-bounds for drilling short geodesics in convex co-compact hyperbolic 3-manifolds
Abstract
We give -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 -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 and a collection of disjoint simple closed geodesics in , the manifold also supports a complete hyperbolic structure . If we insist that and have the same ending data then is unique. If is closed, or more generally finite volume, and the elements of are sufficiently short then Hodgson-Kerckhoff [HK] developed a theory of a hyperbolic cone-manifolds that allows one to continuously interpolate between and 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 and .
One can give precise meaning to comparing the geometry of and . For example, one can compare the length of curves in to those in . 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 -norm. In [Bro], the second author bounded the -norm and these bounds played an important role the in resolving the Bers density conjecture. While -bounds always imply -bounds for all , 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 -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 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: will be a compact, hyperbolizable 3-manifold with boundary with interior and will be a collection of disjoint simple closed curves in . The is a complete, conformally compact hyperbolic structure on where the curves in are geodesics and is a one parameter family of hyperbolic cone-manifolds with cone locus and cone angles . We also assume the conformal boundary is fixed throughout the definition.
Theorem 1.1 ([Bro, Theorem 1.2])
There exists an such that if all geodesics in have length in then the cone deformation exists for where is a complete, hyperbolic structure on .
While the conformal boundary will be a fixed conformal structure the deformation, the complex projective structure on will change. We denote this one parameter family of projective structures by . The derivative of a path of projective structures on is naturally a holomorphic, quadratic differential. We denote the tangent vectors to by the holomorphic quadratic differentials . Our main results is the following bound on the -norm of .
Theorem 1.2
If is the sum of the length the geodesics in in then
As an immediate application we obtain the following -bounds on the change in projective structure.
Theorem 1.3
There exists an and such that the following holds. Let be a conformally compact hyperbolic 3-manifold and a collection of simple closed geodesics in each of length . Let be the unique complete hyperbolic structure on such that the inclusion is an isomorphism of conformal boundaries. If and are the projective structures on the conformal boundaries of and and the holomorphic quadratic differential is Schwarzian derivative between them then
where is the sum of the lengths of the components of in .
We note that the -bounds have universal constants compared to the -bound in Theorem 1.3 in [Bro] which depended on injectivity radius of the boundary hyperbolic structure. In [BW], -bounds on quadratic differentials are obtained from -bounds. These bounds again depend on the injectivity radius but they produce stronger bounds than those obtained in [Bro]. However, in [Bro] cone angles 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 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 -norm can be bounded by the length of the curves in the cone locus. We would like to use the bound on the -norm in the 3-manifold to bound the -norm of the quadratic differentials 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 -norm on the end in terms of the -norm of the quadratic differential. To calculate the -norm of the actual harmonic deformation we would like the model deformation to be orthogonal (in the -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 -norm of the quadratic differential in terms of the -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 -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 be the usual compactification of by . Note that isometries of extend to projective automorphisms of and that the group of isometries/projective transformations is . If is a 3-manifold with boundary a -structure on is an atlas of charts to with transition maps restrictions of elements of . On , the interior of , this a hyperbolic structure. On the boundary this is a complex projective structure. In this paper we will be interested in a special class of -structures, conformally compact hyperbolic cone-manifolds.
Let be a compact 3-manifold with boundary with interior and let be a collection of simple closed curves in the interior of . Let . A hyperbolic cone metric on with cone angle along is a hyperbolic metric on the interior of whose metric completion is homeomorphic to the interior of and in a c of each component the metric is that of a singular hyperbolic metric with cone angle . That is in cylindrical coordinates the metric will locally have the form
where is measured modulo the cone angle and the singular locus is identified with the -axis.
The hyperbolic metric is conformally compact if the hyperbolic structure on extends to a -structure on . We then have:
Theorem 2.1 ([HK, Brm])
Given a cone angle there exists a length such that the following holds. Let be a conformally compact hyperbolic cone-manifold with all cone angles and assume that the tube radius about the singular locus is . If each component of the singular locus has length then there a one parameter family of conformally compact hyperbolic cone-manifolds with the conformally boundary fixed and cone angle .
This one parameter family of cone-manifolds will induces a one parameter family of projective structures on the boundary where the conformal structure of is fixed. We will be interested in controlling the change in this projective structure as the parameter varies.
2.1 Flat -bundles
The Lie algebra can be interpreted geometrically as the space of infinitesimal automorphisms of . These are vector fields on whose flow are elements in so that on , the flow will be isometries of the hyperbolic metric, while on the flow will be projective automorphisms. A -structure on determines a flat -bundle over . We examine this bundle when it is restricted to the hyperbolic structure and when it is restricted to the projective boundary.
Hyperbolic structures
Let be the interior of . Then a -structure is a hyperbolic structure and the bundle has a natural decomposition and metric structure that we now describe. Each fiber is the space of germs of infinitesimal isometries. In particular, is a vector field in a neighborhood of so is a vector in . As is a complex bundle we can multiply by and then is another vector in . Then the map from to given by is bundle isomorphism. In fact, the map is a complex vector bundle isomorphism from to the the complexification of the tangent bundle. This isomorphism from to gives a decomposition of sections of into real and imaginary parts.
If is a vector in , we define such that under our isomorphism from to we have . Then is the infinitesimal translation with axis through and is an infinitesimal rotation about . Note that is real and is imaginary, as one would expect.
As is isomorphic to , the dual bundle is isomorphic to . The hyperbolic metric on determines an isomorphism from to and therefore an isomorphism from to . Note that this isomorphism is -linear but is -anti-linear with respect to the complex structures on and . For sections of we let be the dual section of . When going from to we replace the with a .
As is a complex vector space we have and more generally for alternating tensors with values in we have
In particular, every -valued form is locally the sum of terms where is a complex valued function, is a section of , and is a -valued form. The (and ) operators extend to -valued forms and we have . We also linearly extend the Hodge star operator from real forms to -valued forms so that . This extends to a linear map from to . We the define the inner product
Note that the wedge product of an -valued form and an -valued form is real form so this is a real inner product. We also let be the -norm of an -valued form.
If is either an -valued or -valued form we define the pointwise norm by . Then is the usual -norm of the function .
If is the flat connection for then define the operator on by
Then the formal adjoint for satisfies the formula
Note that if is a flat section () then the real and imaginary parts, and , will not be flat. That is will not preserve our bundle decomposition. Instead we define operators and such that where preserves the bundle decomposition and permutes it. That is for a real section we have that is a real -valued 1-form while is imaginary. We have formulas for both and . If is a vector field then
where is the Riemannian connection for the hyperbolic metric and
where is the Lie bracket. Note that the operator is purely algebraic.
We also have
This is a manifestation of the fact that the -operator is -anti-linear.
The Laplacian for -valued 1-forms is and is harmonic if . If is compact then this is equivalent be closed and co-closed. However, our manifolds will be non-compact so we will define to be a Hodge form if it is closed, co-closed and the real and imaginary parts are symmetric and traceless.
A computation in
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 with and the usual basis and dual basis at each . We also let
be tangent vectors in the complexified tangent space with dual 1-forms and . We can then write any -valued 1-form on as a sum of and terms.
The Lie algebra can be identified with traceless 2-by-2 matrices in , projective vector fields on and infinitesimal isometries of . The reader can check the correspondence given in the following lemma:
Lemma 2.2
An element of given by the matrix
is equivalent to the projective vector
Along the axis in the upper half space model of they are both equivalent to the constant section
At we also have
To calculate we note that
Lemma 2.3
Let be a parabolic vector field on a neighborhood of a point in a hyperbolic manifold . Let be a unit vector orthogonal to the horosphere tangent to , pointing away from the fixed point of and the dual -valued 1-form. Then
where is a -linear 1-form with . Furthermore .
Proof: We can assume that and in the upper half space model. Then and . To calculate we use Lemma 2.2 to write and as matrices and calculate the Lie brackets using matrix multiplication and get
so . We can then compute to see that .
Complex projective structures
We will be interested in complex projective structures that have a fixed underlying conformal structure . The space of such projective structures has a natural affine structure as the space of holomorphic quadratic differentials on . That is the difference of and in is quadratic differential in defined as follows. Let and be charts for and . Then is a conformal map for an open neighborhood in to . The Schwarzian derivative is a holomorphic function on and it determines a holomorphic quadratic differential . Properties of the Schwarzian derivative imply that if is a third projective structure then
This gives a canonical identification of the tangent space with .
If is a conformal metric on and then the ratio is a positive function on . More concretely in a local chart can be written as and the conformal metric can be written as , where is a positive function and is the Euclidean metric. Then the ratio , defined in the chart, is a well defined function on the surface. We denote this function by and let be the -norm of this function with respect to the 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 from our notation.
For every conformal structure there is unique Fuchsian projective structure . We let .
If is a smooth path in then its tangent vectors lie in . To prove our main result, Theorem 1.3, we bound the distances between the endpoints of a path by bounding the norms of the derivative .
We also describe how a quadratic differential determines a cohomology class in . This was originally introduced by the second author in [Brm].
Define a section of by
Then if is represented in a projective chart by we defined an -valued 1-form in the chart by
One can then check that this gives a well defined -valued 1-form on . As both and are holomorphic, is closed and therefore determines a cohomology class in .
Deformations of -structures
Let be a vector field on an open neighborhood in that is conformal on . We then define a section of as follows. For let be the unique infinitesimal isometry that agrees with at and whose curl agrees with the curl of at . On , the vector field is (the real part) of the product of a holomorphic function and . For each let be the complex quadratic polynomial whose 2-jet agrees with at . Then .
Let be a 1-parameter family of -structures on a 3-manifold with boundary with the conformal boundary fixed. For each , the time zero derivative of the path is vector field on that is conformal on . This determines a section of and is an -valued 1-form on . While the sections will not necessarily agree on overlapping charts, the -valued 1-forms will agree and determine an -valued 1-form on . As locally is of a section, is closed and therefore represents an element of .
Since the conformal boundary is some fixed conformal structure , the -structures on determine a family of projective structures . The time zero derivative of will be a holomorphic quadratic differential .
Proposition 2.4 ([Bro, Theorem 2.3])
The restriction of to the projective boundary is .
2.2 Hyperbolic metrics on ends
Let have a -structure. A convex surface in cuts off a conformally compact end if has two components and the outward component is homeomorphic to with . Then is the metric closure of the restriction of the outward component to and it is homeomorphic to with the original convex surface. We also let be the union of with the projective boundary so that .
The unit tangent vectors to the geodesic rays in orthogonal to define a vector field on . We can choose this product structure such that is the time flow of this vector field. If is the the time image of under this normal flow, then the hyperbolic metric for is can be written as a product of the induced metrics on and . However, it will be convenient to parameterize the surfaces in the parameter rather than in the time parameter 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 be a convex surface cutting off a conformally compact end with conformal boundary . Then there exists a conformal metric on and a bundle endomorphism of such that the hyperbolic metric on is given by
where
Proof: Let be the induced metric on and the shape operator and let be the distance normal flow of in . We also let
Then the induced metric on is given by
(see [KS, Lemma 2.2]). To get our representation of the hyperbolic metric in we need to rewrite in terms of a conformal metric on .
The conformal structure on the boundary is induced from the conformal structure on . If we multiply our metric n by the conformal structure doesn’t change but the new metric will extend continuously to
on so is a conformal metric on and . As is convex, the eigenvalues of are non-negative. Therefore we define . It follows that
as claimed.
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 be the a component of the conformal boundary of a conformally compact hyperbolic cone manifold and is the hyperbolic metric on . Then for all there is a convex surface that cuts of an end such that is the metric at infinity for .
The metric in a chart
If is a projective chart for then we can extend it to a chart where is the continuous extension of to a map to that is an isometry on . We say that the chart is adapted to if where the coordinates on the right are in the upper half space model for . We can always construct a chart adapted to by taking any projective chart with and post-composing with an element of .
In a projective chart the metric at infinity is scalar function times the Euclidean metric . If the chart is adapted to then we can calculate the value of this function.
Lemma 2.7
If is a chart adapted to then on the metric is of the form where is smooth and .
Proof: Define a function with . If is the metric for the upper half space model of then extends continuously to the metric on . Since we have that and therefore extends continuously to at .
On a chart for we have the usual coordinate vector fields and along with the vector fields in and in the complexified tangent bundle. On a chart for the end these coordinate vector fields, along with are a basis but, unlike in the upper half space model for , the vector fields and may not be orthogonal or of the same length as the operators are not conformal. In particular, the complex 1-form will not be -linear on the complex structure on induced by the metric . However, we can write down -linear and -anti-linear forms in terms of the Beltrami differential of the endomorphisms that define .
We begin with a computation on a single vector space. The usual Euclidean metric on has a unique -linear extension to . Then and are the usual dual basis for . While they are both -linear on , when restricted to , with the complex structure induced by , is -linear while is -anti-linear. A linear isomorphism has a unique -linear extension to . The Beltrami differential for is
where and are complex numbers with
We have the following:
Lemma 2.8
Let be a linear isomorphism and let and the Beltrami differential for . Then
where and are -linear and -anti-linear on with respect to . If is the Hodge star operator for then
Furthermore
Proof: As and are -linear and -anti-linear for the complex structure on induced by , and are -linear and -anti-linear for the complex structure induced by . As is -linear on and
for complex numbers and we have
Dividing by and by we define and by
so that is -linear and is -anti-linear on the complex structure induced by . Inverting gives our formula for and in terms of and .
As and are -linear and -anti-linear with respect to we have
or
We then have
Multiplying, we obtain the stated formulas. For the norm we note that define . Then
We can apply the above to the metrics . The Beltrami differential for endomorphisms can be written as where
We obtain the following immediate corollary.
Corollary 2.9
Let be a projective chart for with corresponding chart for . Then
where the are the smooth functions on given by
Further for ,
and
2.3 Model deformations
If is a convex surface cutting of a conformally compact end with projective boundary we can use Theorem 2.5 to extend to . Let
be given by . We would like to extend to a bundle map between and . 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 are paths in between and and determine isomorphisms between the fiber of over and the fiber of over . Using this isomorphism we can extend to a bundle map
We then extend to a 1-form in by pulling back via .
Lemma 2.10
Proof: We calculate at a point by taking a chart adapted to . In this chart is written as
While this expression does not depend on , the Hodge star operator and the dual map will. In particular the expression
depends on as both and depend on .
By taking the conjugate of in Corollary 2.9 we have
We also need to find . As we are working in a chart adapted to , we have . Since , by Lemma 2.3 at we have
By Lemma 2.7 we have and and combining our calculations we have the result.
Let be portion of the end cutoff by and let be inner product on . For an -valued form on we then define . Integrating the prior lemma we immediately get:
Corollary 2.11
We have and
The form is not harmonic as . However, we will show that decays rapidly in .
We’ll break the estimate into small calculations.
Lemma 2.12
Let be a neighborhood of and let be a smooth section of such that the function on extends continuously to . Then the projective vector field has a zero at . If then is a is a multiple of and
Proof: We write
where the functions are smooth, complexed valued functions on . By Lemma 2.2 we have
for . If extends continuously to then we must have so, as a projective vector field, is zero at .
If the we must further have that and . Since also exists it follows that
and therefore
Lemma 2.13
Let be a neighborhood of and let be a smooth section of such that the function on extends continuously to . Then
Proof: By Lemma 2.12 the condition that the norm extends to zero on implies that for some smooth complex valued function on . We have that
where is -derivative of and is the volume form for . Therefore
Note that the -derivative of the section is
and has bounded norm on . Therefore, on , the norm of is bounded and if we let be bound of we have
The bound of the norm of is similar once we note that the -derivative of is zero so the norm of the -derivative of is also zero at .
We now prove our bounds on .
Lemma 2.14
Proof: We have with . Therefore to bound we need to bound the norm of and .
In a chart by Corollary 2.9 we have
where the are smooth, complex valued functions on . Therefore
As and are holomorphic in the -coordinate we have that . Therefore
and, as the sections have norm limiting to zero on , by Lemma 2.13
Next we calculate . We will work in a chart adapted to and and a conformal coordinate at . Again applying Corollary 2.9 we have
where the are as above. We use Lemma 2.3 to calculate . At the point we have , and for some scalar . Then
where . Then
where is the volume form and . Then
since . By Corollary 2.9 , giving
As the -operator is an isometry and is compact this implies that there exist a such that
We then have
Here is the area form for the surface and we are using the fact that the area of these surfaces is bounded by for some . Therefore
and the lemma follows.
We now prove the main result of this section.
Theorem 2.15
Let where the section of has finite -norm. The for all
Proof: We have
By Corollary 2.11 we have
For the middle term we have integrate over the compact manifold and let . We have
where the integrals over and are both zero since restricted to these surfaces is zero as it contains a -term.
As , applying Lemma 2.14 we get
By the infinitesimal inflexibility theorem [BB, Theorem 3.6] we have for
Therefore as ,
If is a Hodge form on a conformally compact hyperbolic cone-manifold that is cohomologous to some on and end then, by definition, for some -valued section on . To apply this theorem we need the extra property that has finite -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 for every component 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 such that the following holds. Let be a hyperbolic cone-manifold such that all cone angle , every component of the cone cone locus has length and every component has a tubular neighborhood of radius . Further assume that these neighborhood are mutually disjoint. Then each component of the cone locus of length and cone angle has a tubular neighborhood of radius where
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 such that the following holds. Let be a conformally compact hyperbolic cone-manifold such that all cone angle and there are a disjoint collection of tubular neighborhoods of the components of the cone locus of radius . Further assume that if is a component of the cone locus with cone angle and length then
Then for every component of the projective boundary of we have
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 of projective structures on the hyperbolic disk is equivalent to the space of locally univalent maps with the equivalence if for some . We can identify with the space of quadratic differentials by mapping to its Schwarzian derivative . Then the topology on is the compact-open topology on .
Thurston described a natural parameterization of by the space of measure geodesic laminations on . We briefly review this construction.
A round disk shares a boundary with a hyperbolic plane . Let be the nearest point projection to and be the normal vector to at pointing towards . We can use these maps to define a version of the Epstein map for . In particular define by where is the unique round disk with respect to such that and let . (For the existence of this disk see [KT, Theorem 1.2.7].) We also define and .
The image of is a locally convex pleated plane. More precisely, let be measured geodesic laminations on and the subspace of laminations with finite support. That is if it is the union of a finite collection of disjoint geodesics with positive weights . Then determines a continuous map , unique up to post-composition with isometries of , that is an isometry on the complement of the support of and is “bent” with angle at . By continuity we can extend this construction to a general . An exposition of the following theorem of Thurston can be found in [KT].
Theorem 3.3
Given there exists maps and and a lamination such that is a locally, convex pleated surface pleated along , and the map is a homeomorphism from . Furthermore the maps and 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 and a transverse arc, we let be the -measure of . We then define the average bending norm to be
If is a lift of a measured lamination on a closed hyperbolic surface, then is bounded but in general may be infinite. For simplicity, we will let .
We have the following compactness result;
Lemma 3.4
Given then the set is precompact.
Proof: Let be the space of (unoriented) geodesics in the hyperbolic plane. We define the space of geodesic currents to be the space of non-negative Borel measures on with the weak∗ topology. The topology on is that of a closed subspace of . Given an open geodesic arc , we let be the set of all geodesics transverse to . We define
Then is an open cover of .
We let be the set of continuous functions on with support subordinate to the cover . Then for each there exists a with . We have the map given by . This map is a homeomorphism onto its image.
If then there is a with . Therefore for
Therefore is homeomorphic to a subset of which is compact by Tychanoff’s theorem. Therefore is precompact.
Corollary 3.5
Given there exists an such that if is a locally univalent map with then
.
Proof: We consider the family of with . Then by Thurston, is the image of under a homeomorphism. Therefore is precompact and has compact closure . Therefore there is an such that for all then
Therefore for all . As the set is invariant under isometries of it follows that for all .
3.3 Convex Hull of Cone Manifold
In this section will be a conformally compact hyperbolic cone-manifold with incompressible boundary and all cone angles . We let be the quadratic differential on the conformal boundary given by uniformization. In [Bro], the second author studied the convex core boundary of . This is given by taking the Epstein surface for the projective metric which we denote by . By [Bro, Proposition 6.5] the surface is an embedded locally convex surface in bounding an end of homeomorphic to . Also does not contain any cone axes in its interior. The surface has intrinsic hyperbolic metric and has a bending lamination . We identify the universal cover with the hyperbolic disk and obtain a lamination .
First some elementary lemmas about balls in hyperbolic cone-manifolds.
Lemma 3.6
Let be the unit sphere in . Let be cylindrical coordinates on and for define the spherical cone-surface
If then for some .
Proof: Assume not. We first take the case of . Then the unit sphere. Then letting be an open disk of radius about , we have
It follows that giving our contradiction.
For we take a fundamental wedge domain for in above, and can assume the are in the interior. Then by the spherical case two of the points have . As we obtain our contradiction.
We have the following elementary calculation on half-spaces in ;
Lemma 3.7
Let be given by
Let be half-spaces in such that are disjoint. If each intersects then .
Proof Let be the distance from to and let have spherical radius . Then we have and where is the spherical radius of where is a half-space a distance from . Therefore as each contains a disk or radius , if the are disjoint, then there are 3 disks of radius which are disjoint.
We show that . We let have the spherical metric given by angle subtended at . If then the centers of satisfy contradicting Lemma 3.6.
We have a right-angled hyperbolic triangle with sides and angle between. Let be the length of the other side. Then solving we have
and by the hyperbolic Pythagorean formula
We now consider balls in our cone manifold . We let be the universal cover with convex hull . The end lifts to a component of the complement of with boundary . As has incompressible boundary, then is trivial.
The space is a hyperbolic cone manifold and the cone axes lift to . For we define balls in the usual way, i.e. . We note that may not be topologically a ball or isometric to a hyperbolic ball. For a point , we define to be the maximum radius such that is embedded and isometric to a hyperbolic ball of radius . Note for , and otherwise and equals is the injectivity radius of in . For we further define to be the minimum distance to the cone axes .
We first bound the average bending for points with bounded below.
Lemma 3.8
Let be a conformally compact hyperbolic cone-manifold with incompressible boundary and all cone angles . Let and a closed geodesic arc on with midpoint and length less than . Then
Proof: We let be the 1-parameter family of support half-spaces from to . We consider and disks . We let be the smallest such that have disjoint interiors. Then we have . If there is no such then we have and we’re done.
We now let be the smallest such that have disjoint interiors. Again it follows that giving . If no such exists then and we are also done.
We first show that do not intersect. If do intersect, we extend to a closed curve by joining by a piecewise geodesics on . We note that is simply connected. We get our contradiction by showing that curve in is homotopically non-trivial. The curve is homotopic to a simple closed curve in via a homotopy in . But as arc is transverse to a bending line then separates the points in . Therefore is non-trivial in . As has incompressible boundary, is trivial and we obtain our contradiction. Thus if exists, then do not intersect. We then obtain a contradiction from the above lemma as are disjoint in and intersect .
We use the same argument as above to bound average bending for points close to the cone axes.
Lemma 3.9
Let be a conformally compact hyperbolic cone-manifold with incompressible boundary and all cone angles . Let have an embedded tube of radius and with . If is a closed geodesic arc on with midpoint and length less than then
Proof: We let be the nearest point of to on and consider . If is a geodesic arc of length centered about , then is in . We let , then is a sphere with two cone points. We again consider the 1-parameter family of support planes from to and let . Then are disks in whose interior are disjoint from the cone points. Then analysing as in Lemma 3.7, we obtain 3 disks with disjoint interiors on . By Lemma 3.6 the disks cannot be disjoint which gives a contradiction. Thus we have .
To bound our average bending uniformly for a given length, reduces now to showing that 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 be a conformally compact hyperbolic cone-manifold with incompressible boundary and all cone angles . Let be the cone-axes and for let be the neighborhood of in . Let be such that are embedded and disjoint.
-
•
If then for some and .
-
•
if then and
Proof: We to be the completion of the universal cover of . Then is CAT(0). We let be the completions of the universal covers of .
Let . If is achieved by an arc joining to an axis then and as for then the first statement follows. Otherwise there is a non-trivial geodesic of length in . Therefore must link a finite collection of axes of elements . If links more than one axis then is greater than the length of the shortest closed geodesic linking the axes. As this is at least , then as before. Therefore we can assume links a single axis . Then lifts to a piecewise geodesic in which is invariant under the action of the deck transformation corrseponding to . As is CAT(0) and the are convex and complete, projection onto is distance decreasing (see [BH, Proposition II.2.4]). Thus as commutes with the action of , the curve descends to a curve in of length which is contained in linking with basepoint . Thus and the first item is done.
For then trivially . We now describe the relation between . Then projecting as above, we have that is attained by a curve in . We take a fundamental domain for to be a wedge of a hyperbolic tube of radius about a geodesic with wedge angle and have be on the central radial line of the wedge. Taking the largest ball about that is embedded in it follows that if then we have that . Otherwise and then is the length of the unique shortest geodesic arc with both endpoints . Thus are sides of a right angled triangle, with hypothenuse and side of length facing angle . Therefore for by the hyperbolic sine formula
Lemma 3.11
There is an explicit monotonic increasing function such that the following holds. Let be a conformally compact hyperbolic cone-manifold with incompressible boundary and all cone angles satisfying the conditions of Theorem 3.2. Then for , if with then .
Proof: By assumption for , has embedded tubular neighborhood of radius such that . We let . We lift the tubular neighborhoods to and denote by the lift for .
By the Lemma 3.10, we need only consider points in the neighborhoods . Thus we let . Again by Lemma 3.10 if then giving . Similarly for , we have
This gives a bound on for .
We now consider . By [Bro, Lemma 3.3] all support half-spaces are embedded in . Let be a half space intersecting with distance from the cone axis. We take a wedge fundamental domain with the nearest point of being centered. Then in order for to be embedded in , it cannot intersect the radial sides of the wedge. Therefore we must have where form a right-angled triangle with hypothenuse and angle between the sides . Labeling the other side of the triangle we have by hyperbolic geometry (see [FN, formulas III.5, III.6])
Thus for we have . Therefore substituting
To obtain a bound, we use our assumptions in Theorem 3.2 and applying Theorem 3.1 we have
Therefore
It follows that
As and is monotonic, then
Thus for we have with
Thus combining the bounds, we have with
In [BC1], the first author and Canary proved the following.
Theorem 3.12 ([BC1, Theorem 3])
Let be an embedded convex pleated plane then its bending lamination satisfies
for
Proposition 3.13
Let be a conformally compact hyperbolic cone-manifold with incompressible boundary and all cone angles satisfying the conditions of Theorem 3.2. Then
for any .
Proof: We let where .
If . Then as we have
Therefore we can apply Lemma 3.11 to to get . Therefore for then for any geodesic arc of length less than centered at .
If , as has an embedded tubes of radius then by Lemma 3.9 if is an arc of length then .
Combining the bounds we have
For . As and we can assume , then
We now prove the main result of this section.
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 with a collection of disjoint, simple closed curves in the interior. We will examine a family of conformally compact hyperbolic cone-manifold structures on with cone locus .
Theorem 4.1 ([Bro])
Let be a one parameter family of cone-manifolds given by Theorem 1.1 and let be the length of a component of in and the sum of the .
-
•
-
•
In each there is a union of embedded, disjoint tubular neighborhoods of the components of of radius .
-
•
The time derivative of is represented by a model Hodge form with
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 rather than the larger radii assumed in that proposition.
We are now ready to prove our main theorem bounding the -norm of the derivative of the path of complex projective structures.
Proof of Theorem 1.2: We assume has been fixed throughout the proof.
For the path of complex projective structures on the boundary of , by Theorem 3.2 we have that . Therefore by Theorem 2.6 there is a convex surface in cutting of an end such that is the metric at infinity for . Note that while will be disjoint from the cone locus in it may intersect the tubular neighborhood of the cone locus. To correct this we need to remove the collar of width from . This is the end where .
By Theorem 4.1 we have that
and since this implies that
As is a model Hodge form Theorem 4.1 implies that
As
this gives
where
Our main results now follows immediately.
Proof of Theorem 1.3: Integrating the above, we get the -bound
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.