this subsurface is an essential curve vv and Φv(r)Tv\Phi_{v}(r)\subset T_{v} is a bounded diameter set. Thus, restricting to the set 𝕊𝒜1\mathbb{S}^{1}_{\mathcal{A}} 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 SS 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 Φ\partial\Phi, as well as the discussion of Φ0:𝒞(S)𝒞(S˙)\partial\Phi_{0}\colon\partial{\mathcal{C}}(S)\to\partial{\mathcal{C}}(\dot{S}), 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 v𝒞(S)v\in\mathcal{C}(S), the induced survival map Φv:𝒞s(S˙)\Phi_{v}\colon\mathbb{H}\rightarrow{\mathcal{C}^{s}(\dot{S})} has a continuous, π1(S)\pi_{1}(S)–equivariant extension to

Φ¯v:𝕊𝒜1𝒞¯s(S˙))\overline{\Phi}_{v}:\mathbb{H}\cup\mathbb{S}^{1}_{\mathcal{A}}\rightarrow\overline{\mathcal{C}}^{s}(\dot{S}))

Moreover, the restriction Φv=Φ¯v|𝕊𝒜1:𝕊𝒜1𝒞s(S˙)\partial\Phi_{v}=\overline{\Phi}_{v}|_{\mathbb{S}^{1}_{\mathcal{A}}}\colon\mathbb{S}^{1}_{\mathcal{A}}\to\partial{\mathcal{C}}^{s}(\dot{S}) does not depend on the choice of vv.

By the last claim, we may denote the restriction as Φ:𝕊𝒜1𝒞s(S˙)\partial\Phi\colon\mathbb{S}^{1}_{\mathcal{A}}\to\partial{\mathcal{C}}^{s}(\dot{S}), without reference to the choice of vv. 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 XX and YY be two hyperbolic metric spaces, and F:XYF:X\rightarrow Y a continuous map. Fix a basepoint yYy\in Y and a subset AXA\subset\partial X. Then there is a AA–Cannon-Thurston map

F¯:XAYY\overline{F}:X\cup A\rightarrow Y\cup\partial Y

if and only if for all sAs\in A there is a neighborhood basis iXA\mathcal{B}_{i}\subset X\cup A of ss and a collection of uniformly quasiconvex sets QiYQ_{i}\subset Y such that;

  • F(iX)QiF(\mathcal{B}_{i}\cap X)\subset Q_{i}, and

  • dY(y,Qi)d_{Y}(y,Q_{i})\rightarrow\infty as ii\rightarrow\infty.

Moreover,

iQ¯i=iQi={F¯(s)}\bigcap_{i}\overline{Q}_{i}=\bigcap_{i}\partial Q_{i}=\{\overline{F}(s)\}

determines F¯(s)\overline{F}(s) uniquely, where Qi=Q¯iY\partial Q_{i}=\bar{Q}_{i}\cap\partial Y.

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 γ\gamma in \mathbb{H} so that p(γ)p(\gamma) is a closed geodesic that fills SS (i.e. nontrivially intersects every essential simple closed curve or arc on SS). As in [LMS11], we construct quasi-convex sets from such γ\gamma as follows. Define

𝒳(γ)=Φ(𝒞(S)×γ)\mathcal{X}(\gamma)=\Phi({\mathcal{C}}(S)\times\gamma)

where Φ\Phi is the survival map. Let ±(γ)\mathcal{H}^{\pm}(\gamma) denote the two half spaces bounded by γ\gamma and define the sets

±(γ)=Φ(𝒞(S)×±(γ))\mathscr{H}^{\pm}(\gamma)=\Phi({\mathcal{C}}(S)\times\mathcal{H}^{\pm}(\gamma))

The proofs of the following two facts about these sets are identical to the quoted results in [LMS11].

  • [LMS11, Proposition 3.1]: 𝒳(γ)\mathcal{X}(\gamma), ±(γ)\mathscr{H}^{\pm}(\gamma) are simplicial subcomplexes of 𝒞s(S˙){\mathcal{C}^{s}(\dot{S})} 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).

  • [LMS11, Proposition 3.2]: We have,

    +(γ)(γ)=𝒞s(S˙)\mathscr{H}^{+}(\gamma)\cup\mathscr{H}^{-}(\gamma)={\mathcal{C}^{s}(\dot{S})}

    and

    +(γ)(γ)=𝒳(γ).{\mathscr{H}}^{+}(\gamma)\cap\mathscr{H}^{-}(\gamma)=\mathcal{X}(\gamma).

Now we consider a set {γn}\{\gamma_{n}\} of pairwise disjoint translates of γ\gamma in \mathbb{H} so that the corresponding (closed) half spaces nest:

+(γ1)+(γ2)\mathcal{H}^{+}(\gamma_{1})\supset\mathcal{H}^{+}(\gamma_{2})\supset\cdots

Since the action is properly discontinuously on \mathbb{H}, there is a xx\in\partial\mathbb{H} such that

(5) n=1+(γn)¯={x}.\bigcap^{\infty}_{n=1}\overline{\mathcal{H}^{+}(\gamma_{n})}=\{x\}.

Here, +(γn)¯\overline{\mathcal{H}^{+}(\gamma_{n})} is the closure in ¯\overline{\mathbb{H}}. For such a sequence, we say {γn}\{\gamma_{n}\} nests down on xx.

On the other hand, if rr\subset\mathbb{H} is a geodesic ray ending in some point xx\in\partial\mathbb{H} which is not a parabolic fixed point, p(r)p(r) intersects p(γ)p(\gamma) infinitely many times. Hence, we can find a sequence {γn}\{\gamma_{n}\} which nests down on xx. In particular, for any element x𝕊𝒜1x\in\mathbb{S}^{1}_{\mathcal{A}} has a sequence {γk}\{\gamma_{k}\} that nests down on xx.

The main ingredient in the proof of existence of the extension is the following.

Proposition 8.4.

If {γn}\{\gamma_{n}\} nests down to x𝕊𝒜1x\in\mathbb{S}^{1}_{\mathcal{A}}, then for a basepoint b𝒞s(S˙)b\in{\mathcal{C}^{s}(\dot{S})}, the sets +(γn)\mathscr{H}^{+}(\gamma_{n}) are quasiconvex and we have

ds(b,+(γn))asnd^{s}(b,\mathscr{H}^{+}(\gamma_{n}))\rightarrow\infty\,\,\text{as}\,\,n\rightarrow\infty

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 \mathbb{H}, we have nesting in 𝒞s(S˙){\mathcal{C}}^{s}(\dot{S}),

+(γ1)+(γ2).\mathscr{H}^{+}(\gamma_{1})\supset\mathscr{H}^{+}(\gamma_{2})\supset\cdots.

We must show that for any R>0R>0, there exists N>0N>0 so that ds(b,+(γn))Rd^{s}(b,\mathscr{H}^{+}(\gamma_{n}))\geq R, for all nNn\geq N. The first observation is that because Π:𝒞s(S˙)𝒞(S)\Pi\colon{\mathcal{C}}^{s}(\dot{S})\to{\mathcal{C}}(S) is simplicial (hence 11–Lipschitz), it suffices to find N>0N>0 so that ds(b,+(γn)Π1(BR(Π(b)))Rd^{s}(b,\mathscr{H}^{+}(\gamma_{n})\cap\Pi^{-1}(B_{R}(\Pi(b)))\geq R for all nNn\geq N.

To prove this, one can use an inductive argument to construct an increasing sequence N1<N2<<NR+1N_{1}<N_{2}<\ldots<N_{R+1} so that

𝒳(γNj)Π1(BR(Π(b)))𝒳(γNj+1)=.\mathcal{X}(\gamma_{N_{j}})\cap\Pi^{-1}(B_{R}(\Pi(b)))\cap\mathcal{X}(\gamma_{N_{j+1}})=\emptyset.

Before explaining the idea, we note that this implies that {+(γNj)Π1(BR(Π(b)))}j=1R+1\{\mathscr{H}^{+}(\gamma_{N_{j}})\cap\Pi^{-1}(B_{R}(\Pi(b)))\}_{j=1}^{R+1} are properly nested: a path from bb to +(γNR+1)\mathscr{H}^{+}(\gamma_{N_{R+1}}) inside Π1(BR(Π(b)))\Pi^{-1}(B_{R}(\Pi(b))) must pass through a vertex of +(γNj)\mathscr{H}^{+}(\gamma_{N_{j}}), for each jj, before entering the next set. Therefore, it must contain at least R+1R+1 vertices, and so have length at least RR. This completes the proof by taking N=NR+1N=N_{R+1}, since then a geodesic from bb to a point of +(γNR+1)\mathscr{H}^{+}(\gamma_{N_{R+1}}) will have length at least RR (if it leaves Π1(BR(Π(b)))\Pi^{-1}(B_{R}(\Pi(b))), then it's length is greater than RR).

The main idea to find the sequence N1<N2<NR+1N_{1}<N_{2}<\ldots N_{R+1} is involved in the inductive step. If we have already found N1<N2<Nk1N_{1}<N_{2}<\ldots N_{k-1}, and we want to find NkN_{k}, we suppose there is no such NkN_{k}, and derive a contradiction. For this, assume

𝒳(γNk1)𝒳(γn)Π1(BR(Π(b))),\mathcal{X}(\gamma_{N_{k-1}})\cap\mathcal{X}(\gamma_{n})\cap\Pi^{-1}(B_{R}(\Pi(b)))\neq\emptyset,

for all n>Nk1n>N_{k-1}, and let unu_{n} be a vertex in this intersection. Set vn=Π(un)v_{n}=\Pi(u_{n}), and recall that Φvn1(un)=Un\Phi_{v_{n}}^{-1}(u_{n})=U_{n}\subset\mathbb{H} is a component of the complement a small neighborhood of the preimage in \mathbb{H} of the geodesic representative of vnv_{n} in SS. The fact that un𝒳(γNk1)𝒳(γn)u_{n}\in\mathcal{X}(\gamma_{N_{k-1}})\cap\mathcal{X}(\gamma_{n}) translates into the fact that γNk1Un\gamma_{N_{k-1}}\cap U_{n}\neq\emptyset and γnUn\gamma_{n}\cap U_{n}\neq\emptyset. After passing to subsequences and extracting a limit, we find a geodesic from a point on γNk1\gamma_{N_{k-1}} (or one of its endpoints in \partial\mathbb{H}) to xx, which projects to have empty transverse intersection with vnv_{n} in SS. Since vnv_{n} is contained in the bounded set BR(Π(b))B_{R}(\Pi(b)), any subsequential Hausdorff limit does not contain an ending lamination on SS, 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 x𝕊𝒜1x\in\mathbb{S}^{1}_{\mathcal{A}}, 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 Φ¯v\overline{\Phi}_{v} follows by verifying the hypotheses in Lemma 8.3.

Fix a basepoint b𝒞s(S˙)b\in{\mathcal{C}^{s}(\dot{S})} and let {γn}\{\gamma_{n}\} be a sequence nesting to a point x𝕊𝒜1x\in\mathbb{S}^{1}_{\mathcal{A}}. The collection of sets

{+(γn)¯(𝕊𝒜1)}n=1,\{\overline{\mathcal{H}^{+}(\gamma_{n})}\cap(\mathbb{H}\cup\mathbb{S}^{1}_{\mathcal{A}})\}_{n=1}^{\infty},

is a neighborhood basis of xx in 𝕊𝒜1\mathbb{H}\cup\mathbb{S}^{1}_{\mathcal{A}}. By definition of +(γn)\mathscr{H}^{+}(\gamma_{n})

Φv(+(γn))=Φ({v}×+(γn))+(γn),\Phi_{v}(\mathcal{H}^{+}(\gamma_{n}))=\Phi(\{v\}\times\mathcal{H}^{+}(\gamma_{n}))\subset\mathscr{H}^{+}(\gamma_{n}),

for all nn. By Proposition 8.4, ds(b,+(γn))asnd^{s}(b,\mathscr{H}^{+}(\gamma_{n}))\rightarrow\infty\,\,\text{as}\,\,n\rightarrow\infty. Therefore, by Lemma 8.3 we have a 𝕊𝒜1\mathbb{S}^{1}_{\mathcal{A}}-Cannon–Thurston map Φ¯v\overline{\Phi}_{v} defined on x𝕊𝒜1x\in\mathbb{S}^{1}_{\mathcal{A}} by

{Φ¯v(x)}=n=1+(γn)¯.\{\overline{\Phi}_{v}(x)\}=\bigcap_{n=1}^{\infty}\overline{\mathscr{H}^{+}(\gamma_{n})}.

Since the sets on the right-hand side do not depend on the choice of vv, and since x𝕊𝒜1x\in\mathbb{S}^{1}_{\mathcal{A}}, we also write Φ(x)=Φ¯v(x)\partial\Phi(x)=\overline{\Phi}_{v}(x), and note that Φ:𝕊𝒜1𝒞s(S˙)\partial\Phi\colon\mathbb{S}^{1}_{\mathcal{A}}\to\partial{\mathcal{C}}^{s}(\dot{S}) does not depend on vv. ∎

Observe that for all x𝕊𝒜1x\in\mathbb{S}^{1}_{\mathcal{A}}, we have

(6) Φ(x)=n=1+(γn)\partial\Phi(x)=\bigcap_{n=1}^{\infty}\partial\mathscr{H}^{+}(\gamma_{n})

where {γn}\{\gamma_{n}\} is any sequence nesting down on xx, 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 v𝒞0(S)v\in\mathcal{C}^{0}(S) we have

𝒞s(S˙)Φv()¯\partial{\mathcal{C}^{s}(\dot{S})}\subset\overline{\Phi_{v}(\mathbb{H})}

The analogous statement for SS 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 𝒞s(S˙){\mathcal{C}}^{s}(\dot{S}), to which Klarreich's result does not apply.

Proof.

We first claim that if X𝒞s(S˙)X\subset\partial{\mathcal{C}^{s}(\dot{S})} is closed and PMod(S˙)\operatorname{PMod}(\dot{S})–invariant then either X=X=\emptyset or X=𝒞s(S˙)X=\partial{\mathcal{C}^{s}(\dot{S})}. This is true since the set PA{\rm{PA}} of fixed points of pseudo-Anosov elements of PMod(S˙)\operatorname{PMod}(\dot{S}) is dense in (S˙)\mathcal{EL}(\dot{S}) and (S˙)\mathcal{EL}(\dot{S}) is dense in s(S˙)\mathcal{EL}^{s}(\dot{S}). As a result, PA{\rm{PA}} is dense in 𝒞s(S˙)\partial{\mathcal{C}^{s}(\dot{S})}. Since any nonempty, closed, pure mapping class group invariant subset of 𝒞s(S˙)\partial{\mathcal{C}^{s}(\dot{S})} has to include PA{\rm{PA}}, the claim is true.

Now we will show that 𝒞s(S˙)Φv()¯\partial{\mathcal{C}^{s}(\dot{S})}\cap\overline{\Phi_{v}(\mathbb{H})} contains a PMod(S˙)\operatorname{PMod}(\dot{S})–invariant set. For this, first let PA0PA{\rm{PA}}_{0}\subset{\rm{PA}} be the set of pseudo-Anosov fixed points for elements in π1S<PMod(S˙)\pi_{1}S<\operatorname{PMod}(\dot{S}). Since the π1(S)\pi_{1}(S) action leaves Φv()\Phi_{v}(\mathbb{H}) invariant, and since pseudo-Anosov elements act with north-south dynamics on 𝒞¯s(S˙)\overline{\mathcal{C}}^{s}(\dot{S}), it follows that PA0Φv()¯{\rm{PA}}_{0}\subset\overline{\Phi_{v}(\mathbb{H})}. Next, we need to show that f(PA0)=PA0f({\rm{PA}}_{0})={\rm{PA}}_{0} for fPMod(S˙)f\in\operatorname{PMod}(\dot{S}). For any point xPA0x\in PA_{0}, let γπ1(S)\gamma\in\pi_{1}(S) be a pseudo-Anosov element with γ(x)=x\gamma(x)=x. Then fγf1f\gamma f^{-1} fixes f(x)f(x), but fγf1f\gamma f^{-1} is also a pseudo-Anosov element of π1(S)\pi_{1}(S), since π1(S)\pi_{1}(S) is a normal subgroup of PMod(S˙)\operatorname{PMod}(\dot{S}). So, f(PA0)PA0f({\rm{PA}}_{0})\subset{\rm{PA}}_{0}, since xPA0x\in{\rm{PA}}_{0} was arbitrary. Applying the same argument to f1f^{-1}, we find f(PA0)=PA0f({\rm{PA}}_{0})={\rm{PA}}_{0}. Since fPMod(S˙)f\in\operatorname{PMod}(\dot{S}) was arbitrary, PA0{\rm{PA}}_{0} is PMod(S˙)\operatorname{PMod}(\dot{S})–invariant.

Therefore, PA¯0\overline{\rm{PA}}_{0} is a nonempty closed PMod(S˙)\operatorname{PMod}(\dot{S})–invariant subset of 𝒞s(S˙)Φv()¯\partial{\mathcal{C}^{s}(\dot{S})}\cap\overline{\Phi_{v}(\mathbb{H})}, and so both of these sets equal 𝒞s(S˙)\partial{\mathcal{C}^{s}(\dot{S})}. ∎

To prove the surjectivity, we will need the following proposition. The exact analogue for SS 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 𝒫\mathcal{P}\subset\mathbb{H} denotes the set of parabolic fixed points; see Section 2.5.

Proposition 8.6.

If {xn}\{x_{n}\} is a sequence of points in \mathbb{H} with limit x𝕊𝒜1x\in\partial\mathbb{H}{\smallsetminus}\mathbb{S}^{1}_{\mathcal{A}}, then one of the following holds:

  1. (1)

    Φv(xn)\Phi_{v}(x_{n}) does not converge to a point of 𝒞s(S˙)\partial{\mathcal{C}^{s}(\dot{S})}; or

  2. (2)

    x𝒫x\in\mathcal{P} and Φv(xn)\Phi_{v}(x_{n}) accumulates only on points in 𝒞(𝒲(x))\partial{\mathcal{C}}(\mathcal{W}(x)).

To prove this, we will need the following lemma. For the remainder of this paper, we identify the points of 𝒞s(S˙)\partial{\mathcal{C}}^{s}(\dot{S}) with s(S˙)\mathcal{EL}^{s}(\dot{S}) via Theorem 6.1.

Lemma 8.7.

Suppose x𝒫x\in\mathcal{P} and {xn}\{x_{n}\}\subset\mathbb{H} with xnxx_{n}\to x. If Φv(xn)\Phi_{v}(x_{n})\to\mathcal{L} in 𝒞s(S˙)\partial{\mathcal{C}^{s}(\dot{S})}, then 𝒞(𝒲(x))\mathcal{L}\in\partial{\mathcal{C}}(\mathcal{W}(x)).

Proof.

We suppose Φv(xn)\Phi_{v}(x_{n})\to\mathcal{L} in 𝒞s(S˙)\partial{\mathcal{C}^{s}(\dot{S})}. Let H=HxH=H_{x}\subset\mathbb{H} be the horoball based at xx disjoint from all chosen neighborhoods of geodesics used to define Φ\Phi as in Sections 2.4 and 2.5. Applying an isometry if necessary, we can assume that x=x=\infty in the upper-half plane model and H={zIm(z)1}H=\{z\in\mathbb{C}\mid{\rm{Im}}(z)\geq 1\} is stabilized by the cyclic, parabolic group g<π1(S,z)\langle g\rangle<\pi_{1}(S,z). By Lemma 2.7, the Φv\Phi_{v}–image of HH is a single point Φv(H)={u}\Phi_{v}(H)=\{u\}. Note that if Im(xn)>ϵ>0{\rm{Im}}(x_{n})>\epsilon>0 for some ϵ>0\epsilon>0, then Φv(xn)\Phi_{v}(x_{n}) remains a bounded distance from uu, and hence does not converge to any 𝒞s(S˙)\mathcal{L}\in\partial{\mathcal{C}^{s}(\dot{S})}. Therefore, it must be that Im(xn)0{\rm{Im}}(x_{n})\to 0 and consequently Re(xn)±{\rm{Re}}(x_{n})\to\pm\infty.

We may pass to a subsequence so that the hyperbolic geodesics [xn,xn+1][x_{n},x_{n+1}] nontrivially intersect HH, and from this find a sequence of points ynHy_{n}\in H and curves vn𝒞(S)v_{n}\in{\mathcal{C}}(S) so that un=Φ(vn,yn)u_{n}=\Phi(v_{n},y_{n})\to\mathcal{L} as nn\to\infty (c.f. the proof of [LMS11, Proposition 3.11]). According to Lemma 2.7, Φ(𝒞(S)×H)=𝒞(𝒲(x))\Phi({\mathcal{C}}(S)\times H)={\mathcal{C}}(\mathcal{W}(x)), and so Φ(vn,yn)𝒞(𝒲(x))\Phi(v_{n},y_{n})\in{\mathcal{C}}(\mathcal{W}(x)). Consequently, 𝒞(𝒲(x))\mathcal{L}\in\partial{\mathcal{C}}(\mathcal{W}(x)), as required. ∎

Proof of Proposition 8.6.

We suppose Φv(xn)𝒞s(S˙)\Phi_{v}(x_{n})\to\mathcal{L}\in\partial{\mathcal{C}^{s}(\dot{S})} and argue as in [LMS11]. Specifically, the assumption that x𝕊𝒜1x\in\partial\mathbb{H}{\smallsetminus}\mathbb{S}^{1}_{\mathcal{A}} means that a ray rr ending at xx, after projecting to SS, is eventually trapped in some proper, π1\pi_{1}–injective subsurface YSY\subset S, and fills YY if YY is not an annulus. If YY 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 YY is an annular neighborhood of a puncture, then by Lemma 8.7, 𝒞(𝒲(x))\mathcal{L}\in\partial{\mathcal{C}}(\mathcal{W}(x)), as required. ∎

Theorem 8.8.

The Cannon–Thurston map

Φ:𝕊𝒜1𝒞s(S˙)\partial\Phi:\mathbb{S}^{1}_{\mathcal{A}}\rightarrow\partial{\mathcal{C}^{s}(\dot{S})}

is surjective and PMod(S˙)\operatorname{PMod}(\dot{S})–equivariant.

Proof.

Let 𝒞s(S˙)\mathcal{L}\in\partial{\mathcal{C}^{s}(\dot{S})}. Then, by Lemma 8.5 =limΦv(xn)\mathcal{L}=\lim\Phi_{v}(x_{n}) for some sequence {xn}\{x_{n}\}\in\mathbb{H}. Passing to a subsequence, assume that xnxx_{n}\rightarrow x in \partial\mathbb{H}. If x𝕊𝒜1x\in\mathbb{S}^{1}_{\mathcal{A}} we are done since by continuity at every point of 𝕊𝒜1\mathbb{S}^{1}_{\mathcal{A}} we have,

=limΦv(xn)=Φ¯v(x)=Φ(x).\mathcal{L}=\lim\Phi_{v}(x_{n})=\overline{\Phi}_{v}(x)=\partial\Phi(x).

If x𝕊𝒜1x\notin\mathbb{S}^{1}_{\mathcal{A}}, then by Proposition 8.6, x𝒫x\in\mathcal{P} and 𝒞(W)\mathcal{L}\in\partial{\mathcal{C}}(W), where W=𝒲(x)W=\mathcal{W}(x). By Lemma 6.5, πW(Φv(xn))𝒞(W)\pi_{W}(\Phi_{v}(x_{n}))\to\mathcal{L}\in\partial{\mathcal{C}}(W). Let gπ1(S˙)g\in\pi_{1}(\dot{S}) be the generator of Stabπ1S(x){\rm{Stab}}_{\pi_{1}S}(x). As in the proof of Lemma 8.7, xnx_{n} is not entirely contained in any horoball based at xx, and hence it must be that there exists a sequence {kn}\{k_{n}\} such that gkn(xn)ξg^{k_{n}}(x_{n})\rightarrow\xi where ξ\xi\in\partial\mathbb{H} is some point such that ξx\xi\neq x. Since gg is a Dehn twist in W\partial W, it does not affect πW(Φv(xn))\pi_{W}(\Phi_{v}(x_{n})). Thus πW(Φv(gkn(xn)))\pi_{W}(\Phi_{v}(g^{k_{n}}(x_{n})))\to\mathcal{L} and hence Φv(gkn(xn))\Phi_{v}(g^{k_{n}}(x_{n}))\to\mathcal{L} by another application of Lemma 6.5. But in this case since ξx\xi\neq x does not satisfy either of the possibilities given in Lemma 8.6, and hence ξ𝕊𝒜1\xi\in\mathbb{S}^{1}_{\mathcal{A}}. But this implies

=limnΦv(gkn(xn))=Φ¯v(ξ),\mathcal{L}=\lim_{n\to\infty}\Phi_{v}(g^{k_{n}}(x_{n}))=\overline{\Phi}_{v}(\xi),

again appealing to continuity of Φ¯\overline{\Phi}.

The proof of PMod(S˙)\operatorname{PMod}(\dot{S})–equivariance is identical to the proof of [LMS11, Theorem 1.2]. The idea is to use π1S\pi_{1}S–equivariance, and prove Φ(ϕx)=ϕΦ(x)\partial\Phi(\phi\cdot x)=\phi\cdot\partial\Phi(x) for ϕPMod(S˙)\phi\in\operatorname{PMod}(\dot{S}) and xx in the dense subset of 𝕊𝒜1\mathbb{S}^{1}_{\mathcal{A}} consisting of attracting fixed points of elements δπ1S\delta\in\pi_{1}S whose axes project to filling closed geodesics on SS. The point is that such points xx are attracting fixed points in \partial\mathbb{H} of δ\delta, but their images are also attracting fixed points in 𝒞s(S˙)\partial{\mathcal{C}}^{s}(\dot{S}) since δ\delta is pseudo-Anosov by Kra's Theorem [Kra81], when viewed as an element of PMod(S˙)\operatorname{PMod}(\dot{S}). The fact that ϕ(x)\phi(x) and ϕ(Φ(x))\phi(\partial\Phi(x)) are the attracting fixed points of ϕδϕ1\phi\delta\phi^{-1} in \partial\mathbb{H} and 𝒞s(S˙)\partial{\mathcal{C}}^{s}(\dot{S}), 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 \mathbb{H} of laminations in SS 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 x,y𝕊𝒜1x,y\in\mathbb{S}^{1}_{\mathcal{A}}, Φ(x)=Φ(y)\partial\Phi(x)=\partial\Phi(y) if and only if xx and yy are the ideal endpoints of a leaf or complementary region of p1()p^{-1}(\mathcal{L}) for some (S)\mathcal{L}\in\mathcal{EL}(S).

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) +(γ)(γ)=𝒳(γ)\partial\mathscr{H}^{+}(\gamma)\cap\partial\mathscr{H}^{-}(\gamma)=\partial\mathcal{X}(\gamma)

where as above, γ\gamma is a geodesics that projects to a closed, filling geodesic in SS. The next is the analogue in our setting of [LMS11, Lemma 3.9]. To describe this, recall that the element δπ1S\delta\in\pi_{1}S stabilizing γ\gamma is a pseudo-Anosov mapping class when viewed in PMod(S˙)\operatorname{PMod}(\dot{S}) by a theorem of Kra [Kra81]. Let ±(S˙)𝒞s(S˙)\pm\mathcal{L}\in\mathcal{EL}(\dot{S})\subset\partial{\mathcal{C}}^{s}(\dot{S}) be the attracting and repelling fixed points (i.e. the stable/unstable laminations). Then we have

(8) 𝒳(γ)=Φ^(𝒞(S)×γ){±}\partial\mathcal{X}(\gamma)=\hat{\Phi}(\partial{\mathcal{C}}(S)\times\gamma)\cup\{\pm\mathcal{L}\}

The proofs of these facts are identical to those in [LMS11], and we do not repeat them.

Proof of Theorem 1.5.

Given x,y𝕊𝒜1x,y\in\mathbb{S}^{1}_{\mathcal{A}}, first suppose that there is an ending lamination (S)\mathcal{L}\in\mathcal{EL}(S) and EE\subset\mathbb{H} which is either a leaf or complementary region of p1()p^{-1}(\mathcal{L}), so that xx and yy are ideal vertices of EE. Let {γnx},{γny}\{\gamma^{x}_{n}\},\{\gamma^{y}_{n}\} be π1S\pi_{1}S–translates of γ\gamma that nest down on xx and yy, respectively. Then by (6), we have

Φ(x)=n=1+(γnx) and Φ(y)=n=1+(γny)\partial\Phi(x)=\bigcap_{n=1}^{\infty}\partial\mathscr{H}^{+}(\gamma^{x}_{n})\quad\mbox{ and }\quad\partial\Phi(y)=\bigcap_{n=1}^{\infty}\partial\mathscr{H}^{+}(\gamma^{y}_{n})

By Lemma 7.2, Φ^({}×E)\hat{\Phi}(\{\mathcal{L}\}\times E) is a single point, which we denote Φ^({}×E)=0s(S˙)\hat{\Phi}(\{\mathcal{L}\}\times E)=\mathcal{L}_{0}\in\mathcal{EL}^{s}(\dot{S}). Now observe that because γnx\gamma^{x}_{n} intersects EE for all sufficiently large nn, (8) implies

0n=1Φ^({}×γnx)n=1𝒳(γnx)n=1+(γnx)=Φ(x).\mathcal{L}_{0}\in\bigcap_{n=1}^{\infty}\hat{\Phi}(\{\mathcal{L}\}\times\gamma^{x}_{n})\subset\bigcap_{n=1}^{\infty}\partial\mathcal{X}(\gamma^{x}_{n})\subset\bigcap_{n=1}^{\infty}\partial\mathscr{H}^{+}(\gamma^{x}_{n})=\partial\Phi(x).

Therefore, Φ(x)=0\partial\Phi(x)=\mathcal{L}_{0}. The exact same argument shows Φ(y)=0\partial\Phi(y)=\mathcal{L}_{0}, and hence

Φ(x)=0=Φ(y),\partial\Phi(x)=\mathcal{L}_{0}=\partial\Phi(y),

as required.

Now suppose Φ(x)=Φ(y)=0s(S˙)\partial\Phi(x)=\partial\Phi(y)=\mathcal{L}_{0}\in\mathcal{EL}^{s}(\dot{S}). Again by (6) there are sequences {γnx}\{\gamma^{x}_{n}\} and {γny}\{\gamma^{y}_{n}\} (π1S\pi_{1}S-translates of γ\gamma) nesting down to xx and yy respectively so that

n=1+(γnx)=0=n=1+(γny).\bigcap_{n=1}^{\infty}\partial\mathscr{H}^{+}(\gamma^{x}_{n})=\mathcal{L}_{0}=\bigcap_{n=1}^{\infty}\partial\mathscr{H}^{+}(\gamma^{y}_{n}).

Because the intersections are nested, this implies that for all nn we have

0+(γnx)+(γny).\mathcal{L}_{0}\in\partial\mathscr{H}^{+}(\gamma_{n}^{x})\cap\partial\mathscr{H}^{+}(\gamma_{n}^{y}).

Passing to a subsequence if necessary, we may assume that for all nn, +(γnx)(γny)\mathscr{H}^{+}(\gamma_{n}^{x})\subset\mathscr{H}^{-}(\gamma_{n}^{y}) and +(γny)(γnx)\mathscr{H}^{+}(\gamma_{n}^{y})\subset\mathscr{H}^{-}(\gamma_{n}^{x}). Therefore, for all nn we have

0\displaystyle\mathcal{L}_{0} \displaystyle\in +(γnx)+(γny)\displaystyle\partial\mathscr{H}^{+}(\gamma_{n}^{x})\cap\partial\mathscr{H}^{+}(\gamma_{n}^{y})
=\displaystyle= (+(γnx)(γny))(+(γny)(γnx))\displaystyle\left(\partial\mathscr{H}^{+}(\gamma_{n}^{x})\cap\partial\mathscr{H}^{-}(\gamma_{n}^{y})\right)\cap\left(\partial\mathscr{H}^{+}(\gamma_{n}^{y})\cap\partial\mathscr{H}^{-}(\gamma_{n}^{x})\right)
=\displaystyle= (+(γnx)(γnx))(+(γny)(γny))\displaystyle\left(\partial\mathscr{H}^{+}(\gamma_{n}^{x})\cap\partial\mathscr{H}^{-}(\gamma_{n}^{x})\right)\cap\left(\partial\mathscr{H}^{+}(\gamma_{n}^{y})\cap\partial\mathscr{H}^{-}(\gamma_{n}^{y})\right)
=\displaystyle= 𝒳(γnx)𝒳(γny).\displaystyle\partial\mathcal{X}(\gamma_{n}^{x})\cap\partial\mathcal{X}(\gamma_{n}^{y}).

The last equality here is an application of (7). Combining this with the description of 0\mathcal{L}_{0} above and (8), we have

Φ(x)=Φ(y)=0=n=1(𝒳(γnx)𝒳(γny))=n=1(Φ^(𝒞(S)×γnx)Φ^(𝒞(S)×γny)).\partial\Phi(x)=\partial\Phi(y)=\mathcal{L}_{0}=\bigcap_{n=1}^{\infty}(\partial\mathcal{X}(\gamma_{n}^{x})\cap\partial\mathcal{X}(\gamma_{n}^{y}))=\bigcap_{n=1}^{\infty}(\hat{\Phi}(\partial{\mathcal{C}}(S)\times\gamma_{n}^{x})\cap\hat{\Phi}(\partial{\mathcal{C}}(S)\times\gamma_{n}^{y})).

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 δnx\delta_{n}^{x} and δny\delta_{n}^{y} in π1S\pi_{1}S stabilizing γnx\gamma_{n}^{x} and γny\gamma_{n}^{y}, respectively, are all distinct, hence 0\mathcal{L}_{0} is not one of the stable/unstable laminations.

From the equation above, we have nx,ny(S)\mathcal{L}_{n}^{x},\mathcal{L}_{n}^{y}\in\mathcal{EL}(S) and xnγnx,ynγnyx_{n}\in\gamma_{n}^{x},y_{n}\in\gamma_{n}^{y} so that Φ^(nx,xn)=Φ^(ny,yn)=0\hat{\Phi}(\mathcal{L}_{n}^{x},x_{n})=\hat{\Phi}(\mathcal{L}_{n}^{y},y_{n})=\mathcal{L}_{0}, for all nn. According to Lemma 7.2, there exists (S)\mathcal{L}\in\mathcal{EL}(S) so that nx=ny=\mathcal{L}_{n}^{x}=\mathcal{L}_{n}^{y}=\mathcal{L} for all nn, and there exists a leaf or complementary region EE of p1()p^{-1}(\mathcal{L}) so that xn,ynEx_{n},y_{n}\in E. Since γnx\gamma_{n}^{x} and γny\gamma_{n}^{y} nest down on xx and yy, respectively, it follows that xnxx_{n}\to x and ynyy_{n}\to y as nn\to\infty. Therefore, x,yx,y are endpoints of a leaf of p1()p^{-1}(\mathcal{L}) or ideal endpoints of a complementary region of p1()p^{-1}(\mathcal{L}), as required. ∎

We can now easily deduce the following, which also proves Proposition 1.4.

Proposition 8.9.

Given 0s(S˙)\mathcal{L}_{0}\in\mathcal{EL}^{s}(\dot{S}), Φ1(0)\partial\Phi^{-1}(\mathcal{L}_{0}) is infinite if and only if 0(W)\mathcal{L}_{0}\in\mathcal{EL}(W) for some proper witness WW.

Proof.

Theorem 1.5 implies that for 0s(S˙)\mathcal{L}_{0}\in\mathcal{EL}^{s}(\dot{S}), Φ1(0)\partial\Phi^{-1}(\mathcal{L}_{0}) contains more than two points if and only if there is a lamination (S)\mathcal{L}\in\mathcal{EL}(S) and a complementary region UU of p1()p^{-1}(\mathcal{L}) so that Φ1(0)\partial\Phi^{-1}(\mathcal{L}_{0}) is precisely the set of ideal points of UU. Moreover, in this case the proof above shows that 0=Φ^({}×U)\mathcal{L}_{0}=\hat{\Phi}(\{\mathcal{L}\}\times U).

On the other hand, Lemma 7.3 and the paragraph preceding it tell us that 0s(S˙)\mathcal{L}_{0}\in\mathcal{EL}^{s}(\dot{S}) is contained in (W)\mathcal{EL}(W) for a proper witness WS˙W\subsetneq\dot{S} if and only if it is given by 0=Φ^({}×U)\mathcal{L}_{0}=\hat{\Phi}(\{\mathcal{L}\}\times U) where (S)\mathcal{L}\in\mathcal{EL}(S) and UU is the complementary region of p1()p^{-1}(\mathcal{L}) containing HxH_{x}, where x𝒫x\in\mathcal{P} with W=𝒲(x)W=\mathcal{W}(x).

Finally, we note that a complementary region of a lamination (S)\mathcal{L}\in\mathcal{EL}(S) has infinitely many ideal vertices if and only if it projects to a complementary region of \mathcal{L} containing a puncture, and this happens if and only if it contains a horoball HxH_{x} for some x𝒫x\in\mathcal{P}.

Combining all three of the facts above proves the proposition. ∎

Now we define 𝕊𝒜01𝕊𝒜1\mathbb{S}^{1}_{\mathcal{A}_{0}}\subset\mathbb{S}^{1}_{\mathcal{A}} to be those points that map by Φ\partial\Phi to (S˙)\mathcal{EL}(\dot{S}) and then define Φ0:𝕊𝒜01𝒞(S˙)=(S˙)\partial\Phi_{0}\colon\mathbb{S}^{1}_{\mathcal{A}_{0}}\to\partial{\mathcal{C}}(\dot{S})=\mathcal{EL}(\dot{S}) to be the ``restriction'' of Φ\partial\Phi to 𝕊𝒜01\mathbb{S}^{1}_{\mathcal{A}_{0}}. Theorem 1.2 is a consequence of the Theorem 1.5 since Φ0\partial\Phi_{0} is the restriction of Φ\partial\Phi to 𝕊𝒜01\mathbb{S}^{1}_{\mathcal{A}_{0}}. 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 𝕊𝒜01\mathbb{S}^{1}_{\mathcal{A}_{0}}. For (S)\mathcal{L}\in\mathcal{EL}(S), consider the subset 𝒮\mathcal{S}_{\mathcal{L}}\subset\partial\mathbb{H} consisting of all ideal endpoints of complementary components of p1()p^{-1}(\mathcal{L}) which have infinitely many such ideal endpoints. That is, 𝒮\mathcal{S}_{\mathcal{L}} is the set of ideal endpoints of complementary regions that project to complementary regions of \mathcal{L} 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 𝕊𝒜01𝕊𝒜1\mathbb{S}^{1}_{\mathcal{A}_{0}}\subset\mathbb{S}^{1}_{\mathcal{A}}\subset\partial\mathbb{H} that map to (S˙)s(S˙)\mathcal{EL}(\dot{S})\subset\mathcal{EL}^{s}(\dot{S}) is

𝕊𝒜01=𝕊𝒜1(S)𝒮.\mathbb{S}^{1}_{\mathcal{A}_{0}}=\mathbb{S}^{1}_{\mathcal{A}}{\smallsetminus}\bigcup_{\mathcal{L}\in\mathcal{EL}(S)}\mathcal{S}_{\mathcal{L}}.

\Box

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 33. 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 33-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 FNF_{N}. 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 33-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 33-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.