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

L2L^{2}-bounds for drilling short geodesics in convex co-compact hyperbolic 3-manifolds

Martin Bridgeman  and Kenneth Bromberg M. Bridgeman’s research was supported by NSF grants DMS-1500545, DMS-2005498 and institutional NSF grant DMS-1928930 while the author participated in the Fall 2020 program at MSRI. This work was also supported by a grant from the Simons Foundation (675497, Bridgeman)K. Bromberg’s research supported by NSF grant DMS-1509171, DMS-1906095.
Abstract

We give L2L^{2}-bounds on the change in the complex projective structure on the boundary of conformally compact hyperbolic 3-manifold with incompressible boundary after drilling short geodesics. We show that the change is bounded by a universal constant times the square root of the length of the drilled geodesics. While LL^{\infty}-bounds of this type where obtained in [Bro], our bounds here do not depend on the injectivity radius of the boundary.

1 Introduction

Given a complete, hyperbolic 3-manifold MM and a collection 𝒞\mathcal{C} of disjoint simple closed geodesics in MM, the manifold M𝒞M-\mathcal{C} also supports a complete hyperbolic structure M^\hat{M}. If we insist that MM and M^\hat{M} have the same ending data then M^\hat{M} is unique. If MM is closed, or more generally finite volume, and the elements of 𝒞\mathcal{C} are sufficiently short then Hodgson-Kerckhoff [HK] developed a theory of a hyperbolic cone-manifolds that allows one to continuously interpolate between MM and M^\hat{M} through cone-manifolds. These methods were extended to conformally compact manifolds in [Brm]. By controlling the derivative of this family of cone-manifolds one can obtain comparisons between the geometry of MM and M^\hat{M}.

One can give precise meaning to comparing the geometry of MM and M^\hat{M}. For example, one can compare the length of curves in MM to those in M^\hat{M}. In this paper we will be interested in measuring the change in the projective boundary between the two manifolds. This change is described by a holomorphic quadratic differential given by taking the Schwarzian derivative. The size of this quadratic differential can be measured by taking an LpL^{p}-norm. In [Bro], the second author bounded the LL^{\infty}-norm and these bounds played an important role the in resolving the Bers density conjecture. While LL^{\infty}-bounds always imply LpL^{p}-bounds for all pp, the bounds in [Bro] depended on both the length of the curves being drilled and the injectivity radius of the hyperbolic metric on the boundary. In this paper, we obtain L2L^{2}-bounds on the change in the projective structure that are proportional to the square-root of the total length of the geodesics to be drilled but are independent of the injectivity radius. In particular, this gives uniform control on the L2L^{2} change for drilling short geodesics. In [BBB], this result is used to study the Weil-Petersson gradient flow of renormalized volume and obtain lower bounds on the renormalized volume of a convex cocompact hyperbolic manifold with incompressible boundary in terms of the Weil-Petersson distance between its boundary components.

We have the following setup: N¯\bar{N} will be a compact, hyperbolizable 3-manifold with boundary with interior NN and 𝒞\mathcal{C} will be a collection of disjoint simple closed curves in NN. The M¯\bar{M} is a complete, conformally compact hyperbolic structure on N¯\bar{N} where the curves in 𝒞\mathcal{C} are geodesics and M¯t\bar{M}_{t} is a one parameter family of hyperbolic cone-manifolds with cone locus 𝒞\mathcal{C} and cone angles tt. We also assume the conformal boundary cMt\partial_{c}M_{t} is fixed throughout the definition.

Theorem 1.1 ([Bro, Theorem 1.2])

There exists an L0>0L_{0}>0 such that if all geodesics in 𝒞\mathcal{C} have length L0\leq L_{0} in MM then the cone deformation exists for t[0,2π]t\in[0,2\pi] where M0M_{0} is a complete, hyperbolic structure on N𝒞N-\mathcal{C}.

While the conformal boundary will be a fixed conformal structure XX the deformation, the complex projective structure on XX will change. We denote this one parameter family of projective structures by Σt\Sigma_{t}. The derivative of a path of projective structures on XX is naturally a holomorphic, quadratic differential. We denote the tangent vectors to Σt\Sigma_{t} by the holomorphic quadratic differentials Φt\Phi_{t}. Our main results is the following bound on the L2L^{2}-norm of Φt\Phi_{t}.

Theorem 1.2

If L𝒞L_{\mathcal{C}} is the sum of the length the geodesics in 𝒞\mathcal{C} in M=M2πM=M_{2\pi} then

Φt2cdrillL𝒞.\|\Phi_{t}\|_{2}\leq c_{\rm drill}\sqrt{L_{\mathcal{C}}}.

As an immediate application we obtain the following L2L^{2}-bounds on the change in projective structure.

Theorem 1.3

There exists an L0>0L_{0}>0 and cdrill>0c_{\rm drill}>0 such that the following holds. Let MM be a conformally compact hyperbolic 3-manifold and 𝒞\mathcal{C} a collection of simple closed geodesics in MM each of length L0\leq L_{0}. Let M^\hat{M} be the unique complete hyperbolic structure on M𝒞M-\mathcal{C} such that the inclusion M^M^\hat{M}\hookrightarrow\hat{M} is an isomorphism of conformal boundaries. If Σ\Sigma and Σ^\hat{\Sigma} are the projective structures on the conformal boundaries of MM and M^\hat{M} and the holomorphic quadratic differential Φ=Φ(Σ,Σ^)\Phi=\Phi(\Sigma,\hat{\Sigma}) is Schwarzian derivative between them then

Φ22πcdrillL𝒞\|\Phi\|_{2}\leq 2\pi c_{\rm drill}\sqrt{L_{\mathcal{C}}}

where L𝒞L_{\mathcal{C}} is the sum of the lengths of the components of 𝒞\mathcal{C} in MM.

We note that the L2L^{2}-bounds have universal constants compared to the LL^{\infty}-bound in Theorem 1.3 in [Bro] which depended on injectivity radius of the boundary hyperbolic structure. In [BW], LL^{\infty}-bounds on quadratic differentials are obtained from L2L^{2}-bounds. These bounds again depend on the injectivity radius but they produce stronger bounds than those obtained in [Bro]. However, in [Bro] cone angles >2π>2\pi where allowed which was important for the application to the Bers density conjecture.

We briefly sketch our argument. Following the classical construction of Calabi [Cal] and Weil [Weil] the derivative of the deformation MtM_{t} can be represented by a cohomology class in a certain flat bundle. This bundle has a metric and, in our setting, each cohomology class has a harmonic representative whose L2L^{2}-norm can be bounded by the length of the curves in the cone locus. We would like to use the bound on the L2L^{2}-norm in the 3-manifold to bound the L2L^{2}-norm of the quadratic differentials Φt\Phi_{t} representing the derivative of the projective structures.

To do this we first represent the cohomology class in the ends of the manifold by a certain model deformation which we describe explicitly. This model deformation will differ from the actual deformation by a trivial deformation. For the model deformation we can explicitly calculate the L2L^{2}-norm on the end in terms of the L2L^{2}-norm of the quadratic differential. To calculate the L2L^{2}-norm of the actual harmonic deformation we would like the model deformation to be orthogonal (in the L2L^{2}-inner product) to the trivial deformation. Unfortunately, this is not true, essentially because the model deformation itself is not harmonic. However, we will show that the model deformation is asymptotically harmonic. Using the infinitesimal inflexibility theorem from [BB] we use this asymptotic control to bound the L2L^{2}-norm of the quadratic differential in terms of the L2L^{2}-norm of the deformation of the end.

This would seem to be enough however there is one final complication. Our bounds will depend on how large of an end we can embed in the cone-manifold. This size is controlled by the Schwarzian derivative of the of the projective structure. For smooth hyperbolic manifolds with incompressible boundary the Schwarzian of projective boundary is bounded by a classical theorem of Nehari. In the cone-manifold setting we will not be able to apply Nehari’s theorem. Instead we control the Schwarzian by first controlling the average bending of the boundary of the convex core of the cone manifold. This notion was defined by the first author in [Bri] where it was shown that for smooth hyperbolic 3-manifolds with incompressible boundary the average bending of the boundary of convex core is uniformly (and explicitly) bounded. We see that the argument in [Bri] extends to cone-manifolds (with some restrictions) and then we will derive bounds on the Schwarzian via a compactness argument.

Acknowledgements

The work in this paper was motivated by a joint project with the authors and Jeff Brock. We thank Jeff for many interested discussions related to this paper.

2 Background

The proof relies on an analysis of L2L^{2}-bounds for cohomology classes associated to infinitesimal deformations of hyperbolic cone-manifolds. We now quickly review this theory with an emphasis and what is need for our computations. The original analysis can be found in [Brm] and [Bro] which generalized work of [HK] on the finite volume hyperbolic cone-manifolds to the geometrically finite hyperbolic cone-manifolds.

Let ¯3=3^{\bar{\mathbb{H}}^{3}}={{\mathbb{H}}^{3}}\cup\widehat{{\mathbb{C}}} be the usual compactification of 3{{\mathbb{H}}^{3}} by ^\widehat{{\mathbb{C}}}. Note that isometries of 3{{\mathbb{H}}^{3}} extend to projective automorphisms of ^\widehat{{\mathbb{C}}} and that the group of isometries/projective transformations is 𝖯𝖲𝖫2()\mathsf{PSL}_{2}(\mathbb{C}). If M¯\bar{M} is a 3-manifold with boundary a (𝖯𝖲𝖫2(),¯3)(\mathsf{PSL}_{2}(\mathbb{C}),{\bar{\mathbb{H}}^{3}})-structure on M¯\bar{M} is an atlas of charts to ¯3{\bar{\mathbb{H}}^{3}} with transition maps restrictions of elements of 𝖯𝖲𝖫2()\mathsf{PSL}_{2}(\mathbb{C}). On MM, the interior of M¯\bar{M}, this a hyperbolic structure. On the boundary M¯\partial\bar{M} this is a complex projective structure. In this paper we will be interested in a special class of (𝖯𝖲𝖫2(),¯3)\left(\mathsf{PSL}_{2}(\mathbb{C}),{\bar{\mathbb{H}}^{3}}\right)-structures, conformally compact hyperbolic cone-manifolds.

Let N¯\bar{N} be a compact 3-manifold with boundary with interior NN and let 𝒞\cal C be a collection of simple closed curves in the interior of NN. Let M=N𝒞M=N-{\mathcal{C}}. A hyperbolic cone metric on NN with cone angle α\alpha along 𝒞\mathcal{C} is a hyperbolic metric on the interior of MM whose metric completion is homeomorphic to the interior of NN and in a c of each component the metric is that of a singular hyperbolic metric with cone angle α\alpha. That is in cylindrical coordinates (r,θ,z)(r,\theta,z) the metric will locally have the form

dr2+sinh2(r)dθ2+cosh2(r)dz2.dr^{2}+\sinh^{2}(r)d\theta^{2}+\cosh^{2}(r)dz^{2}.

where θ\theta is measured modulo the cone angle α\alpha and the singular locus is identified with the zz-axis.

The hyperbolic metric is conformally compact if the hyperbolic structure on MM extends to a (𝖯𝖲𝖫2(),¯3)\left(\mathsf{PSL}_{2}(\mathbb{C}),{\bar{\mathbb{H}}^{3}}\right)-structure on M¯=N¯𝒞\bar{M}=\bar{N}-\cal C. We then have:

Theorem 2.1 ([HK, Brm])

Given a cone angle α>0\alpha>0 there exists a length >0\ell>0 such that the following holds. Let (N,g)(N,g) be a conformally compact hyperbolic cone-manifold with all cone angles α\alpha and assume that the tube radius about the singular locus is sinh12\geq\sinh^{-1}\sqrt{2}. If each component of the singular locus has length \leq\ell then t[0,α]t\in[0,\alpha] there a one parameter family of conformally compact hyperbolic cone-manifolds (N,gt)(N,g_{t}) with the conformally boundary fixed and cone angle tt.

This one parameter family of cone-manifolds will induces a one parameter family of projective structures Σt\Sigma_{t} on the boundary where the conformal structure of Σt\Sigma_{t} is fixed. We will be interested in controlling the change in this projective structure as the parameter varies.

2.1 Flat 𝔰𝔩2()\mathfrak{sl}_{2}(\mathbb{C})-bundles

The Lie algebra 𝔰𝔩2()\mathfrak{sl}_{2}(\mathbb{C}) can be interpreted geometrically as the space of infinitesimal automorphisms of ¯3{\bar{\mathbb{H}}^{3}}. These are vector fields on ¯3{\bar{\mathbb{H}}^{3}} whose flow are elements in 𝖯𝖲𝖫2()\mathsf{PSL}_{2}(\mathbb{C}) so that on 3{{\mathbb{H}}^{3}}, the flow will be isometries of the hyperbolic metric, while on ^\widehat{{\mathbb{C}}} the flow will be projective automorphisms. A (𝖯𝖲𝖫2(),¯3)\left(\mathsf{PSL}_{2}(\mathbb{C}),{\bar{\mathbb{H}}^{3}}\right)-structure on M¯\bar{M} determines a flat 𝔰𝔩2()\mathfrak{sl}_{2}(\mathbb{C})-bundle E¯=E(M¯)\bar{E}=E(\bar{M}) over M¯\bar{M}. We examine this bundle when it is restricted to the hyperbolic structure and when it is restricted to the projective boundary.

Hyperbolic structures

Let MM be the interior of M¯\bar{M}. Then a (𝖯𝖲𝖫2(),¯3)\left(\mathsf{PSL}_{2}(\mathbb{C}),{\bar{\mathbb{H}}^{3}}\right)-structure is a hyperbolic structure and the bundle has a natural decomposition and metric structure that we now describe. Each fiber EpE_{p} is the space of germs of infinitesimal isometries. In particular, sEps\in E_{p} is a vector field in a neighborhood of pp so s(p)s(p) is a vector in TpMT_{p}M. As EE is a complex bundle we can multiply ss by ii and then (is)(p)(is)(p) is another vector in TpMT_{p}M. Then the map from EE to TMTMTM\oplus TM given by s(s(p),(is)(p))s\mapsto(s(p),(is)(p)) is bundle isomorphism. In fact, the map ss(p)+i(is)(p)s\mapsto s(p)+i(is)(p) is a complex vector bundle isomorphism from EE to the the complexification TMT^{\mathbb{C}}M of the tangent bundle. This isomorphism from EE to TMTMTM\oplus TM gives a decomposition of sections of EE into real and imaginary parts.

If vv is a vector in TpMT_{p}M, we define v^Ep\hat{v}\in E_{p} such that under our isomorphism from EpE_{p} to TpMTpMT_{p}M\oplus T_{p}M we have v^(v,0)\hat{v}\mapsto(v,0). Then v^\hat{v} is the infinitesimal translation with axis through vv and iv^i\hat{v} is an infinitesimal rotation about vv. Note that v^\hat{v} is real and iv^i\hat{v} is imaginary, as one would expect.

As EE is isomorphic to TMTMTM\oplus TM, the dual bundle EE^{*} is isomorphic to TMTMT^{*}M\oplus T^{*}M. The hyperbolic metric on MM determines an isomorphism from TMTM to TMT^{*}M and therefore an isomorphism from EE to EE^{*}. Note that this isomorphism is \mathbb{R}-linear but is \mathbb{C}-anti-linear with respect to the complex structures on EE and EE^{*}. For sections ss of EE we let ss^{\sharp} be the dual section of EE^{*}. When going from EE^{*} to EE we replace the \sharp with a \flat.

As EpE_{p} is a complex vector space we have Ep=EpE_{p}=E_{p}\otimes\mathbb{C} and more generally for alternating tensors with values in EpE_{p} we have

Λk(TpM;Ep)=EpΛk(TpM)=EpΛk(TpM).\Lambda^{k}(T_{p}M;E_{p})=E_{p}\otimes\Lambda^{k}(T_{p}M)=E_{p}\otimes\Lambda^{k}\left(T^{\mathbb{C}}_{p}M\right).

In particular, every EE-valued form is locally the sum of terms ϕsω\phi s\omega where ϕ\phi is a complex valued function, ss is a section of EE, and ω\omega is a \mathbb{R}-valued form. The \sharp (and \flat) operators extend to EE-valued forms and we have (ϕsω)=ϕ¯sω(\phi s\omega)^{\sharp}=\bar{\phi}s^{\sharp}\omega. We also linearly extend the Hodge star operator from real forms to EE-valued forms so that (ϕsω)=ϕs(ω)\star(\phi s\omega)=\phi s(\star\omega). This extends to a linear map from Ωk(M;E)\Omega^{k}(M;E) to Ω3k(M;E)\Omega^{3-k}(M;E). We the define the inner product

(α,β)=Mα(β).(\alpha,\beta)=\int_{M}\alpha\wedge(\star\beta)^{\sharp}.

Note that the wedge product of an EE-valued form and an EE^{*}-valued form is real form so this is a real inner product. We also let α2=(α,α)\|\alpha\|^{2}=(\alpha,\alpha) be the L2L^{2}-norm of an EE-valued form.

If α\alpha is either an EE-valued or \mathbb{C}-valued form we define the pointwise norm |α||\alpha| by |α|2=(α(α))|\alpha|^{2}=\star(\alpha\wedge(\star\alpha)^{\sharp}). Then α\|\alpha\| is the usual L2L^{2}-norm of the function |α||\alpha|.

If dd^{*} is the flat connection for EE^{*} then define the operator \partial on EE by

ω=(d(ω))β.\partial\omega=\left(d^{*}(\omega^{\sharp})\right)^{\beta}.

Then the formal adjoint δ\delta for dd satisfies the formula

δ=.\delta=\star\partial\star.

Note that if ss is a flat section (ds=0ds=0) then the real and imaginary parts, Res\operatorname{Re}s and Ims\operatorname{Im}s, will not be flat. That is dd will not preserve our bundle decomposition. Instead we define operators DD and TT such that d=D+Td=D+T where DD preserves the bundle decomposition and TT permutes it. That is for a real section ss we have that DsDs is a real EE-valued 1-form while TsTs is imaginary. We have formulas for both DD and TT. If vv is a vector field then

Dv^(w)=wv^D\hat{v}(w)=\widehat{\nabla_{w}v}

where \nabla is the Riemannian connection for the hyperbolic metric and

Tv^(w)=[v^,w^]T\hat{v}(w)=[\hat{v},\hat{w}]

where [,][,] is the Lie bracket. Note that the operator TT is purely algebraic.

We also have

=DT.\partial=D-T.

This is a manifestation of the fact that the \sharp-operator is \mathbb{C}-anti-linear.

The Laplacian for EE-valued 1-forms is Δ=dδ+δd\Delta=d\delta+\delta d and ωΩk(E)\omega\in\Omega^{k}(E) is harmonic if Δω=0\Delta\omega=0. If MM is compact then this is equivalent ω\omega be closed and co-closed. However, our manifolds will be non-compact so we will define ω\omega to be a Hodge form if it is closed, co-closed and the real and imaginary parts are symmetric and traceless.

A computation in EE

We now make a few computations that will be very useful later and will also serve as an example of how to do computations in the bundle. We will work in the upper half space model of 3=×+{{\mathbb{H}}^{3}}=\mathbb{C}\times\mathbb{R}^{+} with {x,y,t}\left\{{\frac{\partial}{\partial x}},{\frac{\partial}{\partial y}},\frac{\partial}{\partial t}\right\} and {dx,dy,dt}\{dx,dy,dt\} the usual basis and dual basis at each Tp3T_{p}{{\mathbb{H}}^{3}}. We also let

z=12(xiy) and z¯=12(x+iy)\frac{\partial}{\partial z}=\frac{1}{2}\left({\frac{\partial}{\partial x}}-i{\frac{\partial}{\partial y}}\right)\mbox{ and }\frac{\partial}{\partial{\overline{z}}}=\frac{1}{2}\left({\frac{\partial}{\partial x}}+i{\frac{\partial}{\partial y}}\right)

be tangent vectors in the complexified tangent space with dual 1-forms dz=dx+idydz=dx+idy and dz¯=dxidyd{\overline{z}}=dx-idy. We can then write any EE-valued 1-form on 3{{\mathbb{H}}^{3}} as a sum of dz,dz¯dz,d{\overline{z}} and dtdt terms.

The Lie algebra 𝔰𝔩2()\mathfrak{sl}_{2}(\mathbb{C}) can be identified with traceless 2-by-2 matrices in \mathbb{C}, projective vector fields on ^\widehat{{\mathbb{C}}} and infinitesimal isometries of 3{{\mathbb{H}}^{3}}. The reader can check the correspondence given in the following lemma:

Lemma 2.2

An element of 𝔰𝔩2()\mathfrak{sl}_{2}(\mathbb{C}) given by the matrix

[abca]\begin{bmatrix}a&b\\ c&-a\end{bmatrix}

is equivalent to the projective vector

2(cz2+2az+b)z.2\left(-cz^{2}+2az+b\right)\frac{\partial}{\partial z}.

Along the axis (0,t)(0,t) in the upper half space model of 3{{\mathbb{H}}^{3}} they are both equivalent to the constant section

2(ct2z¯^+att^+bz^).2\left(ct^{2}\hat{\frac{\partial}{\partial{\overline{z}}}}+at\hat{\frac{\partial}{\partial t}}+b\hat{\frac{\partial}{\partial z}}\right).

At (0,t)(0,t) we also have

|2(ct2z¯^+att^+bz^)|2=4|a|2+2|b|2t2+2t2|c|2.\left|2\left(ct^{2}\hat{\frac{\partial}{\partial{\overline{z}}}}+at\hat{\frac{\partial}{\partial t}}+b\hat{\frac{\partial}{\partial z}}\right)\right|^{2}=4|a|^{2}+\frac{2|b|^{2}}{t^{2}}+2t^{2}|c|^{2}.

To calculate TsTs we note that

Ts=[s,z^]dz+[s,z¯^]dz¯+[s,t^]dt.Ts=\left[s,\hat{\frac{\partial}{\partial z}}\right]dz+\left[s,\hat{\frac{\partial}{\partial{\overline{z}}}}\right]d{\overline{z}}+\left[s,\hat{\frac{\partial}{\partial t}}\right]dt.
Lemma 2.3

Let 𝔭{\mathfrak{p}} be a parabolic vector field on a neighborhood of a point pp in a hyperbolic manifold MM. Let enTpMe_{n}\in T_{p}M be a unit vector orthogonal to the horosphere tangent to 𝔭{\mathfrak{p}}, pointing away from the fixed point of pp and ωn\omega_{n} the dual \mathbb{R}-valued 1-form. Then

T𝔭=e^nω+𝔭ωnT{\mathfrak{p}}=\hat{e}_{n}\otimes\omega+{\mathfrak{p}}\otimes\omega_{n}

where ω\omega is a \mathbb{C}-linear 1-form with ω(en)=0\omega(e_{n})=0. Furthermore |ω|=|𝔭||\omega|=|{\mathfrak{p}}|.

Proof: We can assume that 𝔭=λz2z{\mathfrak{p}}=\lambda z^{2}\frac{\partial}{\partial z} and p=(0,1)p=(0,1) in the upper half space model. Then en=t^e_{n}=\hat{\frac{\partial}{\partial t}} and ωn=dt\omega_{n}=dt. To calculate T𝔭T{\mathfrak{p}} we use Lemma 2.2 to write 𝔭,z^,z¯^{\mathfrak{p}},\hat{\frac{\partial}{\partial z}},\hat{\frac{\partial}{\partial{\overline{z}}}} and t^\hat{\frac{\partial}{\partial t}} as matrices and calculate the Lie brackets using matrix multiplication and get

T𝔭=λ2t^dz+𝔭dtT{\mathfrak{p}}=\frac{\lambda}{2}\hat{\frac{\partial}{\partial t}}\otimes dz+{\mathfrak{p}}\otimes dt

so ω=λdz\omega=\lambda dz. We can then compute to see that |ω|=|𝔭||\omega|=|{\mathfrak{p}}|. \Box

Complex projective structures

We will be interested in complex projective structures Σ\Sigma that have a fixed underlying conformal structure XX. The space P(X)P(X) of such projective structures has a natural affine structure as the space Q(X)Q(X) of holomorphic quadratic differentials on XX. That is the difference of Σ0\Sigma_{0} and Σ1\Sigma_{1} in is quadratic differential in Q(X)Q(X) defined as follows. Let (U,ψ0)(U,\psi_{0}) and (U,ψ1)(U,\psi_{1}) be charts for Σ0\Sigma_{0} and Σ1\Sigma_{1}. Then ψ1ψ01\psi_{1}\circ\psi_{0}^{-1} is a conformal map for an open neighborhood in ^\widehat{{\mathbb{C}}} to ^\widehat{{\mathbb{C}}}. The Schwarzian derivative is a holomorphic function ϕ(Σ0,Σ1)\phi(\Sigma_{0},\Sigma_{1}) on ψ0(U)\psi_{0}(U) and it determines a holomorphic quadratic differential Φ(Σ0,Σ1)\Phi(\Sigma_{0},\Sigma_{1}). Properties of the Schwarzian derivative imply that if Σ2P(X)\Sigma_{2}\in P(X) is a third projective structure then

Φ(Σ0,Σ2)=Φ(Σ0,Σ1)+Φ(Σ1,Σ2).\Phi(\Sigma_{0},\Sigma_{2})=\Phi(\Sigma_{0},\Sigma_{1})+\Phi(\Sigma_{1},\Sigma_{2}).

This gives a canonical identification of the tangent space TΣP(X)T_{\Sigma}P(X) with Q(X)Q(X).

If g^\hat{g} is a conformal metric on XX and ΦQ(X)\Phi\in Q(X) then the ratio |Φ|/g^|\Phi|/\hat{g} is a positive function on SS. More concretely in a local chart Φ\Phi can be written as ϕdz2\phi dz^{2} and the conformal metric can be written as g^=ρgeuc\hat{g}=\rho{g_{\rm euc}}, where ρ\rho is a positive function and geuc{g_{\rm euc}} is the Euclidean metric. Then the ratio |ϕ|/ρ|\phi|/\rho, defined in the chart, is a well defined function on the surface. We denote this function by Φ(z)g^\|\Phi(z)\|_{\hat{g}} and let Φg^,p\|\Phi\|_{\hat{g},p} be the LpL^{p}-norm of this function with respect to the g^\hat{g} metric. We will mostly be interested in the hyperbolic metric but much of what we do will work in a more general setting. When we are using the hyperbolic metric we will drop the metric g^\hat{g} from our notation.

For every conformal structure XX there is unique Fuchsian projective structure ΣFP(X)\Sigma_{F}\in P(X). We let Σg^,=Φ(Σ,ΣF)g^,\|\Sigma\|_{\hat{g},\infty}=\|\Phi(\Sigma,\Sigma_{F})\|_{\hat{g},\infty}.

If Σt\Sigma_{t} is a smooth path in P(X)P(X) then its tangent vectors Φt\Phi_{t} lie in Q(X)=TΣtP(X)Q(X)=T_{\Sigma_{t}}P(X). To prove our main result, Theorem 1.3, we bound the distances between the endpoints of a path Σt\Sigma_{t} by bounding the norms of the derivative Φt\Phi_{t}.

We also describe how a quadratic differential ΦQ(X)=TΣP(X)\Phi\in Q(X)=T_{\Sigma}P(X) determines a cohomology class in H1(Σ;E(Σ))H^{1}(\Sigma;E(\Sigma)). This was originally introduced by the second author in [Brm].

Define a section 𝔭{\mathfrak{p}} of E()E(\mathbb{C}) by

𝔭(z)=(wz)2w=12[zz21z].{\mathfrak{p}}(z)=(w-z)^{2}\frac{\partial}{\partial w}=\frac{1}{2}\begin{bmatrix}-z&z^{2}\\ -1&z\end{bmatrix}.

Then if Φ\Phi is represented in a projective chart by ϕdz2\phi dz^{2} we defined an EE-valued 1-form in the chart by

𝔭(z)ϕ(z)dz.{\mathfrak{p}}(z)\phi(z)dz.

One can then check that this gives a well defined EE-valued 1-form ωΦ\omega_{\Phi} on Σ\Sigma. As both 𝔭{\mathfrak{p}} and ϕ\phi are holomorphic, ωΦ\omega_{\Phi} is closed and therefore determines a cohomology class in H1(Σ;E(Σ))H^{1}(\Sigma;E(\Sigma)).

Deformations of (𝖯𝖲𝖫2(),¯3)\left(\mathsf{PSL}_{2}(\mathbb{C}),{\bar{\mathbb{H}}^{3}}\right)-structures

Let vv be a vector field on an open neighborhood UU in ¯3{\bar{\mathbb{H}}^{3}} that is conformal on U^U\cap\widehat{{\mathbb{C}}}. We then define a section ss of E(U)=U×𝔰𝔩2()E(U)=U\times\mathfrak{sl}_{2}(\mathbb{C}) as follows. For xU3x\in U\cap{{\mathbb{H}}^{3}} let s(x)s(x) be the unique infinitesimal isometry that agrees with vv at xx and whose curl agrees with the curl of vv at xx. On U^U\cap\widehat{{\mathbb{C}}}, the vector field is (the real part) of the product of a holomorphic function ff and z\frac{\partial}{\partial z}. For each zU^z\in U\cap\widehat{{\mathbb{C}}} let fzf_{z} be the complex quadratic polynomial whose 2-jet agrees with ff at zz. Then s(z)=fzzs(z)=f_{z}\frac{\partial}{\partial z}.

Let {(Uα,ψtα)}\left\{\left(U^{\alpha},\psi^{\alpha}_{t}\right)\right\} be a 1-parameter family of (𝖯𝖲𝖫2(),¯3)\left(\mathsf{PSL}_{2}(\mathbb{C}),{\bar{\mathbb{H}}^{3}}\right)-structures on a 3-manifold with boundary M¯\bar{M} with the conformal boundary fixed. For each xUαx\in U^{\alpha}, the time zero derivative of the path ψtα(x)\psi^{\alpha}_{t}(x) is vector field vαv^{\alpha} on ψ0α(Uα)\psi^{\alpha}_{0}(U^{\alpha}) that is conformal on ψ0α(Uα)^\psi^{\alpha}_{0}(U^{\alpha})\cap\widehat{{\mathbb{C}}}. This determines a section sαs^{\alpha} of E¯(ψ0α(Uα))\bar{E}(\psi^{\alpha}_{0}(U^{\alpha})) and ωα=dsα\omega^{\alpha}=ds^{\alpha} is an EE-valued 1-form on ψ0α(Uα)\psi^{\alpha}_{0}(U^{\alpha}). While the sections sαs^{\alpha} will not necessarily agree on overlapping charts, the EE-valued 1-forms ωα\omega^{\alpha} will agree and determine an EE-valued 1-form ω\omega on M¯\bar{M}. As locally ω\omega is dd of a section, ω\omega is closed and therefore represents an element of H1(M¯;E¯)H^{1}(\bar{M};\bar{E}).

Since the conformal boundary is some fixed conformal structure XX, the (𝖯𝖲𝖫2(),¯3)\left(\mathsf{PSL}_{2}(\mathbb{C}),{\bar{\mathbb{H}}^{3}}\right)-structures on M¯\bar{M} determine a family of projective structures ΣtP(X)\Sigma_{t}\in P(X). The time zero derivative of Σt\Sigma_{t} will be a holomorphic quadratic differential ΦQ(X)=TΣ0P(X)\Phi\in Q(X)=T_{\Sigma_{0}}P(X).

Proposition 2.4 ([Bro, Theorem 2.3])

The restriction of ω\omega to the projective boundary is ωΦ\omega_{\Phi}.

2.2 Hyperbolic metrics on ends

Let M¯\bar{M} have a (𝖯𝖲𝖫2(),¯3)\left(\mathsf{PSL}_{2}(\mathbb{C}),{\bar{\mathbb{H}}^{3}}\right)-structure. A convex surface SS in M=intM¯M=\operatorname{int}\bar{M} cuts off a conformally compact end {\mathcal{E}} if M¯S\bar{M}-S has two components and the outward component is homeomorphic to S×[0,1)S\times[0,1) with S×{0}M¯S\times\{0\}\subset\partial\bar{M}. Then {\mathcal{E}} is the metric closure of the restriction of the outward component to MM and it is homeomorphic to S×(0,1]S\times(0,1] with S=S×{1}S=S\times\{1\} the original convex surface. We also let ¯\bar{\mathcal{E}} be the union of {\mathcal{E}} with the projective boundary Σ\Sigma so that ¯=S×[0,1]\bar{\mathcal{E}}=S\times[0,1].

The unit tangent vectors to the geodesic rays in {\mathcal{E}} orthogonal to SS define a vector field on {\mathcal{E}}. We can choose this product structure such that (z,s)(z,s) is the time ss flow of this vector field. If SsS_{s} is the the time ss image of SS under this normal flow, then the hyperbolic metric for {\mathcal{E}} is can be written as a product of the induced metrics gsg_{s} on SsS_{s} and ds2ds^{2}. However, it will be convenient to parameterize the surfaces in the parameter t=est=e^{-s} rather than in the time parameter ss and we will see there is a nice formula for the hyperbolic metric in this product structure. The result is essentially due to Epstein. However, as it is not given in the exact form we need we derive it here.

Theorem 2.5 (C. Epstein, [Eps])

Let SS be a convex surface cutting off a conformally compact end {\mathcal{E}} with conformal boundary XX. Then there exists a conformal metric g^\hat{g} on XX and a bundle endomorphism B^\hat{B} of TXTX such that the hyperbolic metric on =S×(0,1]{\mathcal{E}}=S\times(0,1] is given by

gt×dt2/t2g_{t}\times dt^{2}/t^{2}

where

gt=14t2(Id+t2B^)g^g_{t}=\frac{1}{4t^{2}}\left({\operatorname{Id}}+t^{2}\hat{B}\right)^{*}\hat{g}

Proof: Let gg be the induced metric on SS and BB the shape operator and let StS_{t} be the distance logt-\log t normal flow of SS in {\mathcal{E}}. We also let

At=(1+t2)Id+(1t2)B=(Id+B)+t2(IdB).A_{t}=(1+t^{2})\cdot{\operatorname{Id}}+(1-t^{2})\cdot B=({\operatorname{Id}}+B)+t^{2}\cdot({\operatorname{Id}}-B).

Then the induced metric gtg_{t} on StS_{t} is given by

gt=14t2Atgg_{t}=\frac{1}{4t^{2}}A^{*}_{t}g

(see [KS, Lemma 2.2]). To get our representation of the hyperbolic metric in {\mathcal{E}} we need to rewrite gtg_{t} in terms of a conformal metric on XX.

The conformal structure on the boundary is induced from the conformal structure on {\mathcal{E}}. If we multiply our metric n {\mathcal{E}} by 4t24t^{2} the conformal structure doesn’t change but the new metric will extend continuously to

g^=(A0)g\hat{g}=(A_{0})^{*}g

on S×{0}S\times\{0\} so g^\hat{g} is a conformal metric on XX and A0=Id+BA_{0}={\operatorname{Id}}+B. As SS is convex, the eigenvalues of BB are non-negative. Therefore we define B^=(Id+B)1(IdB)\hat{B}=({\operatorname{Id}}+B)^{-1}({\operatorname{Id}}-B). It follows that

gt=14t2Atg=14t2(A01At)g^=14t2(Id+t2B^)g^g_{t}=\frac{1}{4t^{2}}A_{t}^{*}g=\frac{1}{4t^{2}}\left(A_{0}^{-1}A_{t}\right)^{*}\hat{g}=\frac{1}{4t^{2}}\left({\operatorname{Id}}+t^{2}\hat{B}\right)^{*}\hat{g}

as claimed. \Box

In Theorem 2.5 the conformal metric on the boundary is determined by the convex surfaces in the hyperbolic end. In [Eps], Epstein has a construction that starts with a metric at infinity and produces the convex surfaces. We only will use his construction for the hyperbolic metric. In particular, we have:

Theorem 2.6 ([Bro, Propositions 6.4 and 6.5])

Let XX be the a component of the conformal boundary of a conformally compact hyperbolic cone manifold and g^x\hat{g}_{x} is the hyperbolic metric on XX. Then for all t>log(1+2Σg^X,)t>\log\left(\sqrt{1+2\|\Sigma\|_{\hat{g}_{X},\infty}}\right) there is a convex surface SS that cuts of an end {\mathcal{E}} such that e2tg^Xe^{2t}\hat{g}_{X} is the metric at infinity for SS.

The metric in a chart

If (U,ψ)(U,\psi) is a projective chart for Σ\Sigma then we can extend it to a chart (U×[0,1],Ψ)(U\times[0,1],\Psi) where Ψ\Psi is the continuous extension of ψ\psi to a map to ¯3{\bar{\mathbb{H}}^{3}} that is an isometry on U×(0,1]U\times(0,1]. We say that the chart is adapted to z0Uz_{0}\in U if Ψ(z0,t)=(0,t)\Psi(z_{0},t)=(0,t) where the coordinates on the right are in the upper half space model for 3{{\mathbb{H}}^{3}}. We can always construct a chart adapted to z0z_{0} by taking any projective chart (U,ψ)(U,\psi) with z0Uz_{0}\in U and post-composing with an element of 𝖯𝖲𝖫2()\mathsf{PSL}_{2}(\mathbb{C}).

In a projective chart the metric at infinity g^\hat{g} is scalar function times the Euclidean metric geuc{g_{\rm euc}}. If the chart is adapted to z0z_{0} then we can calculate the value of this function.

Lemma 2.7

If (U×[0,1],Ψ)(U\times[0,1],\Psi) is a chart adapted to z0Σz_{0}\in\Sigma then on ψ(U)\psi(U) the metric g^\hat{g} is of the form ρgeuc\rho{g_{\rm euc}} where ρ:ψ(U)+\rho\colon\psi(U)\to\mathbb{R}^{+} is smooth and ρ(0)=4\rho(0)=4.

Proof: Define a function ρ:Ψ(U×[1,0))+\rho\colon\Psi(U\times[1,0))\to\mathbb{R}^{+} with ρΨ(z,t)=4t2\rho\circ\Psi(z,t)=4t^{2}. If g3g_{{\mathbb{H}}^{3}} is the metric for the upper half space model of 3{{\mathbb{H}}^{3}} then ρg3\rho\cdot g_{{\mathbb{H}}^{3}} extends continuously to the metric g^\hat{g} on ψ(U)\psi(U). Since Ψ(z0,t)=(0,t)\Psi(z_{0},t)=(0,t) we have that ρ(0,t)=4t2\rho(0,t)=4t^{2} and therefore ρg3\rho g_{{{\mathbb{H}}^{3}}} extends continuously to 4geuc4{g_{\rm euc}} at 0. \Box

On a chart (U,ψ)(U,\psi) for Σ\Sigma we have the usual coordinate vector fields x{\frac{\partial}{\partial x}} and y{\frac{\partial}{\partial y}} along with the vector fields in z\frac{\partial}{\partial z} and z¯\frac{\partial}{\partial{\overline{z}}} in the complexified tangent bundle. On a chart (U×[0,1],Ψ)(U\times[0,1],\Psi) for the end {\mathcal{E}} these coordinate vector fields, along with t\frac{\partial}{\partial t} are a basis but, unlike in the upper half space model for 3{{\mathbb{H}}^{3}}, the vector fields x{\frac{\partial}{\partial x}} and y{\frac{\partial}{\partial y}} may not be orthogonal or of the same length as the operators (Id+t2B^)({\operatorname{Id}}+t^{2}\hat{B}) are not conformal. In particular, the complex 1-form dzdz will not be \mathbb{C}-linear on the complex structure on SS induced by the metric gtg_{t}. However, we can write down \mathbb{C}-linear and \mathbb{C}-anti-linear forms in terms of the Beltrami differential of the endomorphisms that define gtg_{t}.

We begin with a computation on a single vector space. The usual Euclidean metric geuc{g_{\rm euc}} on 2\mathbb{R}^{2} has a unique \mathbb{C}-linear extension to 2\mathbb{R}^{2}\otimes\mathbb{C}. Then dzdz and dz¯d{\overline{z}} are the usual dual basis for 2\mathbb{R}^{2}\otimes\mathbb{C}. While they are both \mathbb{C}-linear on 2\mathbb{R}^{2}\otimes\mathbb{C}, when restricted to 2\mathbb{R}^{2}, with the complex structure induced by geuc{g_{\rm euc}}, dzdz is \mathbb{C}-linear while dz¯d{\overline{z}} is \mathbb{C}-anti-linear. A linear isomorphism A:22A\colon\mathbb{R}^{2}\to\mathbb{R}^{2} has a unique \mathbb{C}-linear extension to 2\mathbb{R}^{2}\otimes\mathbb{C}. The Beltrami differential for AA is

μ=Az¯/Az\mu=A_{\overline{z}}/A_{z}

where AzA_{z} and Az¯A_{\overline{z}} are complex numbers with

A(dz)=Azdz+Az¯dz¯.A^{*}(dz)=A_{z}dz+A_{\overline{z}}d{\overline{z}}.

We have the following:

Lemma 2.8

Let A:22A\colon\mathbb{R}^{2}\to\mathbb{R}^{2} be a linear isomorphism and let g=Ageucg=A^{*}{g_{\rm euc}} and μ\mu the Beltrami differential for AA. Then

[dzdz¯]=11|μ|2[1μμ¯1][dwdw¯]\begin{bmatrix}dz\\ d{\overline{z}}\end{bmatrix}=\frac{1}{1-|\mu|^{2}}\begin{bmatrix}1&-\mu\\ -\bar{\mu}&1\end{bmatrix}\begin{bmatrix}dw\\ d\bar{w}\end{bmatrix}

where dwdw and dw¯d\bar{w} are \mathbb{C}-linear and \mathbb{C}-anti-linear on 2\mathbb{R}^{2} with respect to gg. If g\star_{g} is the Hodge star operator for gg then

g[dzdz¯]=i1|μ|2[1μμ¯1][dwdw¯]=i1|μ|2[1+|μ|22μ2μ¯1|μ|2][dzdz¯].\star_{g}\begin{bmatrix}dz\\ d\bar{z}\end{bmatrix}=\frac{-i}{1-|\mu|^{2}}\begin{bmatrix}1&\mu\\ -\bar{\mu}&-1\end{bmatrix}\begin{bmatrix}dw\\ d\bar{w}\end{bmatrix}=\frac{-i}{1-|\mu|^{2}}\begin{bmatrix}1+|\mu|^{2}&2\mu\\ -2\bar{\mu}&-1-|\mu|^{2}\end{bmatrix}\begin{bmatrix}dz\\ d\bar{z}\end{bmatrix}.

Furthermore

|dw|g=|dz|geuc|Az|.|dw|_{g}=\frac{|dz|_{g_{euc}}}{|A_{z}|}.

Proof: As dzdz and dz¯d{\overline{z}} are \mathbb{C}-linear and \mathbb{C}-anti-linear for the complex structure on 2\mathbb{R}^{2} induced by geuc{g_{\rm euc}}, A(dz)A^{*}(dz) and A(dz¯)A^{*}(d{\overline{z}}) are \mathbb{C}-linear and \mathbb{C}-anti-linear for the complex structure induced by gg. As AA is \mathbb{C}-linear on 2\mathbb{R}^{2}\otimes\mathbb{C} and

A(dz)=Azdz+Az¯dz¯A^{*}(dz)=A_{z}dz+A_{{\overline{z}}}d{\overline{z}}

for complex numbers AzA_{z} and Az¯A_{\overline{z}} we have

A(dz¯)=Az¯¯dz+Az¯dz¯.A^{*}(d{\overline{z}})=\bar{A_{{\overline{z}}}}dz+\bar{A_{z}}d{\overline{z}}.

Dividing A(dz)A^{*}(dz) by AzA_{z} and A(dz¯)A^{*}(d{\overline{z}}) by Az¯\bar{A_{z}} we define dwdw and dw¯d\bar{w} by

[dwdw¯]=[1μμ¯1][dzdz¯]\begin{bmatrix}dw\\ d\bar{w}\end{bmatrix}=\begin{bmatrix}1&\mu\\ \bar{\mu}&1\end{bmatrix}\begin{bmatrix}dz\\ d{\overline{z}}\end{bmatrix}

so that dwdw is \mathbb{C}-linear and dw¯d\bar{w} is \mathbb{C}-anti-linear on the complex structure induced by gg. Inverting gives our formula for dzdz and dz¯d{\overline{z}} in terms of dwdw and dw¯d\bar{w}.

As dwdw and dw¯d\bar{w} are \mathbb{C}-linear and \mathbb{C}-anti-linear with respect to gg we have

gdw=idw,gdw¯=idw¯\star_{g}dw=-idw,\qquad\star_{g}d\overline{w}=id\overline{w}

or

g[dwdw¯]=[i00i][dwdw¯].\star_{g}\begin{bmatrix}dw\\ d\bar{w}\end{bmatrix}=\begin{bmatrix}-i&0\\ 0&i\end{bmatrix}\begin{bmatrix}dw\\ d\bar{w}\end{bmatrix}.

We then have

g[dzdz¯]=(11|μ|2[1μμ¯1])[i00i][1μμ¯1][dzdz¯].\star_{g}\begin{bmatrix}dz\\ d\bar{z}\end{bmatrix}=\left(\frac{1}{1-|\mu|^{2}}\begin{bmatrix}1&-\mu\\ -\bar{\mu}&1\end{bmatrix}\right)\begin{bmatrix}-i&0\\ 0&i\end{bmatrix}\begin{bmatrix}1&\mu\\ \bar{\mu}&1\end{bmatrix}\begin{bmatrix}dz\\ d\bar{z}\end{bmatrix}.

Multiplying, we obtain the stated formulas. For the norm |dw|g|dw|_{g} we note that define dW=Adz=AzdwdW=A^{*}dz=A_{z}dw. Then

|Azdw|g=|dW|g=|dz|geuc.|A_{z}dw|_{g}=|dW|_{g}=|dz|_{g_{euc}}.

\Box

We can apply the above to the metrics gtg_{t}. The Beltrami differential for endomorphisms 12t(Id+t2B^)\frac{1}{2t}({\operatorname{Id}}+t^{2}\hat{B}) can be written as t2μtt^{2}\mu_{t} where

μt=B^z¯1+t2B^z.\mu_{t}=\frac{\hat{B}_{\overline{z}}}{1+t^{2}\hat{B}_{z}}.

We obtain the following immediate corollary.

Corollary 2.9

Let (U,ψ)(U,\psi) be a projective chart for Σ=\Sigma=\partial{\mathcal{E}} with corresponding chart (U×[0,1],Ψ)(U\times[0,1],\Psi) for {\mathcal{E}}. Then

dz=idzdtt2it2(t2β0dz+β1dz¯)dtt.\star dz=-idz\wedge\frac{dt}{t}-2it^{2}\left(t^{2}\beta_{0}dz+\beta_{1}d{\overline{z}}\right)\wedge\frac{dt}{t}.

where the βi\beta_{i} are the smooth functions on U×[0,1]U\times[0,1] given by

β0(z,t)=|μt|21t4|μt|2β1(z,t)=μt1t4|μt|2.\beta_{0}(z,t)=\frac{|\mu_{t}|^{2}}{1-t^{4}|\mu_{t}|^{2}}\qquad\beta_{1}(z,t)=\frac{\mu_{t}}{1-t^{4}|\mu_{t}|^{2}}.

Further for dwt=dz+t2μtdz¯dw_{t}=dz+t^{2}\mu_{t}d{\overline{z}},

dz=idwtdttit2(t2β0dwt+β1dw¯t)dtt\star dz=-idw_{t}\wedge\frac{dt}{t}-it^{2}\left(t^{2}\beta_{0}dw_{t}+\beta_{1}d\bar{w}_{t}\right)\wedge\frac{dt}{t}

and

|dwt|gt=2t|dz|g^|1+t2Bz|.|dw_{t}|_{g_{t}}=\frac{2t|dz|_{\hat{g}}}{|1+t^{2}B_{z}|}.

2.3 Model deformations

If SS is a convex surface cutting of a conformally compact end {\mathcal{E}} with projective boundary Σ\Sigma we can use Theorem 2.5 to extend ωΦ\omega_{\Phi} to {\mathcal{E}}. Let

Π:S×(0,1]Σ=S×{0}\Pi\colon S\times(0,1]\to\Sigma=S\times\{0\}

be given by Π(z,t)=z\Pi(z,t)=z. We would like to extend Π\Pi to a bundle map between E()E({\mathcal{E}}) and E(Σ)E(\Sigma). For this we note that for any flat bundle a path between two points in base determines an isomorphism between their fibers as the a flat bundle restricted to a path has a canonical product structure.

In our case the geodesic rays {z}×[0,1]\{z\}\times[0,1] are paths in ¯\bar{{\mathcal{E}}} between (z,t)(z,t) and (z,0)(z,0) and determine isomorphisms between the fiber E(z,t)E_{(z,t)} of E()E({\mathcal{E}}) over (z,t)(z,t) and the fiber EzE_{z} of E(Σ)E(\Sigma) over zz. Using this isomorphism we can extend Π\Pi to a bundle map

Π:E()E(Σ).\Pi_{*}\colon E({\mathcal{E}})\to E(\Sigma).

We then extend ωΦ\omega_{\Phi} to a 1-form in Ω1(,E())\Omega^{1}({\mathcal{E}},E({\mathcal{E}})) by pulling back ωΦ\omega_{\Phi} via Π\Pi_{*}.

Lemma 2.10
ωΦ(ωΦ)=t216Φg^2(1+2t4|μt|21t4|μt|2)dAg^dt/t\omega_{\Phi}\wedge\star\left(\omega_{\Phi}^{\sharp}\right)=\frac{t^{2}}{16}\|\Phi\|^{2}_{\hat{g}}\left(1+\frac{2t^{4}|\mu_{t}|^{2}}{1-t^{4}|\mu_{t}|^{2}}\right)dA_{\hat{g}}\wedge dt/t

Proof: We calculate at a point (z,t)S×(0,1](z,t)\in S\times(0,1] by taking a chart (U,ψ)(U,\psi) adapted to zz. In this chart ωΦ\omega_{\Phi} is written as

ϕ(z)𝔭(z)dz.\phi(z){\mathfrak{p}}(z)dz.

While this expression does not depend on tt, the Hodge star operator and the dual map will. In particular the expression

(ϕ(z)𝔭(z)dz)=ϕ¯(z)𝔭(z)dz¯\star({\phi}(z){\mathfrak{p}}(z)dz)^{\sharp}=\bar{\phi}(z){\mathfrak{p}}(z)^{\sharp}\star d{\overline{z}}

depends on tt as both pp^{\sharp} and dz¯\star d{\overline{z}} depend on tt.

By taking the conjugate of dz\star dz in Corollary 2.9 we have

dzdz¯=i(1+2t4|μt|21t4|μt|2)dzdz¯dt/t.dz\wedge\star d{\overline{z}}=i\left(1+\frac{2t^{4}|\mu_{t}|^{2}}{1-t^{4}|\mu_{t}|^{2}}\right)dz\wedge d{\overline{z}}\wedge dt/t.

We also need to find 𝔭(z)(𝔭(z))=|𝔭|2(z){\mathfrak{p}}(z)^{\sharp}({\mathfrak{p}}(z))=|{\mathfrak{p}}|^{2}(z). As we are working in a chart adapted to zz, we have Ψ(z,t)=(0,t)\Psi(z,t)=(0,t). Since 𝔭(0)=w2w{\mathfrak{p}}(0)=w^{2}\frac{\partial}{\partial w}, by Lemma 2.3 at (0,t)(0,t) we have

|𝔭(0)|2=|t2z¯^|2=t2/2.|{\mathfrak{p}}(0)|^{2}=\left|-t^{2}\hat{\frac{\partial}{\partial{\overline{z}}}}\right|^{2}=t^{2}/2.

By Lemma 2.7 we have Φ(z)g^2=4|ϕ(0)|2\|\Phi(z)\|^{2}_{\hat{g}}=4|\phi(0)|^{2} and dAg^=4dxdy=2idzdz¯dA_{\hat{g}}=4dx\wedge dy=2idz\wedge d{\overline{z}} and combining our calculations we have the result. \Box

Let t=S×(0,t]{\mathcal{E}}_{t}=S\times(0,t] be portion of the end cutoff by StS_{t} and let (,)t(,)_{t} be inner product on t{\mathcal{E}}_{t}. For an EE-valued form ω\omega on {\mathcal{E}} we then define ωt2=(ω,ω)t\|\omega\|^{2}_{t}=(\omega,\omega)_{t}. Integrating the prior lemma we immediately get:

Corollary 2.11

We have ωΦt28t2Φg^,22\|\omega_{\Phi}\|^{2}_{t}\geq 8t^{2}\|\Phi\|^{2}_{\hat{g},2} and

limt01t2ωΦt2=8Φg^,22.\underset{t\to 0}{\lim}\frac{1}{t^{2}}\|\omega_{\Phi}\|^{2}_{t}=8\|\Phi\|^{2}_{{\hat{g}},2}.

The form ωΦ\omega_{\Phi} is not harmonic as δωΦ0\delta\omega_{\Phi}\neq 0. However, we will show that δωΦ2\|\delta\omega_{\Phi}\|_{2} decays rapidly in tt.

We’ll break the estimate into small calculations.

Lemma 2.12

Let UU be a neighborhood of 00\in\mathbb{C} and let ss be a smooth section of E(U×[0,1])E(U\times[0,1]) such that the function |s||s| on U×(0,1]3U\times(0,1]\subset{{\mathbb{H}}^{3}} extends continuously to U×[0,1]¯3U\times[0,1]\subset{\bar{\mathbb{H}}^{3}}. Then the projective vector field s(0,0)s(0,0) has a zero at 0U0\in U. If |s|(0,0)=0|s|(0,0)=0 then s(0,0)s(0,0) is a is a multiple of 𝔭(0){\mathfrak{p}}(0) and

|s|(0,t)=O(t).|s|(0,t)=O(t).

Proof: We write

s(z,t)=(f0(z,t)+f1(z,t)w+f2(z,t)w2)ws(z,t)=\left(f_{0}(z,t)+f_{1}(z,t)w+f_{2}(z,t)w^{2}\right)\frac{\partial}{\partial w}

where the functions fif_{i} are smooth, complexed valued functions on U×[0,1]U\times[0,1]. By Lemma 2.2 we have

|s|2(0,t)=|f0(0,t)|22t2+|f1(0,t)|24+t2|f2(0,t)|22.|s|^{2}(0,t)=\frac{|f_{0}(0,t)|^{2}}{2t^{2}}+\frac{|f_{1}(0,t)|^{2}}{4}+\frac{t^{2}|f_{2}(0,t)|^{2}}{2}.

for t(0,1)t\in(0,1). If |s||s| extends continuously to (0,0)(0,0) then we must have f0(0,0)=0f_{0}(0,0)=0 so, as a projective vector field, s(0,0)s(0,0) is zero at 0.

If |s|(0,0)=0|s|(0,0)=0 the we must further have that f1(0,0)=0f_{1}(0,0)=0 and f0t(0,0)=0\frac{\partial f_{0}}{\partial t}(0,0)=0. Since f1t(0,0)\frac{\partial f_{1}}{\partial t}(0,0) also exists it follows that

|fi(0,t)|=O(t2i)|f_{i}(0,t)|=O\left(t^{2-i}\right)

and therefore

|s(0,t)|=O(t).|s(0,t)|=O(t).

\Box

Lemma 2.13

Let UU be a neighborhood of 00\in\mathbb{C} and let ss be a smooth section of E(U×[0,1])E(U\times[0,1]) such that the function |s||s| on U×(0,1]3U\times(0,1]\subset{{\mathbb{H}}^{3}} extends continuously to U×[0,1]¯3U\times[0,1]\subset{\bar{\mathbb{H}}^{3}}. Then

|dsdz¯dt|(0,t)=O(t3)and|dsdzdt|(0,t)=O(t4).|ds\wedge d{\overline{z}}\wedge dt|(0,t)=O\left(t^{3}\right)\quad\mbox{and}\quad|ds\wedge dz\wedge dt|(0,t)=O\left(t^{4}\right).

Proof: By Lemma 2.12 the condition that the norm |s||s| extends to zero on U×{0}U\times\{0\} implies that s(z,0)=f(z)𝔭(z)s(z,0)=f(z){\mathfrak{p}}(z) for some smooth complex valued function ff on UU. We have that

dsdz¯dt=szdzdz¯dt=2it3szdVds\wedge d{\overline{z}}\wedge dt=s_{z}dz\wedge d{\overline{z}}\wedge dt=-2it^{3}s_{z}dV

where szs_{z} is zz-derivative of ss and dV=idzdz¯dt2t3dV=\frac{idz\wedge d{\overline{z}}\wedge dt}{2t^{3}} is the volume form for 3{{\mathbb{H}}^{3}}. Therefore

|dsdz¯dt|(0,t)=2t3|sz|.|ds\wedge d{\overline{z}}\wedge dt|(0,t)=2t^{3}|s_{z}|.

Note that the zz-derivative of the section 𝔭{\mathfrak{p}} is

𝔭z(z)=2(wz)w{\mathfrak{p}}_{z}(z)=-2(w-z)\frac{\partial}{\partial w}

and has bounded norm on 3{{\mathbb{H}}^{3}}. Therefore, on {0}×(0,1]\{0\}\times(0,1], the norm of sz=fz𝔭+f𝔭zs_{z}=f_{z}{\mathfrak{p}}+f{\mathfrak{p}}_{z} is bounded and if we let cc be bound of 2|sz|2|s_{z}| we have

|dsdz¯dt|(0,t)ct3.|ds\wedge d{\overline{z}}\wedge dt|(0,t)\leq ct^{3}.

The bound of the norm of dsdzdtds\wedge dz\wedge dt is similar once we note that the z¯{\overline{z}}-derivative of 𝔭{\mathfrak{p}} is zero so the norm of the z¯{\overline{z}}-derivative sz¯s_{\overline{z}} of ss is also zero at (0.0)(0.0). \Box

We now prove our bounds on δωΦ\delta\omega_{\Phi}.

Lemma 2.14
limt01t2δωΦt=0.\lim_{t\to 0}\frac{1}{t^{2}}\|\delta\omega_{\Phi}\|_{t}=0.

Proof: We have δ=\delta=\star\partial\star with =DT=d2T\partial=D-T=d-2T. Therefore to bound δωΦt\|\delta\omega_{\Phi}\|_{t} we need to bound the norm of d(ωΦ)d(\star\omega_{\Phi}) and T(ωΦ)T(\star\omega_{\Phi}).

In a chart (U×[0,1],Ψ)(U\times[0,1],\Psi) by Corollary 2.9 we have

dz=idzdtt2it2(t2β0dz+β1dz¯)dtt\star dz=-idz\wedge\frac{dt}{t}-2it^{2}(t^{2}\beta_{0}dz+\beta_{1}d{\overline{z}})\wedge\frac{dt}{t}

where the βi\beta_{i} are smooth, complex valued functions on U×[0,1]U\times[0,1]. Therefore

ωΦ=ϕ𝔭dz=iϕ𝔭dzdtt2it2ϕ𝔭(t2β0dz+β1dz¯)dtt\star\omega_{\Phi}=\phi{\mathfrak{p}}\star dz=-i\phi{\mathfrak{p}}dz\wedge\frac{dt}{t}-2it^{2}\phi{\mathfrak{p}}(t^{2}\beta_{0}dz+\beta_{1}d{\overline{z}})\wedge\frac{dt}{t}

As ϕ\phi and 𝔭{\mathfrak{p}} are holomorphic in the dzdz-coordinate we have that d(iϕ𝔭dzdt/t)=0d(-i\phi{\mathfrak{p}}dz\wedge dt/t)=0. Therefore

d(ωΦ)=2it(t2d(ϕβ0𝔭)dz+d(ϕβ1𝔭)dz¯)dtd(\star\omega_{\Phi})=-2it(t^{2}d(\phi\beta_{0}{\mathfrak{p}})\wedge dz+d(\phi\beta_{1}{\mathfrak{p}})\wedge d{\overline{z}})\wedge dt

and, as the sections ϕβi𝔭\phi\beta_{i}{\mathfrak{p}} have norm limiting to zero on \partial{\mathcal{E}}, by Lemma 2.13

|d(ωΦ)|(0,t)=O(t4).|d(\star\omega_{\Phi})|(0,t)=O\left(t^{4}\right).

Next we calculate T(ωΦ)T(\star\omega_{\Phi}). We will work in a chart adapted to zz and and a conformal coordinate wtw_{t} at (0,t)(0,t). Again applying Corollary 2.9 we have

ωΦ=ϕ𝔭dz=iϕ𝔭((1+t4β0)dwt+t2β1dw¯t)dtt\star\omega_{\Phi}=\phi{\mathfrak{p}}\star dz=-i\phi{\mathfrak{p}}\left((1+t^{4}\beta_{0})dw_{t}+t^{2}\beta_{1}d\bar{w}_{t}\right)\wedge\frac{dt}{t}

where the βi\beta_{i} are as above. We use Lemma 2.3 to calculate T𝔭T{\mathfrak{p}}. At the point (z,t)(z,t) we have en=1tte_{n}=\frac{1}{t}\frac{\partial}{\partial t}, ωn=dtt\omega_{n}=\frac{dt}{t} and ω=λdwt\omega=\lambda dw_{t} for some scalar λ\lambda. Then

T𝔭=λtt^dwt+𝔭dttT{\mathfrak{p}}=\lambda t\hat{\frac{\partial}{\partial t}}\otimes dw_{t}+{\mathfrak{p}}\otimes\frac{dt}{t}

where |λ||dwt|=|𝔭|=t/2|\lambda||dw_{t}|=|{\mathfrak{p}}|=t/\sqrt{2}. Then

T(ωΦ)=iϕλt2β1t^dwtdw¯tdt=iϕt4β1|dwt|2t^dVT(\star\omega_{\Phi})=-i\phi\lambda t^{2}\beta_{1}\hat{\frac{\partial}{\partial t}}\otimes dw_{t}\wedge d\bar{w}_{t}\wedge dt=\frac{-i\phi t^{4}\beta_{1}|dw_{t}|}{\sqrt{2}}\hat{\frac{\partial}{\partial t}}dV

where dVdV is the volume form and dwtdw¯tdt=it|dwt|2dVdw_{t}\wedge d\bar{w}_{t}\wedge dt=it|dw_{t}|^{2}dV. Then

|T(ωΦ)|=t3|ϕ||β1||dwt|/2|T(\star\omega_{\Phi})|=t^{3}|\phi||\beta_{1}||dw_{t}|/\sqrt{2}

since |tt|=1\left|t\frac{\partial}{\partial t}\right|=1. By Corollary 2.9 |dwt|=O(t)|dw_{t}|=O(t), giving

|T(ωΦ)|=O(t4).|T(\star\omega_{\Phi})|=O\left(t^{4}\right).

As the \star-operator is an isometry and SS is compact this implies that there exist a c>0c>0 such that

|δωΦ|(z,t)ct4.|\delta\omega_{\Phi}|(z,t)\leq ct^{4}.

We then have

t0δωΦδωΦ\displaystyle\int_{{\mathcal{E}}_{t_{0}}}\delta\omega_{\Phi}\wedge\star\delta\omega_{\Phi}^{\sharp} =\displaystyle= t0|δωΦ|2𝑑V\displaystyle\int_{{\mathcal{E}}_{t_{0}}}|\delta\omega_{\Phi}|^{2}dV
\displaystyle\leq 0t0Sct8𝑑At𝑑t/t\displaystyle\int_{0}^{t_{0}}\int_{S}ct^{8}dA_{t}dt/t
\displaystyle\leq 0t0Kct5𝑑t=Kct06/6.\displaystyle\int_{0}^{t_{0}}Kct^{5}dt=Kc{t_{0}}^{6}/6.

Here dAtdA_{t} is the area form for the surface StS_{t} and we are using the fact that the area of these surfaces is bounded by K/t2K/t^{2} for some K>0K>0. Therefore

δωΦt2Kct6/6\|\delta\omega_{\Phi}\|_{t}^{2}\leq Kct^{6}/6

and the lemma follows. \Box

We now prove the main result of this section.

Theorem 2.15

Let ω=ωΦ+dτ\omega=\omega_{\Phi}+d\tau where the section τ\tau of E()E({\mathcal{E}}) has finite L2L^{2}-norm. The for all t01t_{0}\leq 1

Φg^,2218t02ωt02||\Phi||^{2}_{\hat{g},2}\leq\frac{1}{8t_{0}^{2}}||\omega||_{t_{0}}^{2}

Proof: We have

(ω,ω)t=(ωΦ+dτ,ωΦ+dτ)t=(ωΦ,ωΦ)t+2Re(dτ,ωΦ)t+(dτ,dτ)t.(\omega,\omega)_{t}=(\omega_{\Phi}+d\tau,\omega_{\Phi}+d\tau)_{t}=(\omega_{\Phi},\omega_{\Phi})_{t}+2\operatorname{Re}(d\tau,\omega_{\Phi})_{t}+(d\tau,d\tau)_{t}.

By Corollary 2.11 we have

Φg^,2218t2(ωΦ,ωΦ)t.\|\Phi\|_{\hat{g},2}^{2}\leq\frac{1}{8t^{2}}(\omega_{\Phi},\omega_{\Phi})_{t}.

For the middle term we have integrate ωΦdτ\omega_{\Phi}\wedge\star d\tau^{\sharp} over the compact manifold S×[s,t]S\times[s,t] and let s0s\to 0. We have

S×[s,t]dτωΦ\displaystyle\int_{S\times[s,t]}d\tau\wedge\star\omega_{\Phi}^{\sharp} =\displaystyle= S×[s,t]τδωΦ+StτωΦSsτωΦ\displaystyle\int_{S\times[s,t]}\tau\wedge\delta\omega_{\Phi}+\int_{S_{t}}\tau\wedge\star\omega_{\Phi}^{\sharp}-\int_{S_{s}}\tau\wedge\star\omega_{\Phi}^{\sharp}
=\displaystyle= S×[s,t]τδωΦ(τ,δωϕ)t\displaystyle\int_{S\times[s,t]}\tau\wedge\delta\omega_{\Phi}\to(\tau,\delta\omega_{\phi})_{t}

where the integrals over StS_{t} and SsS_{s} are both zero since ωΦ\star\omega_{\Phi} restricted to these surfaces is zero as it contains a dtdt-term.

As τ<\|\tau\|<\infty, applying Lemma 2.14 we get

lim supt01t2|(dτ,ωΦ)t|=lim supt01t2|(τ,δωΦ)t|lim supt01t2τtδωΦt=0.\limsup_{t\rightarrow 0}\frac{1}{t^{2}}|(d\tau,\omega_{\Phi})_{t}|=\limsup_{t\rightarrow 0}\frac{1}{t^{2}}|(\tau,\delta\omega_{\Phi})_{t}|\leq\limsup_{t\rightarrow 0}\frac{1}{t^{2}}\|\tau\|_{t}\cdot\|\delta\omega_{\Phi}\|_{t}=0.

By the infinitesimal inflexibility theorem [BB, Theorem 3.6] we have for t<t0t<t_{0}

1t2(ω,ω)t1t02(ω,ω)t0.\frac{1}{t^{2}}(\omega,\omega)_{t}\leq\frac{1}{t_{0}^{2}}(\omega,\omega)_{t_{0}}.

Therefore as (dτ,dτ)t0(d\tau,d\tau)_{t}\geq 0,

Φg^,22lim inft018t2(ωΦ,ωΦ)tlim inft018t2(ω,ω)t18t02(ω,ω)t0\|\Phi\|_{\hat{g},2}^{2}\leq\liminf_{t\rightarrow 0}\frac{1}{8t^{2}}(\omega_{\Phi},\omega_{\Phi})_{t}\leq\liminf_{t\rightarrow 0}\frac{1}{8t^{2}}(\omega,\omega)_{t}\leq\frac{1}{8t_{0}^{2}}(\omega,\omega)_{t_{0}}

\Box

If ω\omega is a Hodge form on a conformally compact hyperbolic cone-manifold that is cohomologous to some ωΦ\omega_{\Phi} on and end {\mathcal{E}} then, by definition, ω=ωΦ+dτ\omega=\omega_{\Phi}+d\tau for some EE-valued section τ\tau on {\mathcal{E}}. To apply this theorem we need the extra property that τ\tau has finite L2L^{2}-norm. We call such a Hodge form a model Hodge form.

3 Nehari type bounds for cone-manifolds

For a smooth, hyperbolic 3-manifold with incompressible boundary the classical Nehari bound on the Schwarzian derivative of univalent maps gives that Σ3/2\|\Sigma\|_{\infty}\leq 3/2 for every component Σ\Sigma of the projective boundary. We are interested in obtaining similar bounds for a hyperbolic cone-manifolds. To do so we need to make some technical assumptions, that will always be satisfied in our applications, but do make the statement somewhat cumbersome.

One of the difficulties is that the usual Margulis lemma does not hold for cone-manifolds. The following statement is a replacement.

Theorem 3.1 ([Bro, Theorem 3.5])

There exists an L0>0L_{0}>0 such that the following holds. Let MM be a hyperbolic cone-manifold such that all cone angle 2π\leq 2\pi, every component of the cone cone locus has length L0\leq L_{0} and every component has a tubular neighborhood of radius sinh12\sinh^{-1}\sqrt{2}. Further assume that these neighborhood are mutually disjoint. Then each component cc of the cone locus of length LcL_{c} and cone angle θc\theta_{c} has a tubular neighborhood of radius RcR_{c} where

θcLcsinh(2Rc)=1.\theta_{c}L_{c}\sinh(2R_{c})=1.

Now we state our version of the Nehari bound. When the cone angle is small it will be important that the cone locus has a large tubular neighborhood where the radius grows as the cone angle decreases. The necessary lower bounds will come from the previous result and to use it we will need to assume that the length of the cone locus is bounded above by a linear function of the cone angle.

Theorem 3.2

There exists an L0>0L_{0}>0 such that the following holds. Let MM be a conformally compact hyperbolic cone-manifold such that all cone angle 2π\leq 2\pi and there are a disjoint collection of tubular neighborhoods of the components of the cone locus of radius sinh12\geq\sinh^{-1}\sqrt{2}. Further assume that if cc is a component of the cone locus with cone angle θc\theta_{c} and length LcL_{c} then

LcθcL0.L_{c}\leq\theta_{c}L_{0}.

Then for every component Σ\Sigma of the projective boundary of MM we have

ΣK.\|\Sigma\|_{\infty}\leq K.

In order to prove this, we will need to consider the Thurston parametrization of projective structures via measured laminations and use the notion of average bending of a measured lamination. We show that the result follows from a compactness argument.

3.1 The Thurston parameterization

The space P(Δ)P(\Delta) of projective structures on the hyperbolic disk is equivalent to the space of locally univalent maps f:Δ^f\colon\Delta\to\widehat{{\mathbb{C}}} with the equivalence fgf\sim g if f=ϕgf=\phi\circ g for some ϕ𝖯𝖲𝖫2()\phi\in\mathsf{PSL}_{2}(\mathbb{C}). We can identify P(Δ)P(\Delta) with the space of quadratic differentials Q(Δ)Q(\Delta) by mapping [f]P(Δ)[f]\in P(\Delta) to its Schwarzian derivative S(f)Q(Δ)S(f)\in Q(\Delta). Then the topology on P(Δ)P(\Delta) is the compact-open topology on Q(Δ)Q(\Delta).

Thurston described a natural parameterization of P(Δ)P(\Delta) by (2){\cal M}{\cal L}({\mathbb{H}}^{2}) the space of measure geodesic laminations on 2{\mathbb{H}}^{2}. We briefly review this construction.

A round disk D^D\subset\widehat{{\mathbb{C}}} shares a boundary with a hyperbolic plane D23{\mathbb{H}}^{2}_{D}\subseteq{\mathbb{H}}^{3}. Let rD:D3r_{D}\colon D\to{\mathbb{H}}^{3} be the nearest point projection to D2{\mathbb{H}}^{2}_{D} and r~D:DT13\tilde{r}_{D}\colon D\to T^{1}{\mathbb{H}}^{3} be the normal vector to D2{\mathbb{H}}^{2}_{D} at rD(z)r_{D}(z) pointing towards DD. We can use these maps to define a version of the Epstein map for ρf\rho_{f}. In particular define Ep~ρf:ΔT13\widetilde{\operatorname{Ep}}_{\rho_{f}}\colon\Delta\to T^{1}{\mathbb{H}}^{3} by Ep~ρf(z)=r~f(D)(f(z))\widetilde{\operatorname{Ep}}_{\rho_{f}}(z)=\tilde{r}_{f(D)}(f(z)) where DD is the unique round disk with respect to ff such that ρD(z)=ρf(z)\rho_{D}(z)=\rho_{f}(z) and let Epρf(z)=πEp~ρf(z)=rf(D)(f(z))\operatorname{Ep}_{\rho_{f}}(z)=\pi\circ\widetilde{\operatorname{Ep}}_{\rho_{f}}(z)=r_{f(D)}(f(z)). (For the existence of this disk see [KT, Theorem 1.2.7].) We also define Ep~etρf=gtEp~ρf\widetilde{\operatorname{Ep}}_{e^{t}\rho_{f}}=g_{t}\circ\widetilde{\operatorname{Ep}}_{\rho_{f}} and Epetρf=πEp~etρf\operatorname{Ep}_{e^{t}\rho_{f}}=\pi\circ\widetilde{\operatorname{Ep}}_{e^{t}\rho_{f}}.

The image of Epρf\operatorname{Ep}_{\rho_{f}} is a locally convex pleated plane. More precisely, let (2){\cal M}{\cal L}({\mathbb{H}}^{2}) be measured geodesic laminations on 2{\mathbb{H}}^{2} and 0(2)(2){\cal M}{\cal L}_{0}({\mathbb{H}}^{2})\subseteq{\cal M}{\cal L}({\mathbb{H}}^{2}) the subspace of laminations with finite support. That is λ0(2)\lambda\in{\cal M}{\cal L}_{0}({\mathbb{H}}^{2}) if it is the union of a finite collection of disjoint geodesics i\ell_{i} with positive weights θi\theta_{i}. Then λ\lambda determines a continuous map pλ:23p_{\lambda}\colon{\mathbb{H}}^{2}\to{\mathbb{H}}^{3}, unique up to post-composition with isometries of 3{\mathbb{H}}^{3}, that is an isometry on the complement of the support of λ\lambda and is “bent” with angle θi\theta_{i} at i\ell_{i}. By continuity we can extend this construction to a general λ(2)\lambda\in{\cal M}{\cal L}({\mathbb{H}}^{2}). An exposition of the following theorem of Thurston can be found in [KT].

Theorem 3.3

Given fP(Δ)f\in P(\Delta) there exists maps cf:Δ2c_{f}\colon\Delta\to{\mathbb{H}}^{2} and pf:23p_{f}\colon{\mathbb{H}}^{2}\to{\mathbb{H}}^{3} and a lamination λf\lambda_{f} such that pfp_{f} is a locally, convex pleated surface pleated along λf\lambda_{f}, Epρf=pfcf\operatorname{Ep}_{\rho_{f}}=p_{f}\circ c_{f} and the map fλff\mapsto\lambda_{f} is a homeomorphism from P(Δ)(2)P(\Delta)\to{\cal M}{\cal L}({\mathbb{H}}^{2}). Furthermore the maps cf:(Δ,ρf)2c_{f}\colon(\Delta,\rho_{f})\to{\mathbb{H}}^{2} and Epρf:(Δ,ρf)3\operatorname{Ep}_{\rho_{f}}\colon(\Delta,\rho_{f})\to{\mathbb{H}}^{3} are 1-Lipschitz.

3.2 Average Bending Bound

Average bending was introduced by the first author in the study of convex hulls of quasifuchsian groups (see [Bri] and [BC1]). This had applications in the work of Epstein, Marden and Markovic in their paper [EMM]. The idea of average bending is to relate the injectivity radius of the convex hull to the amount of bending per unit length along geodesic arcs. In their work, Epstein, Marden and Markovic, used an equivalent formulation of average bending, called roundedness.

Given λ(Δ)\lambda\in\mathcal{ML}(\Delta) and α\alpha a transverse arc, we let λ(α)\lambda(\alpha) be the λ\lambda-measure of α\alpha. We then define the average bending norm to be

λL=sup{λ(α)|α an open geodesic arc of length L}.||\lambda||_{L}=\sup\{\lambda(\alpha)\ |\alpha\mbox{ an open geodesic arc of length }L\}.

If λ\lambda is a lift of a measured lamination on a closed hyperbolic surface, then λL||\lambda||_{L} is bounded but in general μL||\mu||_{L} may be infinite. For simplicity, we will let μ1=μ\|\mu\|_{1}=\|\mu\|.

We have the following compactness result;

Lemma 3.4

Given L,M>0L,M>0 then the set C(L,M)={λ|λLM}C(L,M)=\{\lambda\ |\ ||\lambda||_{L}\leq M\} is precompact.

Proof: Let G(Δ)G(\Delta) be the space of (unoriented) geodesics in the hyperbolic plane. We define the space of geodesic currents 𝒞(Δ){\mathcal{C}}(\Delta) to be the space of non-negative Borel measures on G(Δ)G(\Delta) with the weak topology. The topology on (Δ)\mathcal{ML}(\Delta) is that of a closed subspace of 𝒞(Δ){\mathcal{C}}(\Delta). Given an open geodesic arc α\alpha, we let UαG(Δ)U_{\alpha}\subseteq G(\Delta) be the set of all geodesics transverse to α\alpha. We define

𝒰={Uα|α an open geodesic arc of length L}.\mathcal{U}=\{U_{\alpha}\ |\ \alpha\mbox{ an open geodesic arc of length }L\}.

Then 𝒰\mathcal{U} is an open cover of G(Δ)G(\Delta).

We let 𝒦\mathcal{K} be the set of continuous functions on G(Δ)G(\Delta) with support subordinate to the cover 𝒰\mathcal{U}. Then for each ϕ𝒦\phi\in\mathcal{K} there exists a U𝒰U\in\mathcal{U} with supp(ϕ)Usupp(\phi)\subset U. We have the map I:(Δ)|𝒦|I:\mathcal{ML}(\Delta)\rightarrow\mathbb{R}^{|\mathcal{K}|} given by I(λ)=(λ(ϕ))ϕ𝒦I(\lambda)=(\lambda(\phi))_{\phi\in\mathcal{K}}. This map is a homeomorphism onto its image.

If ϕ𝒦\phi\in\mathcal{K} then there is a U𝒰U\in\mathcal{U} with supp(ϕ)U𝒰supp(\phi)\subset U\in\mathcal{U}. Therefore for λC(M,L)\lambda\in C(M,L)

λ(ϕ)λ(U)M.\lambda(\phi)\leq\lambda(U)\leq M.

Therefore C(M,L)C(M,L) is homeomorphic to a subset of [0,M]|𝒦|[0,M]^{|\mathcal{K}|} which is compact by Tychanoff’s theorem. Therefore C(M,L)C(M,L) is precompact.\Box

Corollary 3.5

Given L,M>0L,M>0 there exists an RR such that if ff is a locally univalent map with λfLM\|\lambda_{f}\|_{L}\leq M then

ϕf<R\|\phi_{f}\|_{\infty}<R

.

Proof: We consider the family F(M,L)F(M,L) of ϕf=S(f)Q(Δ)\phi_{f}=S(f)\in Q(\Delta) with λfC(M,L)\lambda_{f}\in C(M,L). Then by Thurston, F(M,L)F(M,L) is the image of C(M,L)C(M,L) under a homeomorphism. Therefore F(M,L)F(M,L) is precompact and has compact closure K(M,L)K(M,L). Therefore there is an R>0R>0 such that for all λfC(M,L)\lambda_{f}\in C(M,L) then

|ϕf(0)|R/4.|\phi_{f}(0)|\leq R/4.

Therefore ϕf(0)R||\phi_{f}(0)||\leq R for all λfC(M,L)\lambda_{f}\in C(M,L). As the set K(M,L)K(M,L) is invariant under isometries of 2{\mathbb{H}}^{2} it follows that ϕfR||\phi_{f}||_{\infty}\leq R for all λfC(M,L)\lambda_{f}\in C(M,L). \Box

3.3 Convex Hull of Cone Manifold

In this section MM will be a conformally compact hyperbolic cone-manifold with incompressible boundary and all cone angles 2π\leq 2\pi. We let ϕ\phi be the quadratic differential on the conformal boundary given by uniformization. In [Bro], the second author studied the convex core boundary of MM. This is given by taking the Epstein surface for the projective metric which we denote by SS. By [Bro, Proposition 6.5] the surface SS is an embedded locally convex surface in MM bounding an end \mathcal{E} of MM homeomorphic to S×[0,)S\times[0,\infty). Also \mathcal{E} does not contain any cone axes in its interior. The surface SS has intrinsic hyperbolic metric and has a bending lamination βϕ\beta_{\phi}. We identify the universal cover S~\tilde{S} with the hyperbolic disk Δ\Delta and obtain a lamination β~ϕ\tilde{\beta}_{\phi}.

First some elementary lemmas about balls in hyperbolic cone-manifolds.

Lemma 3.6

Let SS be the unit sphere in 3\mathbb{R}^{3}. Let (θ,z)(\theta,z) be cylindrical coordinates on SS and for 0<t2π0<t\leq 2\pi define the spherical cone-surface

St={(θ,z)S| 0θt}/(0,z)(t,z).S_{t}=\{(\theta,z)\in S\ |\ 0\leq\theta\leq t\}/(0,z)\sim(t,z).

If p1,p2,p3Stp_{1},p_{2},p_{3}\in S_{t} then d(pi,pj)2π/3d(p_{i},p_{j})\leq 2\pi/3 for some i,j,iji,j,i\neq j.

Proof: Assume not. We first take the case of t=2πt=2\pi. Then St=SS_{t}=S the unit sphere. Then letting B(p,r)B(p,r) be an open disk of radius rr about pSp\in S, we have

p2,p3B(p1,2π/3)¯c=B(p1,π/3).p_{2},p_{3}\in\overline{B(p_{1},2\pi/3)}^{c}=B(-p_{1},\pi/3).

It follows that dS(p2,p3)2π/3d_{S}(p_{2},p_{3})\leq 2\pi/3 giving our contradiction.

For t<2πt<2\pi we take a fundamental wedge domain WtW_{t} for StS_{t} in SS above, and can assume the pip_{i} are in the interior. Then by the spherical case two of the points have dS(pi,pj)2π/3d_{S}(p_{i},p_{j})\leq 2\pi/3. As dSt(pi,pj)dS(pi,pj)d_{S_{t}}(p_{i},p_{j})\leq d_{S}(p_{i},p_{j}) we obtain our contradiction. \Box

We have the following elementary calculation on half-spaces in 3{{\mathbb{H}}^{3}};

Lemma 3.7

Let f:++f:\mathbb{R}_{+}\rightarrow\mathbb{R}_{+} be given by

f(R)=cosh1(2cosh(R)1+3cosh2(R)).f(R)=\cosh^{-1}\left(\frac{2\cosh(R)}{\sqrt{1+3\cosh^{2}(R)}}\right).

Let H1,H2,H3H_{1},H_{2},H_{3} be half-spaces in 3{{\mathbb{H}}^{3}} such that HiB(x,R)H_{i}\cap B(x,R) are disjoint. If each HiH_{i} intersects B(x,r)B(x,r) then rf(R)r\geq f(R).

Proof Let rir_{i} be the distance from HiH_{i} to xx and let Di=HiB(x,R)D_{i}=H_{i}\cap\partial B(x,R) have spherical radius θi\theta_{i}. Then we have rirr_{i}\leq r and θiθ\theta_{i}\geq\theta where θ\theta is the spherical radius of D=HB(x,R)D=H\cap\partial B(x,R) where HH is a half-space a distance rr from xx. Therefore as each DiD_{i} contains a disk or radius θ\theta, if the DiD_{i} are disjoint, then there are 3 disks of radius θ\theta which are disjoint.

We show that θπ/3\theta\leq\pi/3. We let S=B(x,R)S=\partial B(x,R) have the spherical metric given by angle subtended at xx. If θ>π/3\theta>\pi/3 then the centers pip_{i} of DiD_{i} satisfy d(pi,pj)>2π/3,ijd(p_{i},p_{j})>2\pi/3,i\neq j contradicting Lemma 3.6.

We have a right-angled hyperbolic triangle with sides r,Rr,R and angle θ\theta between. Let ll be the length of the other side. Then solving we have

sinh(l)=sinh(R).sin(θ)3sinh(R)2.\sinh(l)=\sinh(R).\sin(\theta)\leq\frac{\sqrt{3}\sinh(R)}{2}.

and by the hyperbolic Pythagorean formula

cosh(r)=cosh(R)cosh(l)=cosh(R)1+sinh2(l)cosh(R)1+34sinh2(R)=2cosh(R)1+3cosh2(R).\cosh(r)=\frac{\cosh(R)}{\cosh(l)}=\frac{\cosh(R)}{\sqrt{1+\sinh^{2}(l)}}\geq\frac{\cosh(R)}{\sqrt{1+\frac{3}{4}\sinh^{2}(R)}}=\frac{2\cosh(R)}{\sqrt{1+3\cosh^{2}(R)}}.

\Box

We now consider balls in our cone manifold MM. We let M~\tilde{M} be the universal cover with convex hull C(M~)C(\tilde{M}). The end \mathcal{E} lifts to ~\tilde{\mathcal{E}} a component of the complement of C(M~)C(\tilde{M}) with boundary S~\tilde{S}. As MM has incompressible boundary, then π1(~)\pi_{1}(\tilde{\mathcal{E}}) is trivial.

The space M~\tilde{M} is a hyperbolic cone manifold and the cone axes 𝒞\mathcal{C} lift to 𝒞~\tilde{\mathcal{C}}. For pM~p\in\tilde{M} we define balls in the usual way, i.e. B(p,r)={qM~|d(p,q)r}B(p,r)=\{q\in\tilde{M}\ |\ d(p,q)\leq r\}. We note that B(p,r)B(p,r) may not be topologically a ball or isometric to a hyperbolic ball. For a point pp, we define r(p)r(p) to be the maximum radius such that B(x,r(p))B(x,r(p)) is embedded and isometric to a hyperbolic ball of radius r(p)r(p). Note for p𝒞~p\in\tilde{\mathcal{C}}, r(p)=0r(p)=0 and otherwise r(p)>0r(p)>0 and r(p)r(p) equals is the injectivity radius of pp in M~𝒞~\tilde{M}-\tilde{\mathcal{C}}. For pM~p\in\tilde{M} we further define d(p)d(p) to be the minimum distance to the cone axes 𝒞~\tilde{\mathcal{C}}.

We first bound the average bending for points with r(p)r(p) bounded below.

Lemma 3.8

Let MM be a conformally compact hyperbolic cone-manifold with incompressible boundary and all cone angles 2π\leq 2\pi. Let pS~p\in\tilde{S} and α\alpha a closed geodesic arc on S~\tilde{S} with midpoint pp and length less than 2f(r(p))2f(r(p)). Then

β~ϕ(α)<2π\tilde{\beta}_{\phi}(\alpha)<2\pi

Proof: We let HsH_{s} be the 1-parameter family of support half-spaces from α(0)\alpha(0) to α(1)\alpha(1). We consider S=B(p,r(p))S=\partial B(p,r(p)) and disks Ds=HsSD_{s}=H_{s}\cap S. We let s1s_{1} be the smallest ss such that D0,DsD_{0},D_{s} have disjoint interiors. Then we have β~(α([0,s1]))<π\tilde{\beta}(\alpha([0,s_{1}]))<\pi. If there is no such s1s_{1} then we have β~(α)<π\tilde{\beta}(\alpha)<\pi and we’re done.

We now let s2s_{2} be the smallest tt such that Ds1,DsD_{s_{1}},D_{s} have disjoint interiors. Again it follows that β~(α[s1,s2])<π\tilde{\beta}(\alpha[s_{1},s_{2}])<\pi giving β~(α([0,s2]))<2π\tilde{\beta}(\alpha([0,s_{2}]))<2\pi. If no such t2t_{2} exists then β~(α)<2π\tilde{\beta}(\alpha)<2\pi and we are also done.

We first show that D0,Ds2D_{0},D_{s_{2}} do not intersect. If D0,Ds2D_{0},D_{s_{2}} do intersect, we extend α([0,s2])\alpha([0,s_{2}]) to a closed curve α\alpha^{\prime} by joining α(0),α(s2)\alpha(0),\alpha(s_{2}) by a piecewise geodesics on H0Hs2\partial H_{0}\cup\partial H_{s_{2}}. We note that ~\tilde{\mathcal{E}} is simply connected. We get our contradiction by showing that curve α\alpha^{\prime} in ~\tilde{\mathcal{E}} is homotopically non-trivial. The curve α\alpha^{\prime} is homotopic to a simple closed curve α′′\alpha^{\prime\prime} in D0Ds1Ds2BD_{0}\cup D_{s_{1}}\cup D_{s_{2}}\subset\partial B via a homotopy in B(H0Hs1Hs2)~B\cap(H_{0}\cup H_{s_{1}}\cup H_{s_{2}})\subset\tilde{\mathcal{E}}. But as arc α\alpha is transverse to a bending line bb then α′′\alpha^{\prime\prime} separates the points bBb\cap\partial B in B\partial B. Therefore α′′\alpha^{\prime\prime} is non-trivial in M~b\tilde{M}-b. As MM has incompressible boundary, α′′\alpha^{\prime\prime} is trivial M~C(M~)M~b\tilde{M}-C(\tilde{M})\subset\tilde{M}-b and we obtain our contradiction. Thus if s2s_{2} exists, then D0,Ds2D_{0},D_{s_{2}} do not intersect. We then obtain a contradiction from the above lemma as H0,Hs1,Hs2H_{0},H_{s_{1}},H_{s_{2}} are disjoint in B(p,r(p))B(p,r(p)) and intersect B(p,f(r(p))B(p,f(r(p)). \Box

We use the same argument as above to bound average bending for points close to the cone axes.

Lemma 3.9

Let MM be a conformally compact hyperbolic cone-manifold with incompressible boundary and all cone angles 2π\leq 2\pi. Let c~𝒞~\tilde{c}\in\tilde{\mathcal{C}} have an embedded tube Uc~U_{\tilde{c}} of radius RR and pUc~p\in U_{\tilde{c}} with d(p,c)f(R)/2d(p,c)\leq f(R)/2. If α\alpha is a closed geodesic arc on S~\tilde{S} with midpoint pp and length less than f(R)f(R) then

β~ϕ(α)<2π\tilde{\beta}_{\phi}(\alpha)<2\pi

Proof: We let q𝒞~q\in\tilde{\mathcal{C}} be the nearest point of to pp on 𝒞~\tilde{\mathcal{C}} and consider B=B(q,R0)B=B(q,R_{0}). If α\alpha is a geodesic arc of length f(R0)f(R_{0}) centered about pp, then α\alpha is in B(q,f(R0))B(q,f(R_{0})). We let S=BS=\partial B, then SS is a sphere with two cone points. We again consider HsH_{s} the 1-parameter family of support planes from α(0)\alpha(0) to α(1)\alpha(1) and let Ds=HsSD_{s}=H_{s}\cap S. Then DsD_{s} are disks in SS whose interior are disjoint from the cone points. Then analysing as in Lemma 3.7, we obtain 3 disks with disjoint interiors on SS. By Lemma 3.6 the disks cannot be disjoint which gives a contradiction. Thus we have β~(α)<2π\tilde{\beta}(\alpha)<2\pi. \Box

To bound our average bending uniformly for a given length, reduces now to showing that r(p)r(p) is bounded away from zero for points far from the cone axes. This is the purpose of the following two lemmas.

Lemma 3.10

Let MM be a conformally compact hyperbolic cone-manifold with incompressible boundary and all cone angles 2π\leq 2\pi. Let 𝒞\mathcal{C} be the cone-axes and for c~𝒞~\tilde{c}\in\tilde{\mathcal{C}} let Uc~U_{\tilde{c}} be the RR neighborhood of c~\tilde{c} in M~\tilde{M}. Let RR be such that Uc~U_{\tilde{c}} are embedded and disjoint.

  • If pM~c~𝒞~Uc~p\in\tilde{M}-\cup_{\tilde{c}\in\tilde{\mathcal{C}}}U_{\tilde{c}} then r(p)r(q)r(p)\geq r(q) for some qUc~q\in\partial U_{\tilde{c}} and c~𝒞~\tilde{c}\in\tilde{\mathcal{C}}.

  • if pUc~p\in U_{\tilde{c}} then d(p)=d(p,c~)d(p)=d(p,\tilde{c}) and

    r(p)=\displaystyle r(p)= d(p)\displaystyle d(p) θcπ,\displaystyle\theta_{c}\geq\pi,
    sinh(r(p))=\displaystyle\sinh(r(p))= sinh(d(p))sin(θc/2)\displaystyle\sinh(d(p))\sin(\theta_{c}/2) θcπ.\displaystyle\theta_{c}\leq\pi.

Proof: We M^\hat{M} to be the completion of the universal cover of M𝒞M-\mathcal{C}. Then M^\hat{M} is CAT(0). We let U^c~M^\hat{U}_{\tilde{c}}\subseteq\hat{M} be the completions of the universal covers of Uc~U_{\tilde{c}}.

Let pM~c~𝒞~Uc~p\in\tilde{M}-\cup_{\tilde{c}\in\tilde{\mathcal{C}}}U_{\tilde{c}}. If r(p)r(p) is achieved by an arc joining pp to an axis then r(p)Rr(p)\geq R and as r(q)Rr(q)\leq R for qUc~q\in\partial U_{\tilde{c}} then the first statement follows. Otherwise there is a non-trivial geodesic γ\gamma of length 2r(p)2r(p) in M~𝒞~\tilde{M}-\tilde{\mathcal{C}}. Therefore γ\gamma must link a finite collection of axes of elements 𝒞~\tilde{\mathcal{C}}. If γ\gamma links more than one axis then γ\gamma is greater than the length of the shortest closed geodesic linking the axes. As this is at least 2R2R, then r(p)Rr(p)\geq R as before. Therefore we can assume γ\gamma links a single axis c~𝒞~\tilde{c}\in\tilde{\mathcal{C}}. Then γ\gamma lifts to γ^\hat{\gamma} a piecewise geodesic in M^\hat{M} which is invariant under the action of the deck transformation gc~g_{\tilde{c}} corrseponding to c~\tilde{c}. As M^\hat{M} is CAT(0) and the U^c~\hat{U}_{\tilde{c}} are convex and complete, projection πc~\pi_{\tilde{c}} onto U^c~\hat{U}_{\tilde{c}} is distance decreasing (see [BH, Proposition II.2.4]). Thus as πc~\pi_{\tilde{c}} commutes with the action of gc~g_{\tilde{c}}, the curve πc~(γ^)\pi_{\tilde{c}}(\hat{\gamma}) descends to a curve in M~\tilde{M} of length 2r(p)\leq 2r(p) which is contained in Uc~U_{\tilde{c}} linking c~\tilde{c} with basepoint qUc~q\in\partial U_{\tilde{c}}. Thus r(p)r(q)r(p)\geq r(q) and the first item is done.

For pU~c~p\in\tilde{U}_{\tilde{c}} then trivially d(p)=d(p,c~)d(p)=d(p,\tilde{c}). We now describe the relation between d(p),r(p)d(p),r(p). Then projecting as above, we have that r(p)r(p) is attained by a curve in Uc~U_{\tilde{c}}. We take a fundamental domain for Uc~U_{\tilde{c}} to be a wedge WW of a hyperbolic tube of radius RR about a geodesic with wedge angle tt and have pp be on the central radial line of the wedge. Taking the largest ball about pp that is embedded in WW it follows that if θcπ\theta_{c}\geq\pi then we have that r(p)=d(p)r(p)=d(p). Otherwise θcπ\theta_{c}\leq\pi and then 2r(p)2r(p) is the length of the unique shortest geodesic arc with both endpoints pp. Thus r(p),d(p)r(p),d(p) are sides of a right angled triangle, with hypothenuse d(p)d(p) and side of length r(p)r(p) facing angle θc/2\theta_{c}/2. Therefore for θcπ\theta_{c}\leq\pi by the hyperbolic sine formula

sinh(r(p))=sinh(d(p))sin(θc/2).\sinh(r(p))=\sinh(d(p))\sin(\theta_{c}/2).

\Box

Lemma 3.11

There is an explicit monotonic increasing function g:++g:\mathbb{R}_{+}\rightarrow\mathbb{R}_{+} such that the following holds. Let MM be a conformally compact hyperbolic cone-manifold with incompressible boundary and all cone angles 2π\leq 2\pi satisfying the conditions of Theorem 3.2. Then for r12sinh1(2)r\leq\frac{1}{2}\sinh^{-1}(\sqrt{2}), if pC(M~)p\in\partial C(\tilde{M}) with d(p)rd(p)\geq r then r(p)g(r)r(p)\geq g(r).

Proof: By assumption for c𝒞c\in\mathcal{C}, cc has embedded tubular neighborhood UcU_{c} of radius RcR_{c} such that Rcsinh1(2)R_{c}\geq\sinh^{-1}(\sqrt{2}). We let R0=sinh1(2)R_{0}=\sinh^{-1}(\sqrt{2}). We lift the tubular neighborhoods to M~\tilde{M} and denote by Uc~U_{\tilde{c}} the lift for c~𝒞~\tilde{c}\in\tilde{\mathcal{C}}.

By the Lemma 3.10, we need only consider points in the neighborhoods Uc~U_{\tilde{c}}. Thus we let pUc~p\in U_{\tilde{c}}. Again by Lemma 3.10 if θcπ\theta_{c}\geq\pi then r(p)d(p)r(p)\geq d(p) giving r(p)rr(p)\geq r. Similarly for π/2θcπ\pi/2\leq\theta_{c}\leq\pi, we have

sinh(r(p))=sin(θc/2)sinh(d(p))12sinh(r).\sinh(r(p))=\sin(\theta_{c}/2)\sinh(d(p))\geq\frac{1}{\sqrt{2}}\sinh(r).

This gives a bound on r(p)r(p) for θcπ/2\theta_{c}\geq\pi/2.

We now consider θcπ/2\theta_{c}\leq\pi/2. By [Bro, Lemma 3.3] all support half-spaces are embedded in \mathcal{E}. Let HH be a half space intersecting Uc~U_{\tilde{c}} with distance dd from the cone axis. We take a wedge fundamental domain with the nearest point of HH being centered. Then in order for HH to be embedded in Uc~U_{\tilde{c}}, it cannot intersect the radial sides of the wedge. Therefore we must have d>dcd>d_{c} where dc,Rcd_{c},R_{c} form a right-angled triangle with hypothenuse RcR_{c} and angle between the sides θc/2\theta_{c}/2. Labeling the other side of the triangle ll we have by hyperbolic geometry (see [FN, formulas III.5, III.6])

sinh(l)=sinh(Rc)sin(θc/2)tanh(l)=sinh(dc)tan(θc/2).\sinh(l)=\sinh(R_{c})\sin(\theta_{c}/2)\qquad\tanh(l)=\sinh(d_{c})\tan(\theta_{c}/2).

Thus for pUc~p\in U_{\tilde{c}} we have d(p)dcd(p)\geq d_{c}. Therefore substituting

sinh(r(p))sinh(dc)sin(θc/2)=tanh(l)cos(θc/2)12sinh(Rc)sin(θc/2)1+sinh2(Rc)sin2(θc/2)\sinh(r(p))\geq\sinh(d_{c})\sin(\theta_{c}/2)=\tanh(l)\cos(\theta_{c}/2)\geq\frac{1}{\sqrt{2}}\frac{\sinh(R_{c})\sin(\theta_{c}/2)}{\sqrt{1+\sinh^{2}(R_{c})\sin^{2}(\theta_{c}/2)}}

To obtain a bound, we use our assumptions in Theorem 3.2 and applying Theorem 3.1 we have

θcLcsinh(2Rc)=1andLcL0θc.\theta_{c}L_{c}\sinh(2R_{c})=1\qquad\mbox{and}\qquad L_{c}\leq L_{0}\theta_{c}.

Therefore

θc1sinh(2Rc)L0.\theta_{c}\geq\frac{1}{\sqrt{\sinh(2R_{c})L_{0}}}.

It follows that

sinh(Rc)sin(θc/2)1L0sinh(Rc)sinh(2Rc)sin(θc/2)θc1L0tanh(Rc)22πtanh(R0)π2L0.\sinh(R_{c})\sin(\theta_{c}/2)\geq\frac{1}{\sqrt{L_{0}}}\frac{\sinh(R_{c})}{\sqrt{\sinh(2R_{c})}}\frac{\sin(\theta_{c}/2)}{\theta_{c}}\geq\frac{1}{\sqrt{L_{0}}}\sqrt{\frac{\tanh(R_{c})}{2}}\frac{\sqrt{2}}{\pi}\geq\sqrt{\frac{\tanh(R_{0})}{\pi^{2}L_{0}}}.

As tanh(R0)=2/3\tanh(R_{0})=\sqrt{2/3} and x/1+x2x/\sqrt{1+x^{2}} is monotonic, then

sinh(r(p))12tanh(R0)π2L01+tanh(R0)π2L0=12+2π2cotanh(R0)L012+24L0.\sinh(r(p))\geq\frac{1}{\sqrt{2}}\frac{\sqrt{\frac{\tanh(R_{0})}{\pi^{2}L_{0}}}}{\sqrt{1+\frac{\tanh(R_{0})}{\pi^{2}L_{0}}}}=\frac{1}{\sqrt{2+2\pi^{2}\operatorname{cotanh}(R_{0})L_{0}}}\geq\frac{1}{\sqrt{2+24L_{0}}}.

Thus for pUc~p\in U_{\tilde{c}} we have r(p)g(r)r(p)\geq g(r) with

sinhg(r)=min(12+24L0,12sinh(r))\sinh g(r)=\min\left(\frac{1}{\sqrt{2+24L_{0}}},\frac{1}{\sqrt{2}}\sinh(r)\right)

Thus combining the bounds, we have r(p)g(r)r(p)\geq g(r) with

sinhg(r)=min(12+24L0,12sinh(r))\sinh g(r)=\min\left(\frac{1}{\sqrt{2+24L_{0}}},\frac{1}{\sqrt{2}}\sinh(r)\right)

\Box

In [BC1], the first author and Canary proved the following.

Theorem 3.12 ([BC1, Theorem 3])

Let f:23f:{\mathbb{H}}^{2}\rightarrow{{\mathbb{H}}^{3}} be an embedded convex pleated plane then its bending lamination βf\beta_{f} satisfies

βfL<2π\|\beta_{f}\|_{L}<2\pi

for L2sinh1(1).L\leq 2\sinh^{-1}(1).

We now use the Lemmas 3.9 and 3.11 above to generalize Theorem 3.12 for cone-deformations.

Proposition 3.13

Let MM be a conformally compact hyperbolic cone-manifold with incompressible boundary and all cone angles 2π\leq 2\pi satisfying the conditions of Theorem 3.2. Then

βϕL<2π||\beta_{\phi}||_{L}<2\pi

for any L2f(g(f(sinh1(2))/2))=.1529L\leq 2f(g(f(\sinh^{-1}(\sqrt{2}))/2))=.1529.

Proof: We let r=f(R0)/2r=f(R_{0})/2 where R0=sinh1(2)R_{0}=\sinh^{-1}(\sqrt{2}).

If d(p)rd(p)\geq r. Then as f(x)xf(x)\leq x we have

r=f(R0)2R02=12sinh1(2).r=\frac{f(R_{0})}{2}\leq\frac{R_{0}}{2}=\frac{1}{2}\sinh^{-1}(\sqrt{2}).

Therefore we can apply Lemma 3.11 to pp to get r(p)g(r)r(p)\geq g(r). Therefore for L=2f(g(r))L=2f(g(r)) then β~(α)2π\tilde{\beta}(\alpha)\leq 2\pi for any geodesic arc α\alpha of length less than LL centered at pp.

If d(p)r=f(R0)/2d(p)\leq r=f(R_{0})/2, as c𝒞c\in\mathcal{C} has an embedded tubes of radius Rc>R0R_{c}>R_{0} then by Lemma 3.9 if α\alpha is an arc of length Lf(R0)L\leq f(R_{0}) then β~(α)2π\tilde{\beta}(\alpha)\leq 2\pi.

Combining the bounds we have

βL<2π\|\beta\|_{L}<2\pi

For Lmin(2r,2f(g(r)))L\leq\min(2r,2f(g(r))). As R0=sinh1(2)R_{0}=\sinh^{-1}(\sqrt{2}) and we can assume L0<1L_{0}<1, then

min(2r,2f(g(r)))=2f(g(r))=.152958.\min(2r,2f(g(r)))=2f(g(r))=.152958.

\Box

We now prove the main result of this section.

Proof of Theorem 3.2: By the above, there exists an LL such that

βϕL<2π.\|\beta_{\phi}\|_{L}<2\pi.

Therefore by Corollary 3.5

ΣK\|\Sigma\|_{\infty}\leq K

for some KK universal. \Box

4 Proof of Theorems 1.2 and 1.3

We now bring our work together to prove the main results of the paper. Before doing so we will need to summarize the necessary results about deformations of cone-manifolds. As in the introduction we have a compact 3-manifold N¯\bar{N} with a collection 𝒞\mathcal{C} of disjoint, simple closed curves in the interior. We will examine a family of conformally compact hyperbolic cone-manifold structures on N¯\bar{N} with cone locus 𝒞\mathcal{C}.

Theorem 4.1 ([Bro])

Let MtM_{t} be a one parameter family of cone-manifolds given by Theorem 1.1 and let Lc(t)L_{c}(t) be the length of a component cc of 𝒞\mathcal{C} in MtM_{t} and L𝒞L_{\mathcal{C}} the sum of the LcL_{c}.

  • Lc(t)tLc(2π)πL_{c}(t)\leq\frac{tL_{c}(2\pi)}{\pi}
  • In each MtM_{t} there is a union UtU_{t} of embedded, disjoint tubular neighborhoods of the components of 𝒞\mathcal{C} of radius sinh12\geq\sinh^{-1}\sqrt{2}.

  • The time tt derivative of MtM_{t} is represented by a model Hodge form ωt\omega_{t} with

    Mt\Utωt2314L𝒞(t)t3L𝒞(2π)14π.\int_{M_{t}\backslash U_{t}}||\omega_{t}||^{2}\leq\frac{3}{14}\cdot\frac{L_{\mathcal{C}}(t)}{t}\leq\frac{3L_{\mathcal{C}}(2\pi)}{14\pi}.

Note that the statement in the final bullet is not the actual statement of Proposition 4.2 in [Bro] but rather a direct application of the first inequality of the proof where we assume that the radius of the tubular neighborhoods is sinh12\sinh^{-1}\sqrt{2} rather than the larger radii assumed in that proposition.

We are now ready to prove our main theorem bounding the L2L^{2}-norm of the derivative of the path of complex projective structures.

Proof of Theorem 1.2: We assume tt has been fixed throughout the proof.

For the path Σt\Sigma_{t} of complex projective structures on the boundary of MtM_{t}, by Theorem 3.2 we have that ΣtK\|\Sigma_{t}\|_{\infty}\leq K. Therefore by Theorem 2.6 there is a convex surface SS in MtM_{t} cutting of an end {\mathcal{E}} such that (1+2K)g^X(1+2K)\hat{g}_{X} is the metric at infinity for SS. Note that while {\mathcal{E}} will be disjoint from the cone locus in MtM_{t} it may intersect the tubular neighborhood UtU_{t} of the cone locus. To correct this we need to remove the collar of width sinh12\sinh^{-1}\sqrt{2} from {\mathcal{E}}. This is the end η{\mathcal{E}}_{\eta} where η=esinh12\eta=e^{-\sinh^{-1}\sqrt{2}}.

By Theorem 4.1 we have that

Mt\Utωt23L𝒞14π\int_{M_{t}\backslash U_{t}}||\omega_{t}||^{2}\leq\frac{3L_{\mathcal{C}}}{14\pi}

and since ηMt\Ut{\mathcal{E}}_{\eta}\subset M_{t}\backslash U_{t} this implies that

ηωt23L𝒞14π.\int_{{\mathcal{E}}_{\eta}}\|\omega_{t}\|^{2}\leq\frac{3L_{\mathcal{C}}}{14\pi}.

As ωt\omega_{t} is a model Hodge form Theorem 4.1 implies that

Φt(1+2K)g^X,2218η23L𝒞14π.\|\Phi_{t}\|^{2}_{(1+2K)\hat{g}_{X},2}\leq\frac{1}{8\eta^{2}}\frac{3L_{\mathcal{C}}}{14\pi}.

As

Φ^tg^X,22=(1+2K)Φt(1+2K)g^X,22\|\hat{\Phi}_{t}\|^{2}_{\hat{g}_{X},2}=(1+2K)\|\Phi_{t}\|^{2}_{(1+2K)\hat{g}_{X},2}

this gives

Φtg^X,2cdrillL𝒞\|\Phi_{t}\|_{\hat{g}_{X},2}\leq c_{\rm drill}\sqrt{L_{\mathcal{C}}}

where

cdrill=14η3(1+2K)7π.c_{\rm drill}=\frac{1}{4\eta}\sqrt{\frac{3(1+2K)}{7\pi}}.

\Box

Our main results now follows immediately.

Proof of Theorem 1.3: Integrating the above, we get the L2L^{2}-bound

Φ(Σ0,Σ2π)202πΦt2𝑑t2πcdrillL𝒞.||\Phi(\Sigma_{0},\Sigma_{2\pi})||_{2}\leq\int_{0}^{2\pi}||\Phi_{t}||_{2}dt\leq 2\pi c_{\rm drill}\sqrt{L_{\mathcal{C}}}.

\Box

References

  • [Bri] M. Bridgeman. Average bending of convex pleated planes in hyperbolic three-space. Invent. Math. 132(1998), 381–391.
  • [BBB] M. Bridgeman, J. Brock, and K. Bromberg. The Weil-Petersson gradient flow of renormalized volume and 3-dimensional convex cores. preprint (2020), https://arxiv.org/abs/2003.00337.
  • [BC1] M. Bridgeman and R. Canary. Bounding the bending of a hyperbolic three-manifold. Geom. Dedicata 96(2003), 211–240.
  • [BW] M. Bridgeman and Y. Wu. Uniform bounds on harmonic Beltrami differentials and Weil-Petersson curvatures. J. Reine Angew. Math. 770(2021), 159–181.
  • [BH] M. Bridson and A. Haefliger. Metric Spaces of Non-Positive Curvature. Springer-Verlag, 1999.
  • [BB] J. Brock and K. Bromberg. Geometric inflexibility and 3-manifolds that fiber over the circle. Journal of Topology 4(2011), 1–38.
  • [Bro] K. Bromberg. Hyperbolic cone-manifolds, short geodesics and Schwarzian derivatives. J. Amer. Math. Soc. 17(2004), 783–826.
  • [Brm] K. Bromberg. Rigidity of geometrically finite hyperbolic cone-manifolds. Geom. Dedicata 105(2004), 143–170.
  • [Cal] Eugenio Calabi. On compact, Riemannian manifolds with constant curvature. I. In Proc. Sympos. Pure Math., Vol. III, pages 155–180. American Mathematical Society, Providence, R.I., 1961.
  • [Eps] C. Epstein. Envelopes of horospheres and Weingarten surfaces in hyperbolic 3-spaces. preprint Princeton University, (1984), available at https://www.math.upenn.edu/~cle/papers/WeingartenSurfaces.pdf.
  • [EMM] D. B. A. Epstein, A. Marden, and V. Markovic. Quasiconformal homeomorphisms and the convex hull boundary. Ann. of Math. (2) 159(2004), 305–336.
  • [FN] Werner Fenchel and Jakob Nielsen. Discontinuous Groups of Isometries in the Hyperbolic Plane. De Gruyter, 2011.
  • [HK] C. Hodgson and S. Kerckhoff. Rigidity of hyperbolic cone-manifolds and hyperbolic Dehn surgery. J. Diff. Geom. 48(1998), 1–59.
  • [KT] Yoshinobu Kamishima and Ser P. Tan. Deformation spaces on geometric structures. In Aspects of low-dimensional manifolds, volume 20 of Adv. Stud. Pure Math., pages 263–299. Kinokuniya, Tokyo, 1992.
  • [KS] K. Krasnov and J-M. Schlenker. On the renomalized volume of hyperbolic 3-manifolds. Comm. Math. Phys. 279(2008), 637–668.
  • [Weil] A. Weil. On discrete subgroups of Lie groups. Annals of Math. 72(1960), 369–384.