Download as pdf or txt
Download as pdf or txt
You are on page 1of 80

THE WEIL-PETERSSON GEODESIC FLOW IS ERGODIC

K. BURNS, H. MASUR AND A. WILKINSON


arXiv:1004.5343v1 [math.DS] 29 Apr 2010

Abstract. We prove that the geodesic flow for the Weil-Petersson met-
ric on the moduli space of Riemann surfaces is ergodic and has finite,
positive metric entropy.

Contents
Introduction 2
0.1. The case of the punctured torus 8
0.2. A word about notation 9
1. Background on Teichmüller theory, Quasifuchsian space, and
Weil-Petersson geometry 9
1.1. Riemann surfaces and tensors of type (r, s) 9
1.2. Teichmüller and Moduli spaces 10
1.3. The bundle of projective structures on S 13
1.4. Quasifuchsian space 14
1.5. Fenchel-Nielsen coordinates 16
1.6. The Deligne-Mumford compactification of moduli space 17
2. Background on the geodesic flow 19
2.1. Vertical and horizontal subspaces and the Sasaki metric 20
2.2. The geodesic flow and and Jacobi fields 21
2.3. The Jacobi and Riccati equations 22
2.4. Perpendicular Jacobi fields 23
2.5. Consequences of negative curvature 23
2.6. Unstable Jacobi fields 24
2.7. Bounding the derivative of the geodesic flow 24
3. Bounds on the derivative of ϕ1 in the WP metric 27
3.1. Combined length bases 28
3.2. First and second order properties of the WP metric 31
3.3. Curvature estimates along a geodesic 32
3.4. Estimates on rα /fα 35
3.5. Controlled functions and bounds on Riccati solutions 37
3.6. Proof of Theorem 3.1 41
4. Bounding the second derivative of the geodesic flow 41
4.1. More on the Sasaki metric and statement of the general result. 41

Date: April 30, 2010.


Each author partially supported by the NSF..
1
2 K. BURNS, H. MASUR AND A. WILKINSON

4.2. Variations of solutions to linear ODEs 42


4.3. Proof of Proposition 4.2 43
5. Higher order control of the WP metric 47
5.1. Comparison of Weil-Petersson and Teichmüller metrics 48
5.2. Estimates on the WP metric in special coordinates 51
5.3. Proof of Proposition 5.1 54
6. General criteria for ergodicity of the geodesic flow 55
6.1. Proof of Proposition 6.9: verifying the conditions of Katok-
Strelcyn-Ledrappier 67
6.2. Additional conditions in [17] implying finite, positive entropy. 75
7. Ergodicity and finite entropy of the WP geodesic flow 75
7.1. Ergodicity of the flow on T 1 (T /MCG[k]) 76
7.2. Ergodicity of the flow on M1 (S): Proof of Theorem 1 78
References 78

Introduction
This paper is about the dynamical properties of the Weil-Petersson geo-
desic flow for the moduli space of Riemann surfaces. Our main result is that
this flow is ergodic: any invariant set must have volume zero or full volume.
Ergodicity implies that a randomly chosen, unit speed Weil-Petersson geo-
desic in moduli space becomes equidistributed over time. What is more, the
tangent vectors to such a geodesic also become equidistributed in the space
of all unit tangent vectors to moduli space.
To state our result more precisely and to put it in context, we first review
the basic setup from Teichmüller theory. Let S be a surface of genus g ≥ 0
with n ≥ 0 punctures, and let M(S) be the moduli space of conformal
structures on S, up to conformal equivalence. Assume that 3g+n ≥ 4, which
implies that in each conformal class there is complete hyperbolic metric.
Then M(S) has the alternate description of the moduli space of hyperbolic
structures on S, up to isometry. The orbifold universal cover of M(S) is the
Teichmüller space Teich(S) of marked conformal structures on S.
It is a classical result due to Fricke and Klein that Teich(S) is homeomor-
phic to a ball of dimension 6g − 6 + 2n. Teichmüller space carries a natural
complex structure via a special embedding of Teich(S) into a complex repre-
sentation variety QF (S), called quasifuchsian space. Under this map, called
the Bers embedding, the image of Teich(S) sits as a complex subvariety
(indeed there is a biholomorphic equivalence QF (S) ∼ = (Teich(S))2 ). The
orbifold fundamental group of M(S) is the mapping class group MCG(S) of
orientation preserving homeomorphisms of S modulo isotopy. The mapping
class group acts holomorphically on Teich(S). The stabilizer of each point
is finite, which gives M(S) the structure of a complex orbifold.
There are several natural ways to define a MCG(S)-invariant distance on
Teich(S). One that has been particularly well-studied is the Teichmüller
THE WEIL-PETERSSON GEODESIC FLOW 3

metric dT , which measures the distance between two marked conformal


structures in terms of the minimal conformal distortion (“quasiconformal
dilatation”) of maps between the structures preserving the marking. The
Teichmüller metric is a Finsler metric (induced by a norm), which admits
a finite volume on the quotient. This metric induces a volume-preserving
geodesic flow, which embeds in an SL(2, R)-action on the unit cotangent
bundle of Teich(S). The dynamics of this Teichmüller geodesic flow have
been studied extensively: in particular, it is nonuniformly hyperbolic and
ergodic, and these properties turn out to have close connections with the er-
godic and dynamical properties of directional flows on translation surfaces.
See [42] for a survey of the properties of the Teichmüller geodesic flow.
Another naturally defined and well-studied metric on Teichmüller space,
and the focus of this paper, is the Weil-Petersson metric gW P , which is the
Kähler metric induced by the Weil-Petersson symplectic form ωW P and the
almost complex structure J on Teich(S):

gW P (v, w) = ωW P (v, Jw).

We refer to the Weil-Petersson metric as the WP metric, for short. As with


the Teichmüller metric, the WP metric descends to a metric on M(S) with
∧3g−3+n
finite volume determined by the volume form |ωW P |.
A striking feature of the WP metric is its intimate connections with hy-
perbolic geometry, among them:

• the hyperbolic length of a closed geodesic (for a fixed free homotopy


class on S) is a convex function along WP geodesics in Teich(S) [38];
• in Fenchel-Nielsen coordinates (ℓi , τi )3g−3+n
i=1 on Teich(S), the WP
P
symplectic form ωW P has the simple expression ωW P = 21 dℓi ∧ dτi
[34]
• the growth of the hyperbolic lengths of simple closed curves on S is
related to the WP volume of M(S) [19]; and
• the WP metric has a formulation in terms of dynamical invariants
of the geodesic flow on hyperbolic surfaces [5, 22].

The Weil-Petersson metric has several notable features that make it an


interesting geometric object of study in its own right. The WP metric is
negatively curved, but incomplete. The sectional curvatures are neither
bounded away from 0 (except in the simplest cases of (g, n) = (1, 1) and
(0, 4)), nor bounded away from −∞. The WP geodesic flow thus presents a
naturally-occurring example of a singular hyperbolic dynamical system, for
which one might hope to reproduce the known properties of the geodesic flow
for a compact, negatively curved manifold, such as: ergodicity, equidistri-
bution of closed orbits, exponentially fast mixing and decay of correlations,
and central limit theorem.
4 K. BURNS, H. MASUR AND A. WILKINSON

We summarize the previous literature on the WP geodesic flow. Wolpert


[36] showed that the geodesic flow is defined for all time on a full volume sub-
set of the the unit tangent bundle T 1 Teich(S) and thus descends to a volume-
preserving flow on the finite volume quotient M1 (S) := T 1 Teich(S)/MCG(S).
Pollicott, Weiss and Wolpert [27] proved in the case (g, n) = (1, 1) that the
geodesic flow is transitive on M1 (S) and periodic orbits are dense in M1 (S)
[27]. Brock, Masur and Minsky [6] proved transitivity and denseness of peri-
odic orbits for arbitrary (g, n) and also showed that the topological entropy
of the geodesic flow is infinite (that is, unbounded on compact invariant
sets). Hamenstädt [12] proved a measure-theoretic version of density of
closed orbits: the set of invariant Borel probability measures for the Weil-
Petersson flow that are supported on a closed orbit is dense in the space of
all ergodic invariant probability measures.
In this paper, we prove:
Theorem 1. Let S be a Riemann surface of genus g ≥ 0, with n ≥ 0
punctures. Assume that 3g + n ≥ 4. The Weil-Petersson geodesic flow on
M1 (S) is ergodic with respect to WP volume and has finite measure-theoretic
entropy.
Our basic approach to the problem is as follows. The WP geodesic flow ϕt
preserves a finite volume m on M1 (S), and one can show using properties
of the WP metric that log kDϕ1 k is integrable with respect to m. The
Multiplicative Ergodic Theorem of Oseledec then implies that there is a full
volume subset Ω ⊂ M1 (S) such that for every v ∈ Ω and every nonzero
tangent vector ξ ∈ Tv M1 (S), the limit
1
λ(ξ) := lim log kDv ϕt (ξ)k
t→±∞ t

exists and is finite. The real number λ(ξ) is called the Lyapunov exponent of
ϕt at ξ. Observe that if ξ is in the line bundle Rϕ̇(v) tangent to the orbits
of the flow, then λ(ξ) = 0. We say that ϕt is nonuniformly hyperbolic if for
almost every v ∈ Ω and every ξ ∈ Tv M1 (S)\Rϕ̇(v), the Lyapunov exponent
λ(ξ) is nonzero.
Using the fact that the WP sectional curvatures are negative, we prove
that the WP geodesic flow is nonuniformly hyperbolic. Nonuniform hyper-
bolicity is the starting point for a rich ergodic theory of volume-preserving
diffeomorphisms and flows, developed first by Pesin for closed manifolds,
and expanded by Sinai, Katok-Strelcyn, Chernov and others to systems
with singularities, such as the WP geodesic flow. The basic argument for
establishing ergodicity of such systems originates with Eberhard Hopf and
his proof of ergodicity for geodesic flows for closed, negatively curved sur-
faces [13]. His method was to study the Birkhoff averages of continuous
functions along leaves of the stable and unstable foliations of the flow. This
type of argument has been used since then in increasingly general contexts,
and has come to be known as the Hopf Argument.
THE WEIL-PETERSSON GEODESIC FLOW 5

The core of the Hopf Argument is very simple. Suppose that ψt is a C ∞


flow defined on a full measure subset Ω of an open Riemannian manifold V ,
preserving a finite volume on V . For any x ∈ Ω one defines the stable set:
W s (x) = {x′ ∈ Ω : lim d(ψt (x), ψt (x′ )) = 0},
t→∞
and the unstable set
W u (x) = {x′ ∈ Ω : lim d(ψt (x), ψt (x′ )) = 0}.
t→−∞

The stable (respectively unstable) sets partition Ω into measurable subsets.


The first step in the Hopf Argument is to observe that for any continu-
ous function f : V → R with compact support, the forward upper Birkhoff
average
Z
+ 1 T
f = lim sup f ◦ ψt dt
T →+∞ T 0
is constant on any stable set W s (x), and the the backward upper Birkhoff
average
Z
1 T
f − = lim sup f ◦ ψt dt
T →−∞ T 0
is constant on any unstable set W u (x). Both functions f + and f − are
evidently invariant under the flow ψt , and the Birkhoff and von Neumann
Ergodic Theorems imply that f + = f − almost everywhere. To show that
ψt is ergodic it suffices to show that f + is constant almost everywhere, for
every continuous f with compact support. The fundamental idea is to use
the properties of the equivalence relation generated by the stable sets, the
unstable sets, and the flow to conclude that f + = f − must be constant.
In the next step in the Hopf Argument, one assumes some form of hy-
perbolicity of the flow, which will imply that the stable and unstable sets
are in fact smooth manifolds. In the original context of Hopf’s argument,
V = Ω = T 1 S is the unit tangent bundle of a compact, negatively curved
surface S and ψt is the geodesic flow. In this setting, the stable and unstable
sets have a particularly nice description as gradient level sets of Busemann
functions. They are C ∞ , globally defined, and for ∗ ∈ {s, u}, the collection
W ∗ := {W ∗ (v) : v ∈ T 1 S}
defines a C 1 foliation of T 1 S. At each point v ∈ T 1 S, the tangent space
Tv T 1 S is spanned by the tangents to W s (v), W u (v) and the direction ψ̇(v) of
the flow. A local argument in C 1 charts using Fubini’s theorem shows that
any ψt -invariant function that is almost everywhere constant along leaves of
W s and W u must be locally constant, and hence globally constant, since T 1 S
is connected. In particular the function f + is constant for any continuous,
compactly supported f , and so ψt is ergodic.
Hopf’s original argument does not generalize immediately to geodesic
flows for higher dimensional compact, negatively curved manifolds. In this
higher-dimensional setting, the stable and unstable foliations W s and W u
6 K. BURNS, H. MASUR AND A. WILKINSON

exist, are gradient level sets of Busemann functions, and have C ∞ leaves. In
general, however they fail to be C 1 foliations (except when the curvature is
1/4-pinched) and so the argument using Fubini’s theorem in local C 1 charts
fails.
In the late 1960’s Anosov [1] overcame this obstacle by proving that for
any compact, negatively curved manifold, the foliations W s and W u are
absolutely continuous. Absolute continuity, a strictly weaker property than
C 1 , is sufficient to carry out a Fubini-type argument to show that any ψt -
invariant function almost everywhere constant along leaves of W s and W u
is locally constant. See Section 6 for a more detailed discussion of absolute
continuity. Anosov thereby proved that the geodesic flow for any compact
manifold of negative sectional curvatures is ergodic.
There is an extensive literature devoted to extending the Hopf Argument
beyond the uniformly hyperbolic setting of geodesic flows on compact neg-
atively curved manifolds. For smooth flows defined everywhere on compact
manifolds, Pesin [26] introduced an ergodic theory of nonuniformly hyper-
bolic systems. In short, Pesin theory shows that if ψt : V → V preserves a
finite volume and is nonuniformly hyperbolic, then almost everywhere the
stable and unstable sets are smooth manifolds. The family of stable mani-
folds is measurable and absolutely continuous in a suitable sense.
From Pesin theory, one deduces that a nonuniformly hyperbolic diffeo-
morphism of a compact manifold has countably many ergodic components
of positive measure. More information about the flow can be used in some
contexts to deduce ergodicity. The obstruction to using the full Hopf Argu-
ment in this setting is that stable manifolds are defined only almost every-
where, and they may be arbitrarily small in diameter, with poorly controlled
curvatures, etc.
In a somewhat different direction than Pesin theory, Sinai [32] introduced
methods for proving ergodicity of hyperbolic flows with singularities and
applied them in his study of the n-body problem of celestial mechanics.
Here the flow ψt locally resembles the geodesic flow for a compact, negatively
curved manifold, but globally encounters discontinuities and places where
the norms of the derivatives kDψt k and kD 2 ψt k become unbounded.
Introducing new techniques in the Hopf argument, Sinai was able to show
that for several important classes of systems, including some billiards and
flows connected to the n-body system, ergodicity holds. These arguments
have since been generalized to much larger classes of singular hyperbolic
systems and singular nonuniformly hyperbolic systems.
In the singular nonuniformly hyperbolic setting, all aspects of Hopf’s ar-
gument require careful revisiting. The mere existence of local stable mani-
folds is a delicate matter and depends in a strong way on the growth of the
derivative of ψt near the singularities. To give a sense of how delicate these
issues can be, we remark that:
THE WEIL-PETERSSON GEODESIC FLOW 7

• for compact surfaces of nonpositive curvature and genus g ≥ 2, it is


unknown whether the geodesic flow is always ergodic (even though
it is always transitive)
• there exist complete, finite volume surfaces of pinched negative cur-
vature (but unbounded derivative of curvature) whose stable folia-
tions are not even Hölder continuous [3];
• for C 1 nonuniformly hyperbolic systems that are not C 2 , stable sets
can fail to be manifolds [29];
• nonuniformly hyperbolic systems on compact manifolds can fail to
be ergodic and can even have infinitely many ergodic components
with positive measure [10].

A general result providing for the existence and absolute continuity of local
stable and unstable manifolds for singular, nonuniformly hyperbolic systems
was proved by Katok-Strelcyn [17]. We will use this work in an important
way in this paper.
Returning to the context of the present paper, the WP geodesic flow is a
singular, nonuniformly hyperbolic system. To prove that it is ergodic, the
first step is to verify the Katok-Strelcyn conditions to establish existence and
absolute continuity of local stable and unstable manifolds. In particular, one
needs to control the norm of the first two derivatives of the geodesic flow in
a neighborhood of the boundary of M1 (S).
To control the first derivative, we use the asymptotic expansions of Wolpert
for the WP curvature and covariant derivative found in [36, 35, 37], combined
with a careful analysis of the solutions to the WP Jacobi equations. This
is the content of Theorem 3.1. The precise estimates obtained by Wolpert
appear to be essential for these calculations.
Since Wolpert’s expansions of the WP metric are only to second order,
and we need third order control to estimate the second derivative of the
flow, we borrow ideas of McMullen in [21]. There is a nonholomorphic (in
fact totally real) embedding of Teich(S) into quasifuchsian space QF (S),
under which the WP symplectic form has a holomorphic extension. This
holomorphic form is the derivative of a one-form that is bounded in the
Teichmüller metric. Using the Cauchy Integral Formula and a comparison
formula between Teichmüller and WP metrics, one can then obtain bounds
on all derivatives of the WP metric. This is the content of Proposition 5.1.
These bounds are adequate to control the second derivative of the geodesic
flow, using the bounds on the first derivative already obtained.
Once the conditions of [17] have been verified, we are guaranteed the
almost everywhere existence of absolutely continuous families W s and W u
of stable and unstable manifolds. These families are only measurable and
a priori their local geometry can be highly nonuniform. At this point, we
use negative curvature and another key property of the WP metric called
geodesic convexity to show that in fact W s and W u have well-controlled local
geometry.
8 K. BURNS, H. MASUR AND A. WILKINSON

As a by-product of our arguments, we obtain that the WP Busemann


function for almost every tangent direction to Teich(S) is C ∞ (see Propo-
sition 6.10). The local geometry of W s and W u is sufficiently nice that
Hopf’s original argument can be used with small modifications. In par-
ticular, none of the more complicated local ergodicity arguments, such as
the “Hopf chains” developed by Sinai, are necessary. We also obtain pos-
itive, finite entropy of the WP flow using results of Katok-Strelcyn and
Ledrappier-Strelcyn in [17].
The paper does not follow the structure of this outline exactly. Rather
than restricting to the special case of the WP metric, we instead develop
an abstract criterion for ergodicity of the geodesic flow for an incomplete,
negatively curved manifold. This is carried out in Section 6, which may be
read independently of the rest of the paper. The remainder of this paper is
devoted to setting up and verifying the conditions in Section 6.

0.1. The case of the punctured torus. Several interesting features of


the WP metric are already present in the simplest cases (g, n) = (1, 1) and
(0, 4). When S is the once-punctured torus, Teich(S) is the upper half
space H, the Teichmüller metric dT is the hyperbolic metric, and M(S) is
the classical moduli space of elliptic curves H/PSL(2, Z), which is a sphere
with one puncture and two cone singularities of order 2 and 3.
The mapping class group MCG(S) is the modular group SL(2, Z). Due
to the presence of torsion elements in PSL(2, Z), the space M(S) is not a
manifold, but the finite branched cover H/Γ[k], for k ≥ 3 is a manifold,
where Γ[k] is the level-k congruence subgroup
Γ[k] = {A ∈ PSL(2, Z) | A ≡ I mod k}.
The tangent bundle to Teich(S) is canonically identified with PGL(2, R),
and the unit tangent bundle in the Teichmüller metric with PSL(2, R).
There are global coordinates (ℓ, τ ) in Teich(S), the so-called Fenchel-
Nielsen coordinates, which have the asymptotic (first-order) expansions
1 Re(z)
ℓ(z) ∼ , and τ (z) ∼ , as Im(z) → ∞,
Im(z) Im(z)
and the WP form has the first-order asymptotic expansion
1 1
ωW P = dℓ ∧ dτ ∼ dz ∧ dz, as Im(z) → ∞.
2 Im(z)3
Since the complex structure on Teich(S) is the standard one on H, we obtain
the expansion
2 |dz|2
gW P ∼ .
Im(z)3
By contrast,
|dz|2
gT2 = .
Im(z)2
THE WEIL-PETERSSON GEODESIC FLOW 9

A neighborhood of the cusp in M(S) is formed by taking the quotient of


the points above the line Im(z) = Im(z0 ), for Im(z0 ) sufficiently large, by
the mapping class element z 7→ z + 1. A model for this neighborhood is the
surface of revolution for the curve {y = x3 : x > 0}. From the form of the
metric one can see the incompleteness: a vertical ray to the cusp at infinity
has finite length. Nonetheless the metric is convex. Moreover the curvature
goes to −∞ as Im(z) → ∞.
Pollicott and Weiss [28] studied the model case of a negatively curved sur-
face whose singularities coincide with a surface of revolution for a polynomial
and proved ergodicity of the geodesic flow in this case.

0.2. A word about notation. Throughout the paper, we have strived to


minimize confusion in notation, but because we refer heavily in places to
the work of different authors and are combining techniques from disparate
areas, inevitable conflicts arise. Among the most notable are:
• the use of σ to denote a section of the bundle of projective structures
(see §1.3) a simplex in the curve complex (see §1.6), and a geodesic
(see §4);
• the use of Teich(S) to denote the Teichmüller space of the conjugate
Riemann surface S (see §1.4) and of Teich(S) to denote augmented
Teichmüller space (see §1.6)
• the use of α, β, κ and other Greek letters to denote, variously, a
simple closed curve, a variation of geodesics, a real number, a bundle,
a connection, a differential form and a map.
The notation is locally consistent, so it should be clear from the context
which use is intended; the reader should remain alert to these distinctions
in turning from one part of the paper to another.

Acknowledgments. The authors express their appreciation to Scott Wolpert


and Curt McMullen for many helpful conversations during the time this pa-
per was being written. We also thank Nikolai Chernov and Benson Farb for
useful discussions.

1. Background on Teichmüller theory, Quasifuchsian space,


and Weil-Petersson geometry
Much of the discussion in this section is based on McMullen’s paper [21].
Useful background can be found in [25] and the course notes [20].

1.1. Riemann surfaces and tensors of type (r, s). We begin with some
preliminary facts about Riemann surfaces. A Riemann surface is a topo-
logical surface equipped with an atlas of charts into C with holomorphic
transition maps. Suppose that X is a Riemann surface of genus g with n
punctures. Uniformization implies that X is conformally equivalent to a
10 K. BURNS, H. MASUR AND A. WILKINSON

quotient H/Γ, where H denotes the upper half plane, and Γ is a discrete
subgroup of P SL(2, R). The hyperbolic metric ρ̃ on H given by
|dz|
ρ̃(z) =
Imz
descends to a metric ρ on H/Γ of finite area, which is the unique Riemannian
metric of constant curvature −1 on X that induces the same conformal
structure.
Denote by κ the holomorphic cotangent bundle and by κ−1 the holo-
morphic tangent bundle of X, both of which are holomorphic complex line
bundles over X. For r an integer, we denote by κr the |r|-fold complex
tensor product ⊗|r| κ, if r ≥ 0, and ⊗|r| κ−1 if r < 0.
A tensor of type (r, s) on X is a section of the complex line bundle κr ⊗ κs
over X. Equivalently, a tensor of type (r, s) is the assignment of a func-
tion β(z) to each local holomorphic chart z on X such that β(z)dz r dz s is
invariant under local holomorphic transition maps. Thus, if w is another
holomorphic coordinate, and β̂(w) is the representing function, one has:
   s
dz r dz
β̂(w) = β(z(w)) .
dw dw
Since κr ⊗κs has a real analytic structure, it is natural to speak of measurable
or smooth sections; in the special case where s = 0, the bundle is complex
analytic and so the notion of holomorphic or meromorphic sections of κr is
also well-defined.
The sum, product, conjugate and modulus of (r, s) tensors are defined
in the obvious way. Observe that the square ρ2 of the hyperbolic metric
on X is a tensor of type (1, 1) and that the modulusR of any measurable
tensor ψ on X of type (1, 1) has a well-defined integral X |ψ|, which can be
infinite. More generally if ψ is an (r, s)-tensor and p ≥ 1, then ρ2−p(r+s) |ψ|p
is a (1, 1)-tensor. These observations lead to the construction of Lp norms
on the space of measurable (r, s) tensors, defined as follows; for ψ an (r, s)
tensor, and p ≥ 1 we define:
Z 1/p
kψkp := ρ2−p(r+s)
|ψ|p
kψk∞ := ess sup ρ−(r+s) |ψ|.
X X

These norms will give rise to the Teichmüller and WP metrics on Teichmüller
space, which we now define.

1.2. Teichmüller and Moduli spaces. A marked complex structure is a


Riemann surface X together with a homeomorphism f : S → X, where S
is a fixed Riemann surface. Given a marking surface S of genus g with
n punctures, we define the Teichmüller space Teich(S) to be the set of
equivalence classes of marked complex structures f : S → X, where f1 : S →
X and f2 : S → X2 are equivalent if there is a holomorphic map h : X1 → X2
isotopic to f2 f1−1 .
THE WEIL-PETERSSON GEODESIC FLOW 11

The Teichmüller distance dT between equivalence classes [f1 : S → X1 ]


and [f2 : S → X2 ] is defined to be 21 log K, where K is the minimum quasi-
conformal dilatation of a quasiconformal homeomorphism between X1 and
X2 isotopic to f2 f1−1 :
|∂h/∂z| + |∂h/∂z| 1
K= inf sup .
h≃f2 f1−1 X1 |∂h/∂z| − |∂h/∂z|
With respect to the the Teichmüller distance, Teich(S) is a complete metric
space, homeomorphic to R6g−6+2n . Uniformization gives an identification of
Teich(S) with an open component of the representation variety of homomor-
phisms from π1 (S) into the real Lie group P SL(2, R), modulo conjugacy;
this identification gives Teich(S) a real analytic structure. Teich(S) also
carries a compatible complex analytic structure, which we shall describe a
little later.
The mapping class group MCG(S) is the set of equivalence classes of ori-
entation preserving diffeomorphisms of S modulo isotopy, which forms a
discrete group under composition. MCG(S) acts properly by diffeomorph-
isms of Teich(S) via precomposition with the marking homeomorphisms
f : S → X; the quotient M(S) = Teich(S)/MCG(S) is easily seen to be the
moduli space of Riemann surfaces homeomorphic to S, modulo conformal
equivalence. The MCG(S)-stabilizer of any point in Teich(S) is finite. In
denoting an element of Teich(S), we will often omit the marking given by the
equivalence class of maps f : S → X and refer only to the target Riemann
surface X.
Fix an element X of Teich(S); we review the definition of the Teichmüller
and Weil-Petersson norms on the tangent and cotangent spaces TX Teich(S)
and TX∗ Teich(S). These definitions are made by identifying the cotangent
and tangent spaces with appropriate spaces of differentials on X, as we now
explain.
A holomorphic quadratic differential on X is a holomorphic tensor of type
(2, 0), which has a local representation of the form q(z)dz 2 . We define Q(X)
to be the vector space of holomorphic quadratic differentials φ on X having
at most simple poles at the punctures of X. The Riemann-Roch theorem
implies that Q(X) is a vector space of complex dimension 3g − 3 + n. Three
equivalent characterizations of the holomorphic quadratic differentials φ that
are in Q(X) are:
R 1/p
• kφkp = X ρ2−2p |φ|p < ∞ for some p ≥ 1;
R 2−2p p 1/p
• kφkp = X ρ |φ| < ∞ for all p ≥ 1;
−2
• kφk∞ = ess supX ρ |φ| < ∞.
Moreover, since Q(X) is finite dimensional, all of these norms are equivalent,
but there is no uniform comparison between these norms that holds for all

1We remark that every equivalence class [f : S → X] can itself be represented by a


quasiconformal homeomorphism f : S → X of minimal dilatation.
12 K. BURNS, H. MASUR AND A. WILKINSON

X.2 This means the appropriate norm to consider on Q(X) may depend on
the context. We will be particularly interested in the cases p = 1, which will
arise in defining the Teichmüller norm, p = 2, which will arise in defining
the WP norm and p = ∞, which will arise in the statement of Nehari’s
theorem. The space Q(X) is naturally paired with the space of essentially
bounded Beltrami differentials, which we next define.
A Beltrami differential on X is a measurable tensor of type (−1, 1), which
has a local representation of the form b(z)dz/dz. Note that the product of
a Beltrami differential with a quadratic differential is a (1, 1)-tensor. Let
M (X) be the vector spaceR of all measurable Beltrami differentials µ on X
with the property that X |φµ| < ∞, for every φ ∈ Q(X). We then have a
natural complex pairing of the space M (X) with Q(X) given by
Z
hφ, µi = φµ for φ ∈ Q(X), µ ∈ M (X).
X
In view of the fact that elements of Q(X) have finite Lp norm for every
1 ≤ p ≤ ∞, it follows that elements of M (X) are precisely those Beltrami
differentials µ on X of finite Lq norm, for 1 ≤ q ≤ ∞.
The Measurable Riemann Mapping Theorem states that any for measur-
able Beltrami differential µ ∈ M (X) with kµk∞ < 1, there exist a Riemann
surface Xµ and a quasiconformal homeomorphism hµ : X → Xµ such that,
with µ expressed in local coordinates as b(z)dz/dz, we have:
∂hµ /∂z
= b(z).
∂hµ /∂z
The map hµ is called a solution to the Beltrami equation and is unique up
to composition with a conformal isomorphism. Moreover there is analytic
dependence on parameters, so that if f : S → X is a quasiconformal marking
of X in Teich(S), and kµk∞ = 1, then t 7→ (htµ ◦ f : S → Xtµ ) describes
an analytic path in Teich(S) through X at t = 0. One can show that two
Beltrami differentials µ, ν ∈ M (X) of unit L∞ norm describe paths that are
tangent in Teich(S) at t = 0 if and only if hµ, φi = hν, φi, for every φ ∈ Q(X).
This gives rise to the fundamental isomorphisms of vector spaces:
TX Teich(S) ∼
= M (X)/Q(X)⊥ and TX∗ Teich(S) ∼
= Q(X),
where Q(X)⊥ = {µ ∈ M (X) : hµ, φi = 0, ∀φ ∈ Q(X)}.
Having described these identifications, we now can define the Teichmüller
and WP norms. The Teichmüller distance dT on Teich(S) described above is
the path metric for a Finsler structure k·kT whose restriction to TX∗ Teich(S)
is the L1 norm on Q(X):
Z
kφkT = kφk1 = |φ|.
X

2We remark that if X is a surface with infinite hyperbolic area, then Q(X) is not finite
dimensional, and these characterizations are not equivalent.
THE WEIL-PETERSSON GEODESIC FLOW 13

The Teichmüller norm kφkT encodes information about variations of the


conformal structure X induced by the covector φ. The Weil-Petersson met-
ric is defined by the L2 norm:
Z 1/2
−2 2
kφkW P = kφk2 = ρ |φ| .
X
Note that the definition of the WP metric involves both conformal and
hyperbolic data from X; this feature makes the WP metric somewhat tricky
to work with. On the other hand, the hyperbolic input from the metric ρ
leads to the delicate and beautiful connections between the WP metric and
hyperbolic geometry and dynamics discussed in the introduction.
The Teichmüller and WP norms on the tangent space TX Teich(S) are
then induced by this pairing via the formulae:
kvkT = sup Re(hφ, µi); kvkW P = sup Re(hφ, µi),
φ∈Q(X), kφkT =1 φ∈Q(X), kφkW P =1

for any µ ∈ M (X) representing the tangent vector v ∈ TX Teich(S). By


duality of the appropriate Lp norms, it follows that the Teichmüller norm is
induced by the L∞ norm on M (X) and the WP norm is induced by the L2
norm: Z 1/2
2 2
kvkT = kµk∞ ; kvkW P = kµk2 = ρ |µ| ,
X
for any µ ∈ M (X) representing v ∈ TX Teich(S). The Teichmüller and WP
norms are both invariant under the derivative action of MCG(S).
A simple comparison between the WP and Teichmüller norms follows from
the Cauchy-Schwarz and Gauss-Bonnet Theorems (see Prop 2.4 in [21]):
kvkW P ≤ |2πχ(S)|1/2 kvkT ,
for any X ∈ Teich(S) and v ∈ TX Teich(S). We prove in Section 5.1 a
complementary bound: there exists C > 0 such that
kvkW P ≥ Cℓ(X)−1/2 kvkT ,
for any v ∈ TX Teich(S), where ℓ(X) denotes the hyperbolic length of the
shortest closed geodesic on X, also known as the systole of X. This lower
bound will be used to estimate from above the derivatives of the WP metric.
1.3. The bundle of projective structures on S. A projective struc-
ture on a surface X is an atlas of charts into C whose overlaps are Möbius
transformations (elements of P SL(2, C)); note that a projective structure
determines a unique complex structure. Fix as above a Riemann surface
S of genus g with n punctures. A marked projective structure is a homeo-
morphism f : S → X, where X is endowed with a projective structure; we
say that two marked structures f1 : S → X1 and f2 : S → X2 are equivalent
if there is a projective isomorphism from X1 to X2 homotopic to f2 f1−1 .
Denote by Proj(S) the space of equivalence classes of projective structures
marked by S.
14 K. BURNS, H. MASUR AND A. WILKINSON

It is a classical fact that Proj(S) has the structure of a complex manifold


that arises from its embedding into the representation variety of homomor-
phisms from π1 (S) into P SL(2, C), modulo conjugacy (see [14]). The map
that assigns to each marked projective structure the compatible marked
conformal structure defines a fibration π : Proj(S) → Teich(S). The fiber
ProjX (S) over X is an affine space modelled on Q(X). That is, given
X0 ∈ ProjX (S) and φ ∈ Q(X) there is a unique X1 ∈ ProjX (S) and a
conformal map f : X0 → X1 respecting markings such that Sf = φ. Here
Sf is the Schwarzian derivative
 ′′ ′  
f (z) 1 f ′′ (z) 2 2
Sf (z) = − dz .
f ′ (z) 2 f ′ (z)
In particular, there is a well-defined difference β1 − β2 ∈ Q(X), for β1 , β2 ∈
ProjX (S) that defines a holomorphic map from ProjX (S) × ProjX (S) to
Q(X).
In the next subsection we describe how to define a complex structure
on Teich(S) that gives Proj(S) the structure of a holomorphic fiber bun-
dle, and a holomorphic identification of Proj(S) with the cotangent bundle
T ∗ Teich(S). We shall also describe how to construct a holomorphic sec-
tion σQF of this bundle that arises from a process called quasifuchsian uni-
formization. For now we note that the Fuchsian uniformization X ∼ = H/Γ
defines a section σF : Teich(S) → Proj(S). This section is real analytic but
not holomorphic.

1.4. Quasifuchsian space. Let S = H/Γ be a hyperbolic Riemann surface


with Γ < P SL(2, R), and denote by S the hyperbolic Riemann surface
L/Γ, where L is the lower half plane. Since Γ is a Fuchsian group, it acts
on the Riemann sphere Ĉ fixing H, L and the real axis/circle at infinity
R∞ = Ĉ \ (H ∪ L). Following McMullen [21], we define quasifuchsian space
QF (S) to be the product:
QF (S) = Teich(S) × Teich(S).
Then QF (S) parametrizes marked quasifuchsian groups equivalent to
Γ(S); more precisely we have:
Theorem 1.1 (Bers Simultaneous Uniformization). For each pair ([f : S →
X], [g : S → Y ]) ∈ QF (S), there exist a quasiconformal homeomorphism
φ : Ĉ → Ĉ, a Kleinian group Γ(X, Y ) < P SL(2, C), and a homomorphism
h : Γ → Γ(X, Y ) (all unique up to the action of P SL(2, C)) with the following
properties:
(1) for each γ ∈ Γ, we have φ ◦ γ = h(γ) ◦ φ;
(2) φ sends (H ∪ L, R∞ ) to (Ω(X, Y ), Λ(X, Y )), where Λ(X, Y ) ⊂ Ĉ is
a quasicircle;
(3) there is a conformal isomorphism Ω(X, Y )/Γ(X, Y ) ∼ = (X ∪ Y ) mak-
ing the following diagram commute:
THE WEIL-PETERSSON GEODESIC FLOW 15

φ
(H ∪ L) // Ω(X, Y )

mod Γ mod Γ(X,Y )


 f ∪g 
(S ∪ S) // (X ∪ Y )

We define the Fuchsian locus F (S) to be the image of Teich(S) under


the antidiagonal embedding α̂(X) = (X, X) ∈ QF (S). Theorem 1.1 gives a
natural “quasifuchsian uniformization” map
σ : Teich(S) × Teich(S) → Proj(S) × Proj(S)
that sends (X, Y ) to the projective structures on X and Y inherited from
Ω(X, Y ). We write:
σ(X, Y ) = (σQF (X, Y ), σ QF (X, Y )).
Observe that σQF (X, X) = σF (X). The complex structure on Teich(S)
is then defined via the Bers embedding: fixing X ∈ T eich(S), we define
βX : Teich(S) → Q(X) by
βX (Y ) = σQF (X, Y ) − σF (X).
The map βX is an embedding, and the pullback of the complex structure
on Q(X) gives a complex structure on Teich(S) that is independent of X
(that is, two different Xs give isomorphic structures). Recall that Q(X) is
a Banach space when endowed with any Lp norm. A fundamental result of
Nehari implies that the image of Teich(S) under βX is a bounded subset of
Q(X) in the L∞ norm, where the bound is independent of X. This fact is
used in a crucial way in McMullen’s paper and in ours.
We have defined a complex structure on Teich(S), which induces a con-
jugate complex structure on Teich(S). The complex structure on QF (S) is
defined to be the product complex structure. The Fuchsian locus F (S) is
then a totally real submanifold of QF (S). It can be checked that the fibra-
tion Proj(S) → Teich(S) is holomorphic with respect to these structures.
Furthermore, the naturally-defined section σ : QF (S) → Proj(S) × Proj(S)
is holomorphic. Hence, for a fixed Y ∈ Teich(S), the map X 7→ σQF (X, Y )
gives a holomorphic section of Proj(S) over Teich(S); this section gives an
isomorphism between Proj(S) and the cotangent bundle T ∗ Teich(S).
We will use the section σ in a crucial way to estimate higher derivatives
of the W P metric. Some important properties of σ, which we record at this
point are:
• σ is holomorphic;
• σQF (X, X) = σF (X); where σF is the Fuchsian uniformization sec-
tion defined above;
• for any Y, Z ∈ Teich(S), the map X 7→ σQF (X, Y ) − σQF (X, Z)
defines a holomorphic 1-form on Teich(S);
16 K. BURNS, H. MASUR AND A. WILKINSON

• there exists C ≥ 1 such that for any X ∈ Teich(S) and Y, Z ∈


Teich(S), we have
C −1 ≤ kσQF (X, Y ) − σQF (X, Z)kT ≤ C,
where k · kT is the Teichmüller norm on TX∗ Teich(S).
The final item on this list follows from Nehari’s bound: see Theorems 2.2 in
[21].
The main result from McMullen we will use relates the Bers embedding
to the Kähler form ωW P of the WP metric.
Theorem 1.2 ([21], Theorem 7.1). Fix Z ∈ Teich(S), and let θW P be the
(1, 0)-form on Teich(S) given by
θW P (X) = σF (X) − σQF (X, Z)
= −βX (Z).
Then θW P is a primitive for the Weil-Petersson Kähler form; that is,
d(iθW P ) = ωW P .
1.5. Fenchel-Nielsen coordinates. Continue to denote by S a marking
Riemann surface of genus g with n punctures. We define here a natural
system of global coordinates on Teich(S), called Fenchel-Nielsen coordinates,
in which the Kähler form ωW P takes a simple form.
Recall that a curve in S is nonperipheral if it is not homotopic to a loop
surrounding a single puncture. A pants decomposition of S is a collection
P of 3g − 3 + n pairwise disjoint, homotopically nontrivial, nonperipheral
and homotopically distinct simple closed curves. The complement of these
curves is a collection of surfaces called pairs of pants. Topologically, a pair
of pants is a three-times punctured sphere. A pair of pants has one of
three types of conformal structure depending on whether each puncture is
locally modelled on the punctured plane or on the complement of a disk, in
which case we say that the boundary component is a circle. A pair of pants
with j boundary circles has a j-dimensional space of hyperbolic structures,
parametrized by the hyperbolic lengths of the boundary circles.
We introduce notation that will be used throughout the paper. If f : S →
X is a marked Riemann surface and α is a homotopically nontrivial, nonpe-
ripheral, simple closed curve in S, we denote by ℓα (X) the hyperbolic length
in X of the unique geodesic in the homotopy class of f∗ [α]. This geodesic
length function is intimately connected with the WP metric and is used to
define Fenchel-Nielsen coordinates.
Fix a pair of pants decomposition P = {α1 , . . . , α3g−3+n } of S. The
Fenchel-Nielsen coordinates
(ℓα , τα )α∈P : Teich(S) → (R>0 × R)3g−3+n
determined by P are defined as follows. For f : S → X a marked Riemann
surface and α ∈ σ, we define ℓα (X) to be the geodesic length as above and
τα (X) to be the twist parameter, which records the relative displacement in
THE WEIL-PETERSSON GEODESIC FLOW 17

how the pairs of pants are glued together along α to obtain the hyperbolic
metric on X; more precisely, a full Dehn twist about the curve α changes
τα by the amount ℓα . One must adopt a convention for how this relative
displacement τ is measured, as it is intrinsically only well-defined up to a
constant, but this does not introduce any serious issues. These give global
coordinates on Teich(S).
The Fenchel-Nielsen coordinates are natural with respect to the WP met-
ric. Wolpert
P [34] proved that for any pants decomposition P , we have
ωW P = 21 α∈P dℓα ∧ dτα . A component of this formula is the important
fact that the vector field ∂/∂τα , which generates the Dehn twist flow about
α, is the symplectic gradient of the Hamiltonian function 21 ℓα :
1 ∂
dℓα = ωW P ( , );
2 ∂τα
put another way:

grad ℓα = −2J .
∂τα
This fundamental relationship is the starting point for many of Wolpert’s
deep asymptotic expansions for the WP metric, which we discuss in more
detail in Section 3.

1.6. The Deligne-Mumford compactification of moduli space. As


mentioned earlier, Teich(S) is incomplete with respect to the WP distance
[33]. In particular, it is possible to shrink a simple closed curve α to a point
and leave Teichmüller space along a WP geodesic in finite time — indeed,
1/2
the time it takes is on the order of ℓα . It turns out that pinching one
or more closed geodesics to a point is the only way to leave every compact
set of Teich(S) in finite time along a WP geodesic; this fact allows us to
define a completion of Teich(S) called augmented Teichmüller space and
denoted Teich(S). The mapping class group MCG(S) acts on Teich(S) and
the quotient M(S) is the Deligne-Mumford compactification of the moduli
space M(S).
Augmented Teichmüller space Teich(S) is obtained by adjoining lower-
dimensional Teichmüller spaces of noded Riemann surfaces, which gives it
the structure of a stratified space. The combinatorics of this stratification
are encoded by a symplicial complex C(S) called the curve complex. We
review this construction here.
We first define the curve complex C(S), which is a 3g − 4 + n dimensional
simplicial complex. The vertices of C(S) are homotopy classes of homo-
topically nontrivial, nonperipheral simple closed curves on S. We join two
vertices by an edge if the corresponding pair of curves has disjoint repre-
sentatives. More generally, a k simplex σ ∈ C(S) consists of k + 1 distinct
vertices that have disjoint representatives. We note that in the sporadic
cases of the punctured torus (g, n) = (1, 1) and 4-times punctured sphere
(g, n) = (0, 4), C(S) is just an infinite discrete set of vertices, since there do
18 K. BURNS, H. MASUR AND A. WILKINSON

not exist disjoint homotopically distinct curves on the underlying surface S.


Except in these sporadic cases, C(S) is a connected locally infinite complex.3
Note that a maximal simplex in C(S) defines a pants decomposition of S.
The mapping class group MCG(S) acts on C(S).
Masur gave an interpretation of points in the completion of Teich(S) as
marked noded Riemann surfaces [23]. A noded Riemann surface is a complex
space with at most isolated singularities, called nodes, each possessing a
neighborhood biholomorphic to a neighborhood of (0, 0) in the curve
{(z, w) ∈ C2 : zw = 0}.
Removing the nodes of a noded Riemann surface Y yields a (possibly dis-
connected) Riemann surface which we will usually denote by Ŷ . The com-
ponents of Ŷ are called the pieces of Y .
Given a simplex σ ∈ C(S), a marked noded Riemann surface with nodes
corresponding to σ is a noded Riemann surface Xσ equipped with a con-
tinuous mapping f : S → Xσ so that f |S\σ is a homeomorphism to X̂σ .
Two marked noded Riemann surfaces [f1 : S → Xσ1 ] and [f2 : S → Xσ2 ] are
equivalent if there is a map h : Xσ1 → Xσ2 that preserves nodes, is isotopic
to f2 ◦f1−1 rel the nodes, and induces a biholomorphic map from X̂σ1 to X̂σ2 .
We then define Teich(S)4 to be the set of all noded and unnoded Riemann
surfaces marked by S, modulo this equivalence, where the nodes σ range
over all simplices of C(S). Each element of Teich(S) has a representative
[f : S → Xσ ], where either σ ∈ C(S) or σ = ∅, and we adopt the convention
that X∅ is an unnoded Rieman surface. We will usually denote X∅ simply
by X.
Notational convention. As with the elements of Teich(S), we will fre-
quently abuse notation and omit the marking when referring to an element
of Teich(S); hence “let Xσ ∈ Teich(S)” is shorthand for “let [f : S → Xσ ]
be an element of Teich(S).”
To describe a neighborhood of a point [f : S → Xσ ] in Teich(S), we give
coordinates adapted to the simplex σ. For any such σ, let P be maximal
simplex in C(S) (pants decomposition) containing σ, and let (ℓα , τα )α∈P
be the corresponding Fenchel-Nielsen coordinates on Teich(S). Then the
extended Fenchel-Nielsen coordinates for P are obtained by allowing lengths
ℓα to range in R≥0 and taking the quotient by identifying (0, t) with (0, t′ )
in each R factor corresponding to the curves in σ.
This also defines a topology on Teich(S). We note that the space is not
locally compact. A neighborhood of a noded surface allows for the twists τα
corresponding to the curves α ∈ σ to be arbitrary real numbers.
3In the sporadic cases there is more than one possible definition of C(S); in another, very
standard definition in these cases, one adds edges joining curves that intersect minimally
(once in the case of the torus and twice in the case of the sphere). The resulting 1-complex
is the Farey graph in both cases.
4Not to be confused with Teich(S), which was introduced in §1.4.
THE WEIL-PETERSSON GEODESIC FLOW 19

We introduce here a shorthand notation that will be used in various parts


of the paper. If the topological type of the surface S is fixed, we will some-
times write T for Teich(S). In this notation, T will then denote the aug-
mented space Teich(S). For each σ ∈ C(S), we denote by Tσ the set of
equivalence classes [f : S → Xσ ] with nodes at σ. Then
[
T =T ∪ Tσ ,
σ∈C(S)

and we refer to Tσ as a stratum of T . The main stratum, T = T∅ , is


simply the full Teichmüller space T . Each boundary stratum Tσ is naturally
identified with the product of the Teichmüller spaces of the pieces of any
element Xσ ; equivalently, Tσ is the Teichmüller space of the Riemann surface
X̂σ .

2. Background on the geodesic flow


Let M be a Riemannian manifold. As usual hv, wi denotes the inner
product of two vectors and ∇ is the Levi-Civita connection defined by the
Riemannian metric. The covariant derivative along a curve t 7→ c(t) in M
D
is denoted by Dc , dt or simply ′ if it is not necessary to specify the curve; if
V (t) is a vector field along c that extends to a vector field Vb on M , we have

V ′ (t) = ∇ċ(t) Vb .
D
Given a smooth map (s, t) 7→ α(s, t), we let ∂s denote covariant differentia-
D
tion along a curve of the form s 7→ α(s, t) for a fixed t. Similarly ∂t denotes
covariant differentiation along a curve of the form t 7→ α(s, t) for a fixed s.
The symmetry of the Levi-Civita connection means that
D ∂α D ∂α
(s, t) = (s, t)
∂s ∂t ∂t ∂s
for all s and t.
The curve c is a geodesic if it satisfies the equation Dc ċ(t) = 0. Since
this equation is a first order ODE, a geodesic is uniquely determined by
its initial tangent vector. Geodesics have constant speed, since we have
d
hċ(t), ċ(t)i = 2hċ′ (t), ċ(t)i = 0 if c is a geodesic.
dt
The Riemannian curvature tensor R is defined by
R(A, B)C = (∇A ∇B − ∇B ∇A − ∇[A,B] )C.
The sectional curvature of the 2-plane spanned by vectors A and B is defined
by
hR(A, B)B, Ai
K(A, B) = .
kA ∧ Bk2
20 K. BURNS, H. MASUR AND A. WILKINSON

2.1. Vertical and horizontal subspaces and the Sasaki metric. The
tangent bundle T T M to T M may be viewed as a bundle over M in three
natural ways shown in the following commutative diagram:
DπM
T T MK // T M
KK
KK
KπKT M ◦πM
κ KK πM
KK
KK
 πM K%% 
TM // M

The first is via the composition of the natural bundle projections πT M :


T T M → T M and πM : T M → M . The second is via the composition of
the derivative map DπM : T T M → T M with πM . The third involves a map
κ : T T M → T M , often called the connector map, that is determined by the
Levi-Civita connection. If ξ ∈ T T M is tangent at t = 0 to a curve t 7→ V (t)
in T M c(t) = πM (V (t)) is the curve of footpoints of the vectors V (t), then
κ(ξ) = Dc V (0).
The subbundle ker(DπM ) is the vertical subbundle. It is naturally iden-
tified with T M via the map πT M . The subbundle ker(κ) is the horizontal
subbundle. It is naturally identified with T M via the map κ and is trans-
verse to the vertical subbundle. If v ∈ Tp M , we may identify Tv T M with
Tp M × Tp M via the map DπM × κ : T T M → T M × T M .
Each element of Tv T M can thus be represented uniquely by a pair (v1 , v2 )
with v1 ∈ Tp M and v2 ∈ Tp M . Put another way, every element ξ of Tv T M
is tangent to a curve V : (−1, 1) → T M with V (0) = v. Let c = πM ◦
V : (−1, 1) → M be the curve of basepoints of V in M . Then ξ is represented
by the pair
(ċ(0), Dc V (0)) ∈ Tp M × Tp M.
These coordinates on the fibers of T T M restrict to coordinates on T T 1 M .
Regarding T T M as a bundle over M in this way gives rise to a natural
Riemannian metric on T M , called the Sasaki metric. In this metric, the
inner product of two elements (v1 , w1 ) and (v2 , w2 ) of Tv T M is defined:
h(v1 , w1 ), (v2 , w2 )i = hv1 , v2 i + hw1 , w2 i.
This metric is induced by a symplectic form ω on T T M ; for vectors (v1 , w1 )
and (v2 , w2 ) in Tv T M , we have:
ω((v1 , w1 ), (v2 , w2 )) = hv1 , w2 i − hw1 , v2 i.
This symplectic form is the pull back of the canonical symplectic form on the
cotangent bundle T ∗ M by the map from T M to T ∗ M induced by identifying
a vector v ∈ Tp M with the linear function hv, ·i on Tp M .
Sasaki [31] showed that the fibers of the tangent bundle are totally geo-
desic submanifolds of T T M with the Sasaki metric. A parallel vector field
along a geodesic of M (viewed as a curve in T M ) is a geodesic of the Sasaki
metric. Such a geodesic is orthogonal to the fibers of T M . If v ∈ Tp M and
THE WEIL-PETERSSON GEODESIC FLOW 21

v ′ ∈ Tp′ M , we can join them by first parallel translating v along a geodesic


from p to p′ to obtain w ∈ Tp′ M and then moving from w to v ′ along a line
in Tp′ M . If v ′ is close to v, we can choose the geodesic so that its length
is d(p, p′ ). Then dSas (v, v ′ ) is the length of the hypotenuse of a small right
angled triangle whose other two sides have lengths d(p, p′ ) and kw − v ′ k.
Topogonov’s comparison theorem [8, Theorem 2.2] tells us that if −k2 is a
lower bound for the curvature of the Sasaki metric in a neighborhood of v,
then dSas (v, v ′ ) is less than the length of the hypotenuse of a right triangle
in a space form of constant curvature −k2 with sides of length d(p, p′ ) and
kw − v ′ k meeting at the right angle. It follows easily that
dSas (v, v ′ ) ≍ d(p, p′ ) + kw − v ′ k,
as v ′ → v. The notation a ≍ b, here and in the rest of the paper, means
that the ratios a/b and b/a are bounded above by a constant. In this case
the constant is 2.

2.2. The geodesic flow and and Jacobi fields. For v ∈ T M let γv
denote the unique geodesic γv such that γ̇v (0) = v. The geodesic flow
φt : T M → T M is defined by
φt (v) = γ̇v (t),
wherever this is well-defined. If M is an open Riemannian manifold, the geo-
desic flow is always defined locally. Since the geodesic flow is a Hamiltonian
flow, it preserves a natural volume form on T 1 M called the Liouville volume
form. When the integral of this form is finite, it induces a unique probablity
measure on T 1 M called the Liouville measure or Liouville volume.
Consider now a one-parameter family of geodesics, that is a map α :
(−1, 1)2 → M with the property that α(s, ·) is a geodesic for each s ∈
(−1, 1). Denote by J(t) the vector field
∂α
(0, t)
J(t) =
∂s
along the geodesic γ(t) = α(0, t). Then J satisfies the Jacobi equation:
J ′′ + R(J, γ̇)γ̇ = 0,
in which ′ denotes covariant differentiation along γ. Since this is a second or-
der linear ODE, the pair of vectors (J(0), J ′ (0)) ∈ Tγ(0) M ×Tγ(0) M uniquely
determines the vectors J(t) and J ′ (t) along γ(t). A vector field J along a
geodesic γ satisfying the Jacobi equation is called a Jacobi field.
The pair (J(0), J ′ (0)) corresponds in the manner described above to the
tangent vector at s = 0 to the curve V (s) = ∂α ∂t (s, 0). To see this, note that
V (s) is a vector field along the curve c(s) = α(s, 0), so V ′ (0) corresponds to
the pair
∂α D ∂α D ∂α
(ċ(0), Dc (s, 0)) = (J(0), (s, 0)) = (J(0), (s, 0)) = (J(0), J ′ (0)).
∂t ∂s ∂t ∂t ∂s
22 K. BURNS, H. MASUR AND A. WILKINSON

In the same way one sees that (J(t), J ′ (t)) corresponds to the tangent vector
at s = 0 to the curve s 7→ ∂α ′
∂t (s, t) = φt ◦ V (s), which is Dφt (V (0)).
To summarize the preceding discussion, there is a one-one correspondence
between elements of Tv T M and Jacobi fields along the geodesic γ with
γ̇(0) = v.
Note that the pair (J(t), J ′ (t)) defines a section of T T M over γ(t). We
have the following key proposition:
Proposition 2.1. The image of the tangent vector (v1 , v2 ) ∈ Tv T M under
the derivative of the geodesic flow Dv ϕt is the tangent vector (J(t), J ′ (t)) ∈
Tϕt (v) T M , where J is the unique Jacobi field along γ satisfying J(0) = v1
and J ′ (0) = v2 .
2.3. The Jacobi and Riccati equations. Choose an orthonormal basis
e1 = γ̇(0), . . . , en at 0 for the tangent space at γ(0) and parallel transport
the basis along
P γ(t). Any Jacobi field can be written in terms of the basis
as J(t) = y k ek (t) and so the Jacobi equation can be written as
d2 y k X
= y j (t)hR(ej (t), e1 (t))e1 (t), ek (t)i.
dt2
j

A choice of initial condition y k (0), y k (0) determines a solution.
Let J (t) denote any matrix of solutions to the Jacobi equation. Define
the Wronskian of J to be the matrix
W (J , J ) = (J ∗ )′ J − J ∗ J ′ ,
where ∗ denotes transpose. Note that by the Jacobi equation,
(1) W ′ (J , J ) = (J ∗ )′′ J − J ∗ J ′′ = (RJ )∗ J − J ∗ RJ = 0,
since R is symmetric.
Let
U = J ′ J −1 .
It satisfies the matrix Riccati equation:
U ′ = −hR(· γ̇)γ̇, · i − U 2 .
Lemma 2.2. The operator U = J ′ J −1 is symmetric if and only if
W (J (t), J (t)) = 0,
and this is equivalent to the statement that for any two columns Ji , Jj of J ,
we have
ω((Ji , Ji′ ), (Jj , Jj′ )) = 0,
where ω is the symplectic form on T 1 M .
Proof. The first statement follows from (1) and the second follows from the
definition of ω. ⋄
THE WEIL-PETERSSON GEODESIC FLOW 23

2.4. Perpendicular Jacobi fields. There are two natural subbundles of


T T M that are invariant under the derivative Dϕt of the geodesic flow, the
first containing the second. The first is the tangent bundle T T 1 M to the unit
tangent bundle of M . Under the natural identification Tv T M ∼ = (Tx M )2 ,
1 1
for v ∈ Tx M , the subspace Tv T M is the set of all pairs (w0 , w1 ) such
that hv, w1 i = 0. To see this, note that if α(s, t) is a variation of geodesics
d
generating the Jacobi field J along γ, and k dt α(s, t)k2 = 1 for all s, t, then
2
d
d α d2 d
0= = 2h α, αi = 2hJ ′ (0), γ̇(0)i.
ds dt (0,0) ds dt dt (0,0)
The Dϕt -invariance of T T 1 M follows from the ϕt -invariance of T 1 M .
The second natural invariant subbundle is the orthogonal complement
ϕ̇⊥ in T T 1 M to the vector field ϕ̇ generating the geodesic flow. Under the
natural identification Tv T M ∼ = (Tx M )2 , for v ∈ Tx1 M , the subspace ϕ̇⊥ (v)
is the set of all pairs (w0 , w1 ) such that hv, w0 i = hv, w1 i = 0. To see this,
note that in these coordinates, ϕ̇(v) = (v, 0), which generates the tangent
Jacobi field J(t) = γ̇(t) along the geodesic with γ̇(0) = v.
To see that ϕ̇⊥ is Dϕt -invariant, observe that if hJ(0), γ̇(0)i = 0 and
hJ ′ (0), γ̇(0)i = 0, then the Jacobi equation implies that
(2) hJ(t), γ̇(t)i = 0 and hJ ′ (t), γ̇(t)i = 0, for all t.
A Jacobi field J along γ satisfying (2) is called a perpendicular Jacobi field.
To summarize, the space of all perpendicular Jacobi fields along γ corre-
sponds to the orthogonal complement to the direction of the geodesic flow
ϕ̇(v) at the point v = γ̇(0) ∈ T 1 M . To estimate the norm of the derivative
Dϕt on T T 1 M , it suffices to restrict attention to vectors in the invariant
subspace ϕ̇⊥ ; that is, it suffices to estimate the growth of perpendicular
Jacobi fields along geodesics.
2.5. Consequences of negative curvature. If the sectional curvatures of
the Riemannian metric are negative along γ, then it follows from the Jacobi
equation that hJ ′′ , Ji > 0, for any Jacobi field with the property that J(t)
and γ̇(t) are linearly independent. This has the following consequence.
Lemma 2.3. If the sectional curvatures are negative along γ, then the func-
tions kJ(t)k and kJ(t)k2 are strictly convex, for any nontrivial perpendicular
Jacobi field J along γ.
Proof. It follows from the Jacobi equation and negativity of the sectional
curvatures along γ that
 ′′
kJk2
= hJ, J ′ i′ = hJ ′ , J ′ i + hJ ′′ , Ji = hJ ′ , J ′ i − hR(J, γ̇)γ̇, Ji ≥ 0,
2
for any Jacobi field J. We have hJ ′ (t), J ′ (t)i > 0 unless J ′ (t) = 0; and, since
J is perpendicular, we have −hR(J, γ̇(t))γ̇(t), Ji > 0 unless J(t) = 0. But
J(t) and J ′ (t) cannot both vanish because J is nontrivial. Thus (kJk2 )′′ > 0,
which implies that kJ(t)k2 is a strictly convex function of t.
24 K. BURNS, H. MASUR AND A. WILKINSON

To show convexity of kJ(t)k, first note that if J(t0 ) = 0 we have


kJ(t)k
(3) lim = 1.
t→t0 |t − t0 |kJ ′ (t0 )k
When J 6= 0 we have
1 hJ ′ , Ji
(4) kJk′ = (hJ, Ji 2 )′ = 1
hJ, Ji 2
and
1 1
1 hJ, Ji 2 [hJ ′ , J ′ i + hJ ′′ , Ji] − hJ ′ , Ji2 /hJ, Ji 2
′′
(hJ, Ji ) =
2
hJ, Ji
hJ, JihJ , J i − hJ , Ji2 + hJ, JihJ ′′ , Ji
′ ′ ′
= 3
hJ, Ji 2
hJ, JihJ ′′ , Ji
≥ 3 ,
hJ, Ji 2
by the Cauchy Schwarz inequality. Thus kJk′′ ≥ hJ ′′ , Ji/kJk when J 6= 0,
and as we saw above, hJ ′′ , Ji = −hR(J, γ̇)γ̇, Ji > 0 in this case. ⋄

2.6. Unstable Jacobi fields. We continue to assume that the sectional


curvatures of M are negative. We have the following result from [11, Section
1.10].
Proposition 2.4. If the sectional curvatures of the Riemannian metric are
negative along the infinite geodesic ray γ : (−∞, a] → M , then there exists a
continuous family of positive definite, symmetric operators
{U+ (t) : Tγ(t) M → Tγ(t) M : t ∈ (−∞, a]}
such that for every t ≤ a and v ∈ Tγ(t) M , there exists a Jacobi field J(s)
defined for all s ≤ t and satisfying J(t) = v and J ′ (t) = U+ (t)v.
The family of operators U+ satisfies the Riccati equation:
U ′ = −hR(· γ̇)γ̇, · i − U 2 .
This means that for any vector v,
hv, U ′ (v)i = −hR(v, γ̇)γ̇, vi − hv, U (v)i2 .
A Jacobi field satisfying J ′ = U+ J is called an unstable Jacobi field.
2.7. Bounding the derivative of the geodesic flow. In the next propo-
sition we show how the norm of the derivative of the geodesic flow given by
Proposition 2.1 can be estimated from the growth of unstable Jacobi fields.
Proposition 2.5. Let γ : (−∞, ∞) → M be a geodesic in a manifold with
negative sectional curvatures. Any perpendicular Jacobi field J along γ sat-
isfies
k(J(1), J ′ (1))k ≤ (1 + m+ )(1 + u+ )2 k(J(0), J ′ (0))k,
THE WEIL-PETERSSON GEODESIC FLOW 25

where
′ (1))k
k(J+ (1), J+
m+ = max ′ (0))k .
k(J+ (0), J+
and
kJ+′ (0)k
u+ = max ,
kJ+ (0)k
the maxima being taken over all unstable Jacobi fields J+ .
We remark that the point of the lemma is that the constant in the bound
is essentially linear in m+ .
Proof. We begin by showing that if L is a perpendicular Jacobi field with
L(1) = 0, then
(5) kL′ (0)k ≥ kL(0)k ≥ kL′ (1)k.
These quantities are all 0 if L(0) = 0. If L(0) 6= 0, both inequalities are
strict as we now show. Since kLk is convex by Lemma 2.3 and decreases
from kL(0)k to 0 across the interval [0, 1], we have
−kLk′ (0) > kL(0)k > − lim kLk′ (t).
t→1−
Since L(1) = 0 it follows from (3) that
lim kLk′ (t) = − lim kL′ (t)k = −kL′ (1)k.
t→1− t→1−
This implies the right hand inequality in (5). From (4) and the Cauchy-
Schwarz inequality, we have that

kLk′ (0) ≤ kL′ (0)k.
This gives the left hand inequality in (5).
Now let L be a fundamental solution of the Jacobi equation with L(0) = I
and L(1) = 0. Thus for any Jacobi field L as above we have L(t) = L(t)L(0).
Let
U1 = L′ L−1 .
Then for any such Jacobi field L with kL(0)k = 1, the inequality
hU1 (0)L(0), L(0)i = hL′ (0), L(0)i = kL(0)kkLk′ (0) < −1
implies that all of the eigenvalues of U1 (0) are less than −1.
Now we wish to express a perpendicular Jacobi field (J, J ′ ) with k(J(0), J ′ (0))k =
1 as a sum of a perpendicular Jacobi field of the form (L, L′ ) as studied above
and an unstable perpendicular Jacobi field (J+ , J+ ′ ). Then k(L(0), L′ (0))k ≤

1 + k(J+ (0), J+ (0))k and we will obtain
k(J(1), J ′ (1))k ≤ k(J+ (1), J+

(1))k + k(L(1), L′ (1))k

≤ m+ k(J+ (0), J+ (0))k + k(L(0), L′ (0))k

≤ (1 + m+ )k(J+ (0), J+ (0))k + 1
≤ u+ (1 + m+ )kJ+ (0)k + 1.
26 K. BURNS, H. MASUR AND A. WILKINSON

Recall from Proposition 2.4 that J+ ′ = U J , where U is symmetric and


+ + +
positive definite. Hence U+ (0) − U1 (0) is symmetric and has all eigenvalues
greater than 1. Consequently [U+ (0) − U1 (0)]−1 is a contraction.
Let v = J(0), v ′ = J ′ (0) and w = [U+ (0) − U1 (0)]−1 (v ′ − U+ (0)v). Note
that
kwk ≤ kv ′ k + kU+ (0)vk ≤ 1 + u+ ,
since k(v, v ′ )k = 1. Now we obtain
(v, v ′ ) = (v, U+ (0)v) + (0, v ′ − U+ (0)v)
= (v, U+ (0)v) + (w − w, [U+ (0) − U1 (0)]w)
= (v + w, U+ (0)(v + w)) − (w, U1 (0)w).
This is the desired decomposition of (J, J ′ ) at t = 0. We have
kv + wk ≤ kvk + kwk ≤ 1 + (1 + u+ ) = 2 + u+ .
Putting everything together finally gives
k(J(1), J ′ (1))k ≤ u+ (1 + m+ )(2 + u+ ) + 1 ≤ (1 + m+ )(1 + u+ )2 .

Hence to estimate the norm of the derivative of the geodesic flow, it suffices
to estimate the growth of unstable Jacobi fields. This task is simplified by
considering the Riccati equation.
Lemma 2.6. For any unstable Jacobi field J we have
Z 1 
k(J(1), J ′ (1))kSas p
2
≤ 1 + kU+ (1)k exp kU+ (t)k dt .
k(J(0), J ′ (0))kSas 0

Proof. For any unstable Jacobi field J, we have J ′ = U+ J. Hence


p p
k(J(1), J ′ (1))kSas = J(1)2 + J ′ (1)2 = kJ(1)k 1 + kU+ (1)k2 .
Since kJk′ (t) ≤ kJ ′ (t)k by (4), we have
Z 1 ′  Z 1 
kJ(1)k kJ (t)k
≤ exp dt = exp kU+ (t)k dt .
kJ(0)k 0 kJ(t)k 0
Putting these last two inequalities together gives the desired estimate.

R1
To estimate the quantities 0 kU+ k(s)ds and kU+ (1)k, and hence the
derivative of the time 1 map of the geodesic flow, we have the following
theorem.
Theorem 2.7. Let M be a negatively curved manifold, and let γ : [−1, 1] →
M be a geodesic. Let P be an arbitrary parallel unit vector field along γ,
and let
uP = hU+ P, P i, and kP2 = −hR(P, γ̇)γ̇, P i.
Then
2
(1) u′P ≤ k − u2P , and
THE WEIL-PETERSSON GEODESIC FLOW 27

(2) for any 0 ≤ t ≤ 1,


p Z t 
kDφt (0)k ≤ (1 + u(0))2 (1 + 1 + u(t)2 ) exp u(s)ds ,
0

where k(t) = supv∈T 1 M −hR(v, γ̇(t)), γ̇(t), vi and u = supP uP .


γ(t)

Proof. Since U+ is positive definite, we have that uP > 0 and kU+ k ≤


supP uP . Since P is parallel and has unit length, we have

u′P = hU+′ P, P i = kP2 − hU+2 P, P i = kP2 − hU+ P, U+ P i

2 2
≤ k − hU+ P, P i2 = k − u2P .
This proves the first conclusion.
The second conclusion follows from Proposition 2.1 using the bounds given
by Proposition 2.5, Lemma 2.6 and the bound kU+ k ≤ u. ⋄

In the next section, we find bounds for solutions uP of the differential


2
inequality u′P ≤ k − u2P for the WP metric.

3. Bounds on the derivative of ϕ1 in the WP metric


The WP metric is negatively curved, so the analysis in the previous section
applies in this setting. In this section, we fix a Riemann surface S and denote
by T the Teichmüller space Teich(S), by T the augmented Teichmüller space
Teich(S), and by ∂T the boundary T \T . We will omit the dependence on S
in the mapping class group MCG = MCG(S) and curve complex C = C(S).
For σ ∈ C, the boundary stratum corresponding to the noded surface Sσ is
denoted by Tσ . We denote by π : T T → T the natural projection and by ϕt
the geodesic flow on T 1 T .
For each unit WP tangent vector v ∈ T 1 T and t ≥ 0, we denote by ρt (v)
the minimum WP distance from the geodesic segment π(ϕ[−t,t] (v)) in T to
the singular locus ∂T . The main result of this section is:

Theorem 3.1. There exist constants β > 0, ǫ0 > 0 and C > 1 with the
following property. For every t ∈ [0, 1] and v ∈ T 1 T with ρt (v) ∈ (0, ǫ0 ), we
have:
kDv ϕt kW P ≤ C(ρt (v))−β .

Since we will be dealing only with the WP metric in this section and the
next, we will omit the subscript “WP” from the notation for inner product,
norm and distance functions. Since Section 5 is largely concerned with
comparisons between the WP and Teichmüller metric, the subscripts used
in notation will return at that point.
28 K. BURNS, H. MASUR AND A. WILKINSON

3.1. Combined length bases. The next two sections summarize work of
Wolpert in a constellation of papers [35, 36, 37, 38]. If χ is an arbitrary
finite collection of vertices in C and X ∈ T we define:
ℓχ (X) = min ℓβ (X), and ℓχ (X) = max ℓβ (X).
β∈χ β∈χ

For X ∈ T , we continue to denote by ℓ(X) the systole of X, which is the


hyperbolic length of the shortest closed geodesic in X. By a “simple closed
curve in S” we will always mean a vertex of C; that is, the homotopy class
of a nonperipheral, homotopically nontrivial, simple closed curve in S. Two
curves are disjoint if they are connected by an edge in C. Let B be the set
of pairs (σ, χ), where σ ∈ C and χ is a collection of simple closed curves
in S such that each β ∈ χ is disjoint from every α ∈ σ (we allow for the
possibility that χ = ∅).
For each simple closed curve α in S, the root length function
ℓ1/2
α : T → R>0
plays an important role in various asymptotic expansions of the WP metric.
1/2
Wolpert proved that the functions ℓα and ℓα are convex along WP geodesics
in T (see Corollary 3.4 and Example 3.5 of [38] and Corollary 8.2 of [39]).
1/2
The WP gradient of ℓα defines a vector field
λα = grad ℓ1/2
α .
Following Wolpert, we say that (σ, χ) ∈ B is a combined (short and relative)
length basis at X ∈ T if the collection
{λα (X), Jλα (X), grad ℓβ (X)}α∈σ,β∈χ
is a basis for TX T .
For each η > 0, let
U (η) = {X ∈ T | ℓ(X) < η},
which is a deleted open neighborhood of ∂T in T .
Proposition 3.2. There exist constants c > 1, η, δ > 0 and a countable
collection U of open sets in T with the following properties.
(1) For each U ∈ U , there exists a combined length basis (σ, χ) ∈ B such
that, for every X ∈ U :
1/c < ℓχ (X) ≤ ℓχ (X) < c.
(2) For each X ∈ U (η), there exists U ∈ U such that for any Y ∈ T ,
d(X, Y ) < δ =⇒ Y ∈ U ;
in particular, the sets in U cover U (η).
Before proving this proposition, we discuss further the properties of the
WP metric in a neighborhood of the boundary strata of T . Let σ ∈ C
be a simplex, and consider a marked noded Riemann surface f : S → Xσ
THE WEIL-PETERSSON GEODESIC FLOW 29

representing an element of the boundary stratum Tσ . Recall that the hy-


perbolic surface X̂σ is obtained from Xσ by deleting its nodes. If β is a
simple closed curve in S that is disjoint from the curves in σ, then f∗ [β] is
uniquely represented as a closed geodesic on X̂σ . In this way, the definition
of ℓβ extends continuously to the boundary stratum Tσ ; for such β, we define
ℓβ ([f : S → Xσ ]) to be the hyperbolic length of the geodesic representative
of f∗ [β] on X̂σ . For Xσ ∈ Tσ , we can also define a relative systole ℓ(Xσ ) to
be the infimum of ℓβ (X̂σ ), taken over all curves β disjoint from the curves
in σ.
Recall that the boundary stratum Tσ is isomorphic to the product of
Teichmüller spaces. In particular, Tσ itself carries a WP metric, which is
the product of the WP metrics on the pieces of Xσ , for any Xσ ∈ Tσ . Let
Xσ be a marked noded surface in Tσ . We say that χ is a relative length
basis at Xσ if (σ, χ) ∈ B and the functions {ℓβ }β∈χ give local coordinates
for Tσ at Xσ . Equivalently, χ is a relative length basis at Xσ if the vectors
{grad ℓβ (Xσ )}β∈χ in the induced W P metric on Tσ span the tangent space
TXσ Tσ .
Proposition 3.3 (Existence of relative length bases). For each σ ∈ C, and
each marked noded Riemann surface Xσ ∈ Tσ , there exists (σ, χ) ∈ B such
that χ is a relative length basis at Xσ .
We remark that, unlike Fenchel-Nielsen coordinates, the local coordinates
{ℓβ }β∈χ never extend to a global coordinate system on Tσ ; the reason is that
there are points in T where the geodesic representatives of the curves in χ
cross each other orthogonally. At these points, the coordinate system hits a
singularity. Proposition 3.3 ensures, however, that if one works locally these
issues can be ignored. Wolpert proves:
Theorem 3.4 ([38], Corollary 4.5.). The WP metric is comparable to a sum
of differentials of geodesic-length functions for a simplex σ of short geodesics
and corresponding relative length basis χ as follows
X X
h , i≍ (dℓ1/2 2 1/2
α ) + (dℓα ◦ J) +
2
(dℓβ )2 ,
α∈σ β∈χ

where, given Xσ ∈ Tσ and χ there is a neighborhood U of Xσ in T in which


the comparison holds uniformly.
This has the immediate corollary:
Corollary 3.5. If χ is a relative length basis at Xσ ∈ Tσ , then there is
a neighborhood V of Xσ in T such that for every X ∈ V ∩ T , (σ, χ) is a
combined length basis at X.
We introduce more notation that will be used in the proof of Proposi-
tion 3.2. For σ ∈ C and η > 0, we set
U (σ, η) = {X ∈ T | ℓσ (X) < η},
30 K. BURNS, H. MASUR AND A. WILKINSON

so that [
U (η) = U (σ, η).
σ∈C
Let P : T → M be the quotient map from T to the Deligne-Mumford com-
pactification M under the action of the mapping class group MCG. Note
that P (U (η)) is a deleted open neighborhood of ∂M in M.
Proof of Proposition 3.2. Denote by Ck the set of all k-simplices in C, where
k = 0, . . . , 3g+n−4. The mapping class group has finitely many orbits in Ck .
For each k, let Σk be a minimal (necessarily finite) collection of k-simplices
with the property that [
Ck = g∗ (Σk ).
g∈MCG

For each k, let Mk ⊂ M be the union over σ ∈ Σk of P (Tσ ).


Let σ ∈ Σ3g+n−4 be a maximal simplex. Then Tσ contains a single point
Xσ and χ = ∅ is a relative length basis at Xσ . Corollary 3.5 implies there
exists η = η3g+n−4 > 0 such that the collection of open sets:
V3g+n−4 = {U (σ, 2η) | σ ∈ Σ3g+n−4 }
has the property that (σ, ∅) is a combined length basis in V , for every V ∈
V3g+n−4 . Moreover, the elements of V3g+n−4 project to an open cover of the
compact set [
P (U (σ, η))
σ∈Σ3g+n−4

in M.
Working inductively, suppose that there exist constants ck > 1, ηk > 0,
and a finite collection Vk of open sets in the augmented space T such that
(Ik ) each V ∈ Vk has a combined length basis (σ, χ) for some σ ∈ Σk ,
satisfying
1/ck < ℓχ (X) ≤ ℓχ (X) < ck ,
for allSX ∈ V ∩ T ;
S
(IIk ) j≥k V ∈Vj P (V ) contains the compact neighborhood
[ [
P (U (σ, ηk ))
j≥k σ∈Σj

of ∪j≥k Mj in M.
S S
The set N = Mk−1 \ j≥k V ∈Uj P (V ) is compact in M. Proposition 3.3
and Corollary 3.5 imply that for each X ∈ N there is an open set V in T that
projects to a neighborhood of X in M and a relative length basis (σ, χ) for
V ∩ T , for some σ ∈ Σk−1 . The projections of these open sets to M cover
N ; extracting a finite subcover and taking Vk−1 to be the corresponding
collection of open sets in T , we obtain that Ik−1 and IIk−1 hold for some
ck−1 > 1, ηk−1 > 0.
THE WEIL-PETERSSON GEODESIC FLOW 31

We set c = supj≥0 cj , η = inf j≥0 ηj and let


[
V= Vj .
j≥0

Let δ > 0 be the Lebesgue number of the open covering V of the compact
set P (U (η)) in M.
To finish the proof, we define the countable cover U to be the collection
U = {g(V ) ∩ T | g ∈ MCG, and V ∈ V}.
Note that if (σ, χ) is the combined length basis for V ∈ V satisfying
1/c < ℓχ (X) ≤ ℓχ (X) < c,
for all X ∈ V ∩ T , then (g∗−1 σ, g∗−1 χ) is a combined length basis for g(V )
satisfying
1/c < ℓg∗−1 χ (X) ≤ ℓg∗−1 χ (X) < c,
for all X ∈ g(V ) ∩ T . Since MCG acts by W P isometries, the result now
follows. ⋄

3.2. First and second order properties of the WP metric. For each
c > 1, and (σ, χ) ∈ B, let
Ω(σ, χ, c) = {X ∈ T | ℓσ∪χ (X) < c, and 1/c < ℓχ (X) }.
Wolpert proved key estimates on the WP metric in Ω(σ, χ, c), which we
summarize in the following three propositions.
The first set of estimates expand upon and refine the statement in Theo-
rem 3.4.
Proposition 3.6 (First order estimates). [35] Fix c > 1. For any (σ, χ) ∈ B,
the following estimates hold uniformly on Ω(σ, χ, c):
(1) if α, α′ ∈ σ, then
1
hJλα , Jλα′ i = hλα , λα′ i = δα,α′ + O((ℓα ℓα′ )3/2 );

(2) if α, α′ ∈ σ and β ∈ χ, then
hλα , Jλα′ i = hJλα , grad ℓβ i = 0;
(3) if β, β ′ ∈ χ, then
hgrad ℓβ , grad ℓβ ′ i ≍ 1;
moreover, hgrad ℓβ , grad ℓβ ′ i extends continuously to Tσ ;
(4) if α ∈ σ and β ∈ χ, then
hλα , grad ℓβ i = O(ℓ3/2
α );
32 K. BURNS, H. MASUR AND A. WILKINSON

(5) if X ∈ Ω(σ, χ, c), then


!1/2 !
X X
d(X, Tσ ) = 2π ℓα (X) +O ℓ5/2
α (X) .
α∈σ α∈σ

The second set of Wolpert’s estimates are formulae for covariant deriva-
tives, which are described in the next proposition. In each formula in the
next proposition, the error term is a vector, and the expression v = O(a)
means that the WP length of v is O(a).
Proposition 3.7 (Second order estimates). [35] Fix c > 1. For any (σ, χ) ∈
B, the following estimates hold uniformly on Ω(σ, χ, c):
(1) for any vector v ∈ T Ω(σ, χ, c), and α ∈ σ, we have
3
∇ v λα = 1/2
hv, Jλα iJλα + O(ℓ3/2
α kvkW P );
2πℓα
(2) for β ∈ χ and α ∈ σ, we have
∇λα grad ℓβ = O(ℓ1/2
α ), ∇Jλα grad ℓβ = O(ℓα1/2 );
(3) for β, β ′ ∈ χ, ∇grad ℓβ grad ℓβ ′ extends continuously to Tσ .
The final set of Wolpert’s estimates we use involve the WP curvature
tensor.
Proposition 3.8 (Bounds on curvature). [35] Fix c > 1. For any (σ, χ) ∈ B,
the following estimates hold uniformly on Ω(σ, χ, c). For all α ∈ σ we have
3
(6) hR(λα , Jλα )Jλα , λα i = + O(ℓα ).
16π 2 ℓα
Moreover for any quadruple (v1 , v2 , v3 , v4 ) ∈ {λα , Jλα , grad ℓβ }4α∈σ,β∈χ that
is not a curvature-preserving permutation of (λα , Jλα , Jλα , λα ) for some
α ∈ σ, we have:
(7) hR(v1 , v2 )v3 , v4 i = O(1).5
3.3. Curvature estimates along a geodesic. Fix a unit speed WP ge-
odesic γ : I → T in Teichmüller space. For each simple closed curve α we
define functions fα = fα,γ : I → R>0 and rα = rα,γ : I → R>0 by
fα (t) = ℓ1/2 2 2 2
α (γ(t)), and rα (t) = hλα , γ̇(t)i + hJλα , γ̇(t)i .
Roughly, rα measures of the speed of the geodesic γ in the complex line field
spanned by {λα , Jλα }. Wolpert used the function rα to study the behavior
of geodesics terminating in the boundary strata of T . We will use rα and
fα to bound sectional curvatures along γ. We summarize in the next few
lemmas the key properties of rα and fα that will be used in the sequel.
5Wolpert actually proves more: each vector v appearing in this expression that is of
i
1/2
the form λα or Jλα introduces a multiplicative bound O(ℓα ) in the curvature tensor.
This means that there are sectional curvatures that are arbitrarily close to 0.
THE WEIL-PETERSSON GEODESIC FLOW 33

The first property is an immediate consequence of part (5) of Proposi-


tion 3.6 and explains the significance of the quantity fα .
Lemma 3.9. Fix c > 1. For every (σ, χ) ∈ B and any γ, if γ(t) ∈ Ω(σ, χ, c),
then
!1/2
X
d(γ(t), Tσ ) = 2π fα2 (t) + O(fα5 (t)).
α∈σ

The next two lemmas will allow us to bound the variations of rα and fα
along a geodesic.
Lemma 3.10. Fix c > 1. For every (σ, χ) ∈ B and any γ, if γ(t) ∈
Ω(σ, χ, c), then
rα′ (t) = O(fα3 (t)),
for every α ∈ σ.
Proof. Since the WP metric is Kähler, the almost complex structure J is
parallel, and so we have
D D
2rα (t)rα′ (t) = 2hλα , γ̇(t)ihλα , γ̇(t)i + 2hJλα , γ̇(t)ihJ λα , γ̇(t)i.
∂t ∂t
Now by part (1) of Proposition 3.7, we have
D 3
λα = hγ̇, Jλα i Jλα + O(fα3 )
∂t 2πfα
and
D 3
J λα = −hγ̇, Jλα i λα + O(fα3 ).
∂t 2πfα
Plugging this in to the formula for 2rα (t)rα′ (t), and noting that
max{|hλα , γ̇i|, |hJλα , γ̇i|} < rα ,
we get:
3 3
2rα (t)rα′ (t) = hλα , γ̇ihγ̇, Jλα i2 − hλα , γ̇ihγ̇, Jλα i2 + O(rα fα3 )
πfα πfα

= O(rα fα3 ).

Lemma 3.11. Fix c > 1. For every (σ, χ) ∈ B and any γ, if γ(t) ∈
Ω(σ, χ, c), then

rα2 (t) = (fα′ (t))2 + fα (t)fα′′ (t) + O(fα4 (t)),
3
for every α ∈ σ.
34 K. BURNS, H. MASUR AND A. WILKINSON

1/2
Proof. Since λα = grad ℓα , it follows that fα′ = hλα , γ̇i; differentiating this
expression, we obtain using part (1) of Proposition 3.7:
d 3
fα′′ = hλα , γ̇i = h∇γ̇ λα , γ̇i = hγ̇, Jλα i2 + O(fα3 ).
dt 2πfα (t)
Now multiplying this last expression by fα and adding it to fα′2 , the result
then follows from the definition of rα2 . ⋄

Let
2
k (t) = sup −hR(v, γ̇(t))γ̇(t), vi.
1 T
v∈Tγ(t)

2
We next bound k in terms of rα and fα ; the preceding estimates on rα
and fα will then allow us to bound solutions to the differential inequality
2
u′ ≤ k − u2 .
Lemma 3.12. Fix c > 1. For any (σ, χ) ∈ B and any unit speed geodesic
γ, if (σ, χ) is a combined length basis in U ⊂ Ω(σ, χ, c), and γ(t) ∈ U , then
2 X  r 2 (t) 
α
k (t) = O 2
.
α∈σ
f α (t)

Proof. Since (σ, χ) is a combined length basis, we can write v ∈ T 1 Ω(σ, χ, c)


and γ̇ as
X X
v= (aα λα + bα Jλα ) + cβ grad ℓβ
α∈ σ β∈χ

and
X X
γ̇ = (a′α λα + b′α Jλα ) + c′β grad ℓβ .
α∈ σ β∈χ

Now v and γ̇ are unit vectors, the above estimates on the metric say that all
coefficients aα , bα , cβ , a′α , b′α , c′β are O(1). Moreover by these same estimates
and the definition of rα , we have
1
rα2 = (a′2 + b′2 3
α ) + O(fα ).
4π 2 α
It now follows from Proposition 3.8 that
X
−hR(v, γ̇)γ̇, vi = − (a2α b′α 2 + a′α 2 b2α )hR(λα , Jλα )Jλα , λα i + O(1)
α∈σ
X  
rα2
= O + O(1).
α∈σ
fα2

THE WEIL-PETERSSON GEODESIC FLOW 35

3.4. Estimates on rα /fα . In the next two subsections, we establish how


2
bounds on k can be used to bound solutions to the Riccati inequality u′ ≤
2 2
k − u2 . The main result in this subsection will give us control over k .
Proposition 3.13. Fix c > 1. There is a constant A = A(c) > 0 such that
for any (σ, χ) ∈ B, for any unit speed WP segment γ : [−δ, δ] → Ω(σ, χ, c)
and any α ∈ σ, we have
 
rα (t) rα (t0 )
≤ A max 1, for 0 ≤ t ≤ δ,
fα (t) rα (t0 )|t − t0 | + fα (t0 )
where t0 is the unique time in [0, δ] such that fα (t) ≥ fα (t0 ) for 0 ≤ t ≤ δ.
Proof of Proposition 3.13. The time t0 is uniquely defined since fα (t) is a
convex function of t. It will suffice to prove the proposition under the addi-
tional assumption that fα (t) is increasing for t ≥ 0. If t0 = 0, we apply this
restricted form of the proposition directly to the geodesic γ; if t0 = δ, we
apply it to the geodesic t 7→ γ(δ − t); and if 0 < t0 < δ, we consider both of
the geodesics t 7→ γ(t − t0 ) and t 7→ γ(t0 − t).
We choose C ≥ 1 large enough so that:

(C1 ) the O(fα4 ) term in the equation rα2 = (fα′ )2 + fα fα′′ + O(fα4 ) given
3
by Lemma 3.11 is at most Cfα4 ;
rα 1
(C2 ) ≤ ;
C 2
(C3 ) |rα′ | ≤ Cfα3 (which is possible by Lemma 3.10).
Conditions (C1 ) and (C2 ) give a lower bound on fα′′ when rα /fα ≥ C and
|fα′ |
is small.
rα rα 3 rα2
Lemma 3.14. If ≥ C and |fα′ | ≤ , then fα′′ ≥ .
fα 2 4π fα
Proof. By (C1 ) and (C2 ),
2π r2 2π r2 r2 2π
rα2 = (fα′ )2 + fα fα′′ + O(fα4 ) ≤ α + fα fα′′ + α ≤ α + fα fα′′ .
3 4 3 4C 2 3

We continue with the proof of Proposition 3.13. Recall we are assuming


t0 = 0. We have that f (t) increasing for t ≥ 0. We shall show that
 
rα (t) 32πrα (0)
≤ max 4C, for 0 ≤ t ≤ δ.
fα (t) fα (0) + trα (0)
rα (t)
If ≤ 4C for 0 ≤ t ≤ δ we are done. Otherwise, let
fα (t)
rα (t)
b = sup{t ∈ [0, δ] : ≥ 4C}.
fα (t)
36 K. BURNS, H. MASUR AND A. WILKINSON

rα (t)
Since ≤ 4C for b ≤ t ≤ δ, it will suffice to show that
fα (t)
rα (t) 32πrα (0)
(8) ≤ for 0 ≤ t ≤ b.
fα (t) fα (0) + trα (0)
The function rα is approximately constant and rα /fα is large on the
interval [0, b].
Lemma 3.15. For 0 ≤ t ≤ b we have:
rα (0)
(i) ≤ rα (t) ≤ 2rα (0);
2
rα (t)
(ii) ≥ C.
fα (t)
Proof. By (C3 ), |rα′ | ≤ Cfα3 on the interval [0, b]. Since b ≤ 1 and fα is
increasing on [0, b], we have |rα (b) − rα (t)| ≤ Cfα3 (b) for 0 ≤ t ≤ b. The
rα (b) 1
definition of b and (C2 ) ensure that fα (b) ≤ ≤ by (C2 ). Hence
2C 4
|rα (b) − rα (t)| Cfα3 (b) 1 1
≤ ≤ fα2 (b) ≤ .
rα (b) 2Cfα (b) 2 32
31 rα (t) 33
Thus ≤ ≤ for 0 ≤ t ≤ b, and (i) follows easily. Claim (ii)
32 rα (b) 32
follows from (i) since rα (b)/fα (b) ≥ 2C and fα is increasing on [0, b]. ⋄

Using this lemma we see that inequality (8) will follow if we prove
(9) 16πfα (t) ≥ fα (0) + trα (0) for 0 ≤ t ≤ b.
fα (0)
Lemma 3.15 ensures that rα (0) > 0, so we can set a = . Now for
rα (0)
0 ≤ t ≤ min(a, b), we have
fα (0) + trα (0) ≤ fα (0) + arα (0) = 2fα (0) ≤ 2fα (t),
since fα is increasing on [0, δ]. This gives (9) for 0 ≤ t ≤ min(a, b).
We are done if a ≥ b. It remains to show that if a ≤ b, then inequality (9)
also holds for a ≤ t ≤ b. Since fα is convex and (9) already holds for t = a,
it will suffice to show that if a ≤ b that
(10) 16πfα′ (a) ≥ rα (0)
We may assume that 4fα′ (a) ≤ rα (0), since otherwise there is nothing to
prove. Then fα′ (t) ≤ fα′ (a) ≤ rα (0)/4 for 0 ≤ t ≤ a, because fα is convex
and increasing on [0, a]. Since a ≤ b, we can now apply Lemma 3.15 to see
that on [0, a] we have
rα (t)
fα′ (t) ≤ rα (0)/4 ≤ rα (t)/2 and ≥ C.
fα (t)
THE WEIL-PETERSSON GEODESIC FLOW 37

Thus both hypotheses of Lemma 3.14 are satisfied on [0, a]. Lemmas 3.14
and 3.15 give us
3 rα2 (t) 3 rα2 (0) 1 rα2 (0)
fα′′ (t) ≥ ≥ >
4π fα (t) 16π fα(t) 8π fα (t)
for 0 ≤ t ≤ a. Since fα′ ≤ rα (0)/4 on [0, a], we have
fα (a) ≤ fα (0) + arα (0)/4 = fα (0) + fα (0)/4 < 2fα (0),
and hence
1 rα2 (0)
fα′′ (t) ≥ ,
16π fα (0)
for 0 ≤ t ≤ a. Finally, since fα is increasing on [0, a], we have fα′ (0) ≥ 0 and
a rα2 (0) rα (0)
fα′ (a) ≥ = ,
16π fα 16π
which is the desired inequality (10).

Combining Lemma 3.12 and Proposition 3.13 we obtain the immediate


corollary:
Corollary 3.16. Fix c > 1. There is a constant B = B(c) > 0 such
that for any (σ, χ) ∈ B, if (σ, χ) is a combined length basis in an open set
U ⊂ Ω(σ, χ, c) and γ : [−δ, δ] → U is a unit-speed WP geodesic segment,
then
 
rα (tα )
k(t) ≤ B max 1, for 0 ≤ t ≤ δ,
α∈σ rα (tα )|t − tα | + fα (tα )
where tα is the unique time in [−δ, δ] such that fα (t) ≥ fα (tα ) for −δ ≤ t ≤
δ.
3.5. Controlled functions and bounds on Riccati solutions.
Lemma 3.17. Let κ : (−δ, δ) → R>0 be a continuous function such that the
right derivative
κ(t + h) − κ(t)
DR κ(t) = lim
h→0+ h
is defined for all t ∈ (δ, δ), for some δ > 0. Suppose there is a constant
Q > 1 such that
DR κ(t) ≥ −(Q − 1)κ2 (t)
for all t ∈ (−δ, δ).
Let u(t) be a solution of the differential inequality
(11) u′ ≤ κ2 − u2
that is defined for all t ∈ (δ, δ). Then
 
1
u(t) ≤ 2Q max , κ(t)
δ
for all t ∈ (0, δ).
38 K. BURNS, H. MASUR AND A. WILKINSON

Proof. Replacing κ by κ̂ = max{δ−1 , κ}, we obtain a new function satisfying


DR κ̂(t) ≥ −(Q − 1)κ̂2 (t)
for all t ∈ (−δ, δ). Moreover, any solution to (11) is also a solution to
u′ ≤ κ̂2 − u2 .
We may therefore assume that κ(t) > δ−1 , for all t ∈ (−δ, δ). Moreover,
increasing the value of Q slightly, we may assume that
DR κ(t) > −(Q − 1)κ2 (t).
Suppose that u is solution of (11) with u(t0 ) > 2Qκ(t0 ) for some t0 ∈
(0, δ); we prove that u(t) is not defined for all t < t0 . It will suffice to show
that if v0 is a solution of differential equation
(12) v ′ = κ2 − v 2
with v0 (t0 ) > 2Qκ(t0 ) for some t0 ∈ (0, t0 ), then v0 (t) blows up in time < δ
as t decreases from t0 . We begin by finding a solution v1 (t) that is defined
for all t ∈ (−δ, δ) and satisfies 0 ≤ v1 ≤ Qκ.
Consider a solution v of (12). If v(t) = Qκ(t) for some t, then
v ′ (t) ≤ (1 − Q2 )κ2 (t) ≤ (Q − Q2 )κ2 (t) = −Q(Q − 1)κ2 (t) < QDR κ(t).
It follows that if v is a solution of (12) defined on an interval containing a
time t0 such that v(t0 ) ≤ Qκ(t0 ), then v(t) ≤ Qκ(t) for all t ≥ t0 in the
interval. Moreover if v(t) = 0 for some t, then v ′ (t) = κ2 (t) ≥ 0. Thus
if 0 ≤ v(t0 ) ≤ Qκ(t0 ) for some t0 , then v(t) is defined for t0 ≤ t < δ and
satisfies 0 ≤ v ≤ Qκ on the interval [t0 , δ). Hence the solution v1 with initial
condition v1 (−δ) = 0 satisfies 0 ≤ v1 ≤ Qκ on (−δ, δ).
We now show that v0 (t) and v1 (t) diverge rapidly as t decreases from
t0 . Since v0 (t0 ) > v1 (t0 ) and the graphs of two solutions of (12) cannot
intersect, we have
(13) v0 > v1 ≥ 0
on the interval where v0 is defined.
Let ∆(s) = v0 (t0 −s)−v1 (t0 −s). Observe that ∆(0) > 2Qκ(t0 )−Qκ(t0 ) =
Qκ(t0 ) > 1δ . Then ∆(s) > 0 and it follows from (12) and (13) that
∆′ (s) = v02 (t0 − s) − v12 (t0 − s) ≥ [v0 (t0 − s) − v1 (t0 − s)]2 = ∆(s)2 .
Hence for s ≥ 0 we have
1
∆(s) ≥ ,
(1/∆(0)) − s
which blows up as s → 1/∆(0) < δ. Since v1 stays bounded, it follows that
v0 (t) → ∞ as t ց t0 − 1/∆(0) ∈ (−δ, t0 ). ⋄

Let us say that a function κ is Q-controlled if it satisfies the hypotheses


in the previous lemma. If κ is Q-controlled, then it is Q′ -controlled, for all
Q′ > Q. If κ is Q-controlled, then so is t 7→ κ(t − t0 ) for any t0 , and for any
THE WEIL-PETERSSON GEODESIC FLOW 39

A > 0, the function Aκ is A(Q − 1) + 1-controlled. The maximum of two Q-


controlled functions is Q-controlled. Moreover κ is 1-controlled if κ(t) = 1
1
and 2-controlled if κ(t) = where a > 0.
|t| + a
Proposition 3.18. There exist constants C ≥ 2, δ ∈ (0, 1) such that for
any positive δ′ < δ and any geodesic segment γ : (−δ′ , δ′ ) → T , there exists
a C-controlled function κ : (−δ′ , δ′ ) → R>0 such that for every t ∈ (−δ′ , δ′ ):
2
(1) k (t) ≤ κ2 (t), where
2
k (t) = sup −hR(v, γ̇(t))γ̇(t), vi;
1 T
v∈Tγ(t)

(2)
Z δ′
κ(s) ds ≤ C| ln(ρδ′ (γ̇(0)))|,
−δ′
and
κ(δ′ ) ≤ C(ρδ′ (γ̇(0)))−1 ,
where ρδ′ (γ̇(0)) is the distance from the geodesic segment γ[−δ′ , δ′ ]
to ∂T .
Proof. Let c, η and δ be the constants and let U be the collection of open
sets in T given by Proposition 3.2. We write
T = U (η) ⊔ Θ;
the set Θ = T \ U (η) lies in the thick part of Teichmüller space in which
the WP sectional curvatures are negative and bounded below by a constant
−b2 . By shrinking the value of δ if necessary, we may assume that for every
X ∈ Θ, and Y ∈ T , if d(X, Y ) < δ, then:
sup −hR(v, w)w, vi < b2 .
v,w∈TY1 T

Let B = B(c) > 0 be the constant given by Corollary 3.16.


Fix δ′ < δ. It follows from Proposition 3.2 that if γ : (−δ′ , δ′ ) → T is a
unit-speed WP geodesic, then either γ(0) ∈ Θ, or
γ(−δ′ , δ′ ) ⊂ U,
for some U ∈ U .
If γ(0) ∈ Θ, then we define κ to be the constant function b. Then by
2
construction we have k ≤ κ2 . Since the W P distance from any point in T
to ∂T is bounded above by a uniform constant, it also follows that in this
case: Z ′ δ
κ(s) ds = 2bδ′ = O(| ln(ρδ′ (γ̇(0)))|),
−δ′
and
κ(δ′ ) = O(ρδ′ (γ̇(0)))−1 .
40 K. BURNS, H. MASUR AND A. WILKINSON

Suppose on the other hand that γ(δ′ , δ′ ) ⊂ U , for some U ∈ U . Let (σ, χ)
be the combined length basis in U given by Proposition 3.2 satisfying:
1/c < ℓχ (X) ≤ ℓχ (X) < c,
for every X ∈ U . For α ∈ σ, define κα : (−δ′ , δ′ ) → R>0 by:
rα (tα )
κα (t) = ,
rα (tα )|t − tα | + fα (tα )
where tα is the unique time in [−δ′ , δ′ ] such that fα (t) ≥ fα (tα ) for t ∈
[−δ′ , δ′ ]. Observe that κα is a 2-controlled function and attains its maximum
value of frαα (tα)
(tα ) at t = tα .
Applying Corollary 3.16, we obtain that for all t ∈ (−δ′ , δ′ ):
k(t) ≤ B max{1, κα }.
α∈σ

We define κ : (−δ′ , δ′ ) → R>0 by:


κ = B max{1, κα }.
α∈σ
Since κα is 2-controlled, for each α, it follows that κ is B + 1-controlled. By
2
its construction κ satisfies the inequality k < κ2 on (−δ′ , δ′ ).
It remains to estimate the integral of κ over the interval (−δ′ , δ′ ). Simple
integration shows that
Z δ′
κα (s) ds = O(max{δ′ , |ln(fα (tα ))|)},
−δ′
since rα (tα ) = O(1).
1/2
Note that fα (tα ) is the minimum value of the function ℓα along the
geodesic segment γ[−δ, δ]. Lemma 3.9 implies that there exists a constant
r > 0 such that fα (tα ) ≥ rρt (γ̇(0)). This implies that
Z t Z t
κ(s) ds ≤ B max κα (s) ds
0 α∈σ 0
= O(| ln(ρt (γ̇(0)))|).
Similarly,
rα (tα )
κ(0) ≤ B max
α∈σ fα (tα )

= O(| ln(ρt (γ̇(0)))|).


Combining Proposition 3.18 and Lemma 3.17, we obtain:


Proposition 3.19. There exists C > 1 such that for any infinite geodesic
ray γ : (−∞, 1) → T , if u(t) is a solution to the differential inequality
2
(14) u′ ≤ k − u2
THE WEIL-PETERSSON GEODESIC FLOW 41

that is defined for all t < 1, then for every t ∈ (0, 1):
Z t
u(s) ds ≤ C| ln(ρt (γ̇(0)))|,
0
and
u(0) ≤ C(ρt (γ̇(0)))−1 .
3.6. Proof of Theorem 3.1. The theorem follows immediately from Propo-
sition 2.1, which bounds the derivative of the time t map in terms of solutions
2
u to the differential inequality u′ ≤ k − u2 , and the estimates in Proposi-
tion 3.19 for u and its integral.

4. Bounding the second derivative of the geodesic flow


The results in this section will be used to derive bounds on the second
derivative of the W P geodesic flow, on the order of an inverse power of the
distance to the singular locus. Since these results hold in a more general
context, we return to the setting of Riemannian geometry.
4.1. More on the Sasaki metric and statement of the general result.
Let M be a Riemannian manifold, and let π : T M → M be the canonical
projection. The Sasaki metric on T 1 M induces a Sasaki metric on T T 1 M ,
which for brevity we will also call the Sasaki metric (although strictly speak-
ing it is some sort of Sasaki Sasaki metric). In general we will denote the
Sasaki distance on T 1 M by dSas and on T T 1 M by dSas .
Recall that for v ∈ T 1 M , each vector ξ ∈ Tv T 1 M can be naturally identi-
1 M )2 with w ⊥ v. The distance d
fied with a pair (u, w) ∈ (Tπ(v) 1
Sas on T T M
induced by this metric can be estimated as follows. Let ξ0 = (u0 , w0 ) ∈
1
(Tπ(v M )2 and ξ1 = (u1 , w1 ) ∈ (Tπ(v
1 M )2 be tangent vectors in T T 1 M
0) 1)
based at v0 and v1 respectively. Let σ be a Sasaki geodesic in T 1 M from v0
1
to v1 . Let Pσ : Tπ(v M → Tπ(v1 M be parallel translation along the curve
0) 1)
of basepoints π ◦ σ in M . The following lemma follows from the discussion
in Section 2.
Lemma 4.1. For each v0 there exists an ǫ > 0 such that if dSas (v0 , v1 ) < ǫ,
then
dSas (ξ0 , ξ1 ) ≤ dSas (v0 , v1 ) + ku1 − Pσ (u0 )k + kw1 − Pσ (w0 )k
≤ 2 dSas (ξ0 , ξ1 ).
If R is the curvature tensor of a metric on M and x ∈ N , we define
kRx k = sup kRx (v1 , v2 )v3 k.
v1 ,v2 ,v3 ∈Tx1 N

and
k∇Rx k = sup k∇v1 Rx (v2 , v3 )v4 k.
v1 ,v2 ,v3 ,v4 ∈Tx1 N
The main result in this section is:
42 K. BURNS, H. MASUR AND A. WILKINSON

Proposition 4.2. Let M be an m-dimensional Riemannian manifold, pos-


sibly incomplete, and let t0 ≤ 1 be a positive number. Let γ : [−t0 , t0 ] → M
be a unit-speed geodesic segment whose image lies in the interior of M .
Suppose that there exist constants C1 , C2 , C3 > 1 and ǫ0 > 0 such that for
all t ∈ (−t0 , t0 ):
(1) if v ∈ T 1 M satisfies dSas (v, γ̇(0)) < ǫ0 , then
max{kDv ϕt k, kDϕt (v) ϕ−t k} ≤ C1 ,
(2) if p ∈ M satisfies d(p, γ(t)) < ǫ0 , then
kRp k ≤ C2 and k∇Rp k ≤ C3 .
Then there exists ǫ1 > 0 such that for every t ∈ (−t0 , t0 ), for every pair
v0 , v1 ∈ T 1 M , with dSas (vi , γ̇(0)) < ǫ1 , and for all ξi ∈ Tv1i T 1 M , (i = 0, 1),
we have:
dSas (Dϕt (ξ0 ), Dϕt (ξ1 )) ≤ (8mC14 C22 C3 ) dSas (ξ0 , ξ1 ).
4.2. Variations of solutions to linear ODEs. To prove Proposition 4.2,
we first treat the linearized version of the problem. We begin with a basic
fact about solutions to linear ODEs. Recall that if A(t) : [0, T ] → L(Rm ) is
any continuous family of linear transformations of Rm , then the ODE
(15) x′ (t) = A(t)x(t); x : [0, T ] → Rm
has a unique fundamental solution F (t) : [0, T ] → L(R), which satisfies
F (0) = I and
F ′ (t) = A(t)F (t);
every solution to (15) with initial condition x(0) = x0 then takes the form
x(t) = F (t)x0 .
Lemma 4.3. Let Fi : [0, T ] → L(Rm ) be the fundamental solution to the
differential equation x′ (t) = Ai (t)x(t), for i = 0, 1. Then
kF0 (T ) − F1 (T )k ≤ T kF0 k0 kF0−1 k0 kA0 − A1 k0 kF1 k0 ,
where the k·k0 is the supremum of the operator norm over the interval [0, T ].
Proof. For t ∈ [0, T ], let W (t) = F0 (t) − F1 (t). Since Fi′ (t) = Ai (t)Fi (t) and
Fi (0) = I for i = 0, 1, we obtain by subtraction that W is a solution to the
affine initial value system:
W ′ = A0 F0 − A1 F1 = A0 W + B W (0) = 0,
where B(t) = (A0 (t) − A1 (t))F1 (t). This system can be solved explicitly; we
obtain that Z t 
W (t) = F0 (t) F0 (s)−1 B(s) ds ,
0
as can be verified directly by differentiation (W ′ = F0′ F0−1 W + B = A0 W +
B). The desired bound on kW (T )k now follows immediately. ⋄
THE WEIL-PETERSSON GEODESIC FLOW 43

Consider next the second-order linear ODE


(16) x′′ (t) = −R(t)x(t)
where R : [0, T ] → L(Rm ); in our application R(t) will be a matrix repre-
senting the sectional curvature operator along a geodesic γ and (16) will be
the Jacobi equation in a suitably chosen coordinate system along γ.
Then (16) can be transformed into afirst order  system in the standard
x(t)
way by introducing the variable z(t) = ∈ R2m and the additional
y(t)
constraint x′ (t) = y(t). Then z satisfies the first order ODE:
(17) z ′ (t) = A(t)z(t),
where
 
0 I
A(t) = .
−R(t) 0
The fundamental solution F (t) to this equation has the property that if
x(t) is asolution to (16) ′
  with initial values x(0) = x0 , x (0) = y0 , then
x(t) x0
= F (t) .
x′ (t) y0
We obtain the corollary:
Corollary 4.4. Let Fi : [0, T ] → L(R2m ) be the fundamental solution to the
differential equation x′′ (t) = −Ri (t)x(t), for i = 0, 1. Then
kF0 (T ) − F1 (T )k ≤ T kF0 k0 kF0−1 k0 kR0 − R1 k0 kF1 k0 .
4.3. Proof of Proposition 4.2. We now return to the setting of differential
geometry and finish the proof of Proposition 4.2. Let γ : [−t0 , t0 ] → M be
given. We start with a lemma.
Lemma 4.5. With the assumptions of Proposition 4.2 suppose p0 , p1 ∈ M
satisfy d(pi , γ(t)) < ǫ0 , then for all vi , wi ∈ Tp1i M , the curvature tensor R
satisfies:
kR(v0 , w0 )w0 k ≤ C2 ,
and
dSas (R(v0 , w0 )w0 , R(v1 , w1 )w1 ) ≤ C3 (dSas (v0 , v1 ) + dSas (w0 , w1 )).
Proof. This follows in a straightforward way from the Mean Value Theorem
and the hypotheses that kRp k ≤ C2 and k∇Rp k ≤ C3 , for all p ∈ M with
d(p, γ(t)) < ǫ0 . ⋄

Let v0 , v1 ∈ T 1 M be unit tangent vectors in a neighborhood of γ̇(0), and


let σ : (−2, 2) → T 1 M be a Sasaki geodesic with σ(0) = v0 and σ(1) =
v1 . Each σ(s) determines a unit speed geodesic γs : (−t0 , t0 ) → M with
γ̇s (0) = σ(s). In this way σ determines a variation of geodesics α : (−2, 2) ×
(−t0 , t0 ) → M with the property that α(s, t) = γs (t).
44 K. BURNS, H. MASUR AND A. WILKINSON

We may assume that the norms of the derivatives of α are uniformly


bounded from above by a constant, say 1. For s ∈ (−2, 2), let Ls (t) =
∂α/∂s(s, t) be the induced Jacobi field along along γs . Choose ǫ1 such that
if dSas (vi , γ̇(0)) < ǫ1 for i = 0, 1, then dSas (γ̇s (t), γ̇s (t)) < ǫ0 for all (s, t) ∈
(−2, 2) × (−t0 , t0 ), where ǫ0 is given by the hypotheses of the proposition.
If dSas (vi , γ̇(0)) < ǫ1 , then for any (s, t) ∈ (−2, 2) × (−t0 , t0 ) we have
Z s
dSas (γ̇s (t), γ̇0 (t)) ≤ k(Lu (t), L′u (t)))kSas du.
0
Since σ is a Sasaki geodesic the above inequality is an equality in the case
of t = 0; that is
Z s
kLu (0), L′u (0)kSas du = dSas (γ̇s (0), γ̇0 (0)).
0
By the bounds (1) on the first derivative of the geodesic flow (which bounds
the growth of Jacobi fields), we also have that for any u, t,
k(Lu (t), L′u (t))kSas ≤ C1 k(Lu (0), L′u (0))kSas ,
and so
Z s Z s
k(Lu (t), L′u (t)))kSas du ≤ C1 kLu (0), L′u (0)kSas du.
0 0
Putting these inequalities together, we obtain:
(18) dSas (γ̇s (t), γ̇0 (t)) ≤ C1 dSas (γ̇s (0), γ̇0 (0)).
Our goal is to bound the Lipschitz norm of the derivative of the time-t map
of the geodesic flow ϕt at γ̇0 (0). The conclusion of Proposition 4.2 will follow
if we show that for any (s, t) ∈ (−2, 2) × (−t0 , t0 ), and any ξ0 ∈ Tγ̇10 (0) T 1 M
and ξs ∈ Tγ̇1s (0) T 1 M , we have:

(19) dSas (Dϕt (ξ0 ), Dϕt (ξs )) ≤ (4mC14 C22 C3 ) dSas (ξ0 , ξs ).
Recall that under the standard identification of ξs ∈ Tγ̇s (0) T M with a pair
(us , ws ) ∈ (Tγs (0) M )2 , the vector Dγ̇s (0) ϕt (ξs ) is identified with (Js (t), Js′ (t)),
where Js is the solution to the (second-order) Jacobi equation
(20) J ′′ + R(J, γ̇s )γ̇s = 0
with initial condition (Js (0), Js′ (0)) = (us , ws ).
To analyze the variation of solutions to this ODE, we fix convenient coor-
dinates for the tangent bundle to the geodesic γs in order to express (20) as
a matrix equation of the form (16). To this end, let {ej (0, 0) : j = 1, . . . , m}
be an orthonormal frame at γ0 (0) = α(0, 0) spanning the tangent space
Tγ0 (0) M . We first parallel translate this frame along α(s, 0) to obtain an
orthonormal frame {ej (s, 0)} at γs (0), for s ∈ (−2, 2). We next parallel
translate the frame {ej (s, 0)} along γs (t), for t ∈ (−t0 , t0 ) to obtain a frame
{ej (s, t)} at each point α(s, t).
THE WEIL-PETERSSON GEODESIC FLOW 45

Lemma 4.6. For j ∈ {1, . . . , m}, we have:


dSas (ej (0, t), ej (s, t)) ≤ d(γ0 (0), γs (0)) + 2tC1 C2 dSas (γ̇0 (0), γ̇s (0)),
for all (s, t) ∈ (−2, 2) × (−t0 , t0 ).
Proof. Fix j. Our construction of ej (using parallel translation) gives that
for all s, t:
D D
(21) ej (s, 0) = 0, and ej (s, t) = 0;
∂s ∂t
D
we would like to estimate ∂s ej (s, t) for general s, t. To do this, we first
D D
estimate ∂t ∂s ej (s, t).
It follows directly from the definition of the Riemannian curvature tensor
and the joint integrability of the pair {Ls , γ̇s } that
DD DD
R(Ls (t), γ̇s (t))ej (s, t) = ej (s, t) − ej (s, t)
∂t ∂s ∂s ∂t
DD
= ej (s, t),
∂t ∂s
where we have used the second part of (21) in the last step. Applying the
bound kR(Ls (t), γ̇s (t))ej k ≤ C2 kLs (t)k, we obtain that
DD
k ej (s, t)k ≤ C2 kLs (t)k.
∂t ∂s
Integrating this expression with respect to t, we then have the bound:
Z t Z t
D D
k ej (s, t)k ≤ k ej (s, 0)k + C2 kLs (u)k du = C2 kLs (u)k du.
∂s ∂s 0 0

Integrating again, this time with respect to s, and using Lemma 4.1 and
(18), we obtain:
Z s
D
dSas (ej (0, t), ej (s, t)) ≤ d(γ0 (t), γs (t)) + k ej (u, t)k du
0 ∂s
Z sZ t
≤ d(γ0 (0), γs (0)) + kL′w (u)k du dw
0 0
Z sZ t
+C2 kLw (u)k du dw
0 0
Z sZ t
≤ d(γ0 (0), γs (0)) + 2C2 k(Lw (u), L′w (u))kSas du dw
0 0
Z s
≤ d(γ0 (0), γs (0)) + 2tC1 C2 k(Lw (0), L′w (0))kSas dw
0
= d(γ0 (0), γs (0)) + 2tC1 C2 dSas (γ̇0 (0), γ̇s (0)),
which is the desired bound. ⋄
46 K. BURNS, H. MASUR AND A. WILKINSON

For (s, t) ∈ (−2, 2) × (−t0 , t0 ), the frame {ej (s, t)} gives an isometric
isomorphism between Rm and Tα(s,t) M :
m
X
(x1 , . . . , xm ) 7→ xj ej (s, t).
j=1

This in turn induces for each (s, t) an isometric isomorphism


Is,t : R2m → Tγ̇ (t) T M ∼
= (Tα(s,t) M )2 .
s

Lemma 4.6 has the following immediate corollary.


Corollary 4.7. For each (s, t) ∈ (2, 2) × (−t0 , t0 ) and (Euclidean) unit
vector z ∈ R2m , we have
dSas (Is,t (z), I0,t (z)) ≤ d(γ0 (0), γs (0)) + 2tC1 C2 dSas (γ̇0 (0), γ̇s (0)).
Expressing the Jacobi equation (20) along γs in the coordinates on Tγs Teich(s)
given by Is,t , we obtain the ODE:
(22) x′′ (t) = −Rs (t)x(t),
where
(Rs (t))i,j = h R (ei (s, t), γ̇s (t)) γ̇s (t), ej (s, t) i.
It is here one can see the convenience in choosing the frame {ej } to be
parallel along the geodesic γs ; otherwise, the Jacobi equation would include
terms involving x′ , which would complicate the linear analysis.
Denote by Fs : (−t0 , t0 ) → L(R2m ) the fundamental solution to (22).
Corollary 4.4 implies that for any (s, t) ∈ (−2, 2) × (−t0 , t0 ), we have
kF0 (t) − Fs (t)k ≤ tkF0 k0 kFs−1 k0 kR0 − Rs k0 kFs k0 .
Now the main hypotheses of the proposition, when combined with 4.6 and
(18), and the triangle inequality can be seen to give the upper bound:
kR0 (t) − Rs (t)k ≤ (mC1 C22 C3 ) dSas (γ̇0 (0), γ̇s (0)),
(where we omit the details) and so
kF0 (t) − Fs (t)k ≤ (tmC1 C22 C3 )kF0 k0 kFs k0 kFs−1 k0 dSas (γ̇0 (0), γ̇s (0)),
for all (s, t) ∈ (−2, 2)×(−t0 , t0 ). The bounds on the first derivative of ϕt im-
ply that for all (s, t) ∈ (−2, 2)×(−t0 , t0 ), we have max{kFs (t)k, kFs−1 (t)k} ≤
C1 , which implies that
kF0 (t) − Fs (t)k ≤ (tmC14 C22 C3 ) dSas (γ̇0 (0), γ̇s (0)).
Finally, suppose that ξ0 = I0,0 (z0 ) and ξs = Is,0 (zs ) are arbitrary unit
tangent vectors to T 1 M based at γ̇0 (0) and γ̇s (0), respectively (where z0 , zs
D
are Euclidean unit vectors in R2m ). Since ∂s Is,0 = 0, Lemma 4.1 implies that
the Sasaki distance dSas (ξ0 , ξs ) between ξ0 and ξs is uniformly comparable
to kz0 − z1 k + dSas (γ̇0 (0), γ̇s (0)); in particular:
(23) kz0 − z1 k + dSas (γ̇0 (0), γ̇s (0)) ≤ 2 dSas (ξ0 , ξs ).
THE WEIL-PETERSSON GEODESIC FLOW 47

We may then conclude using Corollary 4.7 and our previous estimates
that:
dSas (Dϕt (ξ0 ), Dϕt (ξs )) = dSas (I0,t (F0 (t)z0 ) , Is,t (Fs (t)zs ))
≤ dSas (I0,t (F0 (t)z0 ) , I0,t (Fs (t)zs ))
+ dSas (I0,t (Fs (t)zs ) , Is,t (Fs (t)zs ))
≤ kF0 (t)z0 − Fs (t)zs k
+kFs (t)zs k (d(γ0 (0), γs (0)) + 2tC1 C2 dSas (γ̇0 (0), γ̇s (0)))
≤ k(F0 (t) − Fs (t))z0 k + kFs (t)(z0 − zs )k
+C1 (d(γ0 (0), γs (0)) + 2tC1 C2 dSas (γ̇0 (0), γ̇s (0)))
≤ (tmC14 C22 C3 ) dSas (γ̇0 (0), γ̇s (0)) + C1 kz0 − zs k
+C1 (d(γ0 (0), γs (0)) + 2tC1 C2 dSas (γ̇0 (0), γ̇s (0))) .
Using the fact that t0 < 1 we obtain
dSas (Dϕt (ξ0 ), Dϕt (ξs )) ≤ (mC14 C22 C3 ) dSas (γ̇0 (0), γ̇s (0)) + C1 kz0 − zs k
+3C12 C2 dSas (γ̇0 (0), γ̇s (0))
≤ (4mC14 C22 C3 ) (kz0 − zs k + dSas (γ̇0 (0), γ̇s (0)))
≤ (8mC14 C22 C3 ) dSas (ξ0 , ξs ),
where we used (23) in the last step. This proves the desired inequality (19)
and completes the proof of Proposition 4.2.

5. Higher order control of the WP metric


In this section, we show how to control higher order derivatives of the
WP metric. If R is the curvature tensor of a Riemannian metric, we define
k∇Rx k as in the previous section and let
k∇2 Rx k = sup k∇2v1 ,v2 Rx (v3 , v4 )v5 k,
v1 ,...,v5 ∈Tx1 N

where ∇2 R is the second covariant derivative of the curvature tensor:



∇2X,Y R = ∇X ∇Y − ∇Y ∇X − ∇[X,Y ] R.
The main result in this section is
Proposition 5.1. There exist C, β1 > 0 such that for any X0 ∈ T , the W P
curvature tensor RW P satisfies:
max{k(∇RW P )X0 k, k(∇2 RW P )X0 k} ≤ Cρ−β
0
1

where ρ0 = ρ0 (X0 ) is the distance from X0 to the singular locus ∂T .


We remark that similar bounds on higher derivatives of the W P curvature
tensor can also be obtained using the methods in this section. The prelim-
inary step in proving this proposition is to establish comparisons between
the WP and Teichmüller metrics.
48 K. BURNS, H. MASUR AND A. WILKINSON

5.1. Comparison of Weil-Petersson and Teichmüller metrics. The


goal of this section is to prove the following proposition, which compares
Teichmüller and Weil-Petersson lengths of vectors. Much of the material for
this section is taken from [36] and [23]. There are similar computations to
be found in [21].
For a given Riemann surface X, recall that ℓ(X) denotes the length of
the shortest simple closed curve in the hyperbolic metric.
Proposition 5.2. There exists C > 0 such that for any X ∈ Teich(S) and
any tangent vector [µ] ∈ TX Teich(S), we have
kµkW P ≥ Cℓ(X)1/2 kµkT .
The proposition follows from the following lemma.
Lemma 5.3. For any holomorphic quadratic differential φ ∈ Q(X),
kφkW P ≤ Cℓ(X)−1/2 kφkT .
We now give the proof of the proposition assuming Lemma 5.3.
Proof. We have
kµkT = sup |φ(µ)|
kφkT =1
≤ sup kφkW P kµkW P
kφkT =1

≤ sup Cℓ(X)−1/2 kφkT kµkW P = Cℓ(X)−1/2 kµkW P .


kφkT =1

We now begin the discussion leading to the proof of Lemma 5.3. We again
use notation x ≍ y to indicate that the two quantities x, y are comparable
up to multiplicative constants that depend only on g, n.
5.1.1. Blowing up nodes and (s, t) coordinates. Now suppose [f : S → Xσ ]
is a marked noded Riemann surface in Tσ , where the simplex σ has p ≤
3g − 3 + n nodes.
For each small η consider the neighborhood Ω(σ, η) of [f : S → Xσ ] in
T . Now let Γσ be the abelian group generated by the Dehn twists about
the curves in σ, and let Vσ = Ω(σ, η)/Γσ . Let Pσ : Vσ → Ω(σ, η) denote the
projection map. We make our computations using coordinates in Vσ which
we now describe. In what follows the marking is irrelevant.
Separating the nodes of Xσ into pairs of punctures, denoted Pi , Qi , pro-
duces the Riemann surface X̂σ . Choose conformal neighborhoods Vi = {zi :
0 < |zi | < 1} and Wi = {wi : 0 < |wi | < 1} of Pi and Qi . These may be
taken to be mutually disjoint. Choose a nonempty open set O on X̂σ disjoint
from ∪i (Vi ∪ Wi ). (If X̂σ is disconnected O consists of an open set in each
component). There exist Beltrami differentials ν1 , . . . , ν3g−3+n−p supported
in O whose equivalence classes form a basis for TX̂σ Tσ . This implies that for
THE WEIL-PETERSSON GEODESIC FLOW 49

any Ŷσ sufficiently close to X̂σ , there is a tuple s(Ŷσ ) = (s1 , . . . , sn3g−3+n−p )
of complex numbers close to 0 and a quasiconformal map h : X̂σ → Ŷσ such
that the dilatation µ(h) of h satisfies
3g−3+n−p
X
µ(h) = s i νi .
i=1

This gives a parametrization {X̂σ (s)} of surfaces in a neighborhood of X̂σ


that lie in ∂T by a neighborhood D 3g−3+n−p of 0 in C3g−3+n−p .
Since the map h (on each component) is conformal in Ui ∪ Vi the coor-
dinates zi , wi are local holomorphic coordinates in neighborhoods Vi , Wi of
the punctures on each X̂σ (s).
We now parametrize Vσ by pairs (s, t) where s ∈ D 3g−3+n−p and t =
(t1 , . . . , tp ) are small complex numbers. For each surface X̂s and for each 1 ≤
i ≤ p remove the disc of radius |ti |1/2 from each of Vi and Wi and then glue
zi to ti /wi , identifying the circle |zi | = |ti |1/2 with the circle |wi | = |ti |1/2 .
We denote the resulting surface by X(s, t); this gives a parametrization
{X(s, t) : s ∈ D 3g−3+n−p , t ∈ D p }
of Vσ . We note that in this notation we have X(s, 0) = X̂σ (s); that is, if all
ti = 0, then there are no discs to remove.
For each ti we can form the annulus
Ati = {zi : |ti | ≤ |zi | ≤ 1} = {wi : |ti | ≤ |wi | ≤ 1}
as a subset of X(s, t).

5.1.2. Metric in coordinates. At points X(s, t) with ti 6= 0 for all i, we may


regard
{∂/∂t1 , . . . , ∂/∂tp , ∂/∂s1 , . . . , ∂/∂s3g−3+n−p }
as a basis for the tangent space to Vσ and therefore as (equivalence classes
of) Beltrami differentials on these surfaces. There exists a dual basis
φ1 , . . . , φp , φp+1 , . . . , φ3g−3+n
of quadratic differentials for the cotangent space at X(s, t) whose expressions
in terms of the coordinates zi and wi on the surface X(s, t) take the following
form:
dz 2
φj = 2i (ci,j (s, t) + fi,j (zi , s, t) + gi,j (wi , s, t)),
zi
where
(a) fi,j (zi , s, t) is holomorphic in |zi | < 1 and t and real analytic in s;
(b) gi,j (wi , s, t) is holomorphic in |wi | < 1 and t and real analytic in s;
(c) fi,j (0, s, t) = gi,j (0, s, t) = 0;
(d) if j > p or j ≤ p and i 6= j, then ci,j (s, t) = 0; and
(e) if j ≤ p, then cj,j (s, t) is bounded above and below away from 0.
50 K. BURNS, H. MASUR AND A. WILKINSON

For each j ≤ p the term cj,j (s, t) is called the residue of φj . It is the
coefficient of the 1/zj2 term in the Laurent expansion of φj in Atj . As we
will see below, it is the dominant term in the calculation of integrals in Atj .
Now we prove Lemma 5.3.

Proof. We first note that the hyperbolic metric ρ(zi )|dzi | on X(s, t) re-
stricted to the annulus Ati satisfies
1
(24) ρ(zi ) ≍ .
|zi |(log |zi |)
The hyperbolic length of the short curve αi is given by
Z
|dzi | 1
(25) ℓαi (X) ≍ ≍− .
|zi |=|ti |1/2 −|zi | log |zi | log |ti |
Now suppose φ ∈ QD(X(s, t)) satisfies
Z
kφkT = |φ| = 1.
X(s,t)

Write
p
X 3g−3+n
X
φ= aj φj + bj φj ,
j=1 j=p+1

and set
3g−3+n
X
ψ= bj φj .
j=p+1

We note that
dzi2 dw2
2 = 2i
zi wi
under the change of variables wi = ti /zi . This means that we can use zi co-
ordinates on Ati when calculating the integral of fi,j and the wi coordinates
for calculating the integral of gi,j . Now (a) and (c) say that fi,j /zi2 has at
most a simple pole at zi = 0, so
Z
fi,j 2

z 2 |dzi | = O(1),
At i i

and similarly for gi,j . Combining this with (d), for j 6= i we obtain
Z
(26) |φj | = O(1).
At i

On the other hand it follows from (e) that for i ≤ p, we have


Z
(27) |φi | ≍ − log |ti |.
At i
THE WEIL-PETERSSON GEODESIC FLOW 51

Each φj has L1 norm bounded below away from 0 in the complement of


the union of the Ati . This implies that |bj | = O(1), and
(28) 1 = kφkT ≍ max{kψkT , kai φi kT }.
i

The residue of φj is 0 for j ≥ p + 1. Then the expansion (24) for the


hyperbolic metric in Ati shows that
Z Z 1
|ψ|2
2
≍ (log |zi |)2 |dzi |2 = O(1).
At i ρ |ti |

On the other hand, we also have


Z
|ψ| = O(1),
At i

which shows that


kψkW P ≍ kψkT .
Thus if the maximum in (28) is given by kψkT we are done.
Therefore we are reduced to the case that the maximum is given by
kai φi kT , for some i ≤ p. Then from (27) we see that
1
|ai | ≍ .
− log |ti |
Thus
Z Z
kφi k2W P ≍ |φi |2 /ρ2 ≍ |dzi |2 (log |zi |)2 /|zi |2
At i At i
Z 1
= dr(log r)2 /r ≍ (− log |ti |)3 ,
|ti |

which by (25) and the above estimate for |ai | gives the desired estimate:
kai φi kW P ≤ C(− log |ti |)1/2 ≤ ℓ(X)−1/2 kφkT .
This completes the proof of Lemma 5.3. ⋄

5.2. Estimates on the WP metric in special coordinates. Follow-


ing [21], we introduce coordinates on Teich(S) in which we can bound the
derivatives of the WP metric. In the section we denote by N the complex
dimension of Teich(S). Let ∆N denote the Euclidean unit polydisk in CN .
We will denote by z = (z1 , . . . , zN ) an element of ∆N , where zk is a complex
coordinate, and by xk = Re(zk ), yk = Im(zk ) the real coordinates. Let ei
be the vector field ∂/∂xi , for 1 ≤ i ≤ N , and ∂/∂yi−N , for N + 1 < i ≤ 2N .
The following lemma is proved in [21] and follows from Nehari’s bound
and the fact that the Teichmüller and Kobayashi metrics agree on the image
of the Bers embedding.
52 K. BURNS, H. MASUR AND A. WILKINSON

Lemma 5.4. [21, cf. Theorem 2.2 and Proof of Theorem 8.2] There exists
C ≥ 1 such that for any X0 ∈ Teich(S), there is a holomorphic embedding
ψ = ψX0 : ∆N → Teich(S), sending 0 ∈ ∆N to X0 and such that for every
v ∈ T ∆N , we have:
1
kvk ≤ kDψ(v)kT ≤ Ckvk
C
where k·k is the Euclidean norm on ∆N and k·kT is the Teichmüller Finsler
norm on Teich(S).
Fix a point X0 ∈ Teich(S), and let ψ = ψX0 be the holomorphic em-
bedding given by this lemma. Since the metric gW P on Teich(S) is Kähler
with respect to the 2-form ωW P , and ψ is holomorphic, it follows that the
pullback metric ψ ∗ gW P on ∆N is Kähler with respect to the pullback form
ψ ∗ ωW P and the standard almost complex structure on ∆N . For z ∈ ∆N ,
denote by G(z) the matrix with entries
Gij (z) = (ψ ∗ gW P )z (ei , ej ).
The main content of this section is the proof of the following proposition.
Proposition 5.5. There exists C > 0 such that, for every X0 ∈ Teich(S),
and every z ∈ ∆N with kzk ≤ 1/2 we have:
(1) kG−1 (z)k ≤ Cℓ(X0 )−1 , and
(2) for any i, j ∈ {1, . . . , 2N } and any k ≥ 0, we have

∂ k Gi,j
sup (z) ≤ Ck! .

k ∂ξ1 · · · ∂ξk
(ξ1 ,...,ξk )∈{x1 ,...,xN ,y1 ,...,yN }

We will use Proposition 5.5 to bound the covariant derivatives of the W P


curvature in terms of the the distance to the singular strata. An immediate
corollary that we record now is:
Corollary 5.6. For each X0 ∈ T eich(S), the image of the polydisk {|z| ≤
1/2} ⊂ ∆N under the embedding ψ = ψX0 given by Lemma 5.4 contains the
WP metric ball around X0 of radius (2C)−1 ℓ(X0 )1/2 , where C is given by
Proposition 5.5.
Proof. We prove Proposition 5.5. Part (1) is just Propositon 5.2 and Lemma 5.4.
The proof of part (2) uses in a crucial way results of McMullen in [21].
Using the embedding ψ, we define an embedding Ψ : ∆N × ∆N → QF (S)
by
Ψ(z, w) = (ψ(z), ψ(w)).
Since ψ is holomorphic and X 7→ X is antiholomorphic, the map Ψ is
holomorphic. Note that the image of the antidiagonal A = {(z, z) : z ∈ ∆N }
under Ψ lies in the Fuchsian locus F (S) ⊂ QF (S). Denote by α : ∆n →
∆N × ∆N the antidiagonal embedding α(z) = (z, z), and by α̂ : Teich(S) →
QF (S) the antidiagonal embedding α̂(X) = (X, X). Then we have the
following commutative diagram:
THE WEIL-PETERSSON GEODESIC FLOW 53

∆N N
99 × ∆MMM
ss MMM
α sss
ss MMΨ
ss MMM
sss MMM
sss M&&
∆N K K QF (S)
88
KK
KKK ψ qqqq
KKK α̂ qqq
KKK
qqqqq
KK
%% qqq
Teich(S)
Note that the maps α and α̂ are not holomorphic, although their deriva-
tives are bounded in the Euclidean and Teichmüller metrics, respectively.
Since Teich(S) and QF (S) are complex manifolds, so are their cotan-
gent bundles T ∗ Teich(S) and T ∗ QF (S), and T ∗ QF (S) = T ∗ Teich(S) ⊕
T ∗ Teich(S). Fixing Z ∈ Teich(S) we define a map τ : QF (S) → T ∗ Teich(S)
by:
τ (X, Y ) = σQF (X, Y ) − σQF (X, Z).
Since T ∗ Teich(S) embeds as the first factor in T ∗ QF (S), we may regard τ
as a 1-form on QF (S). McMullen proves that τ has the following properties:
(1) τ is holomorphic.
(2) the 1-form θ = −α̂∗ τ on Teich(S) is a primitive for the WP Kähler
form:
d(iθ) = ωW P .
(3) For any (X, Y ) ∈ Teich(S), the covector τ (X, Y ) is bounded in the
Teichmüller Finsler norm on Teich(S):
kτ (X, Y )kT = sup{|τ (X, Y )(µ)| : µ ∈ TX Teich(S), kµkT = 1} ≤ C,
where C is independent of (X, Y ).
Pulling the holomorphic 1-form τ back to ∆N × ∆N , we thus obtain a
holomorphic 1-form κ = Ψ∗ τ .
Lemma 5.7. κ is bounded in the Euclidean metric on ∆N × ∆N .
Proof. τ is bounded in the Teichmüller metric, and the Euclidean metric is
comparable to the Ψ-pullback of the Teichmüller metric, by Lemma 5.4. ⋄

Lemma 5.8. The holomorphic 2-form Ω = d(i κ) on ∆N × ∆N satisfies


α∗ Ω = ψ ∗ ωW P , which is the Kähler 2-form for the pullback metric ψ ∗ gW P .
Proof. This follows from the commutativity of the diagram above. ⋄

We now finish the proof of Proposition 5.5. In complex coordinates


(z1 , . . . , zn , w1 , . . . , wN ) on ∆N × ∆N one can write
N
X
κ= ai dzi ,
i=1
54 K. BURNS, H. MASUR AND A. WILKINSON

where ai : ∆N × ∆N → C are bounded holomorphic functions. Now


N
X ∂aj ∂aj
Ω = d(i κ) = i dzk ∧ dzj + i dwk ∧ dzj .
∂zk ∂wk
j,k=1

Finally
N
X ∂aj ∂aj
α∗ Ω = i dzk ∧ dzj + i dz k ∧ dzj .
∂zk ∂z k
j,k=1

The Euclidean coefficients of the Kähler metric ψ ∗ gW P are hence linear


combinations, with bounded coefficients, of ∂aj /∂zk and ∂aj /∂z k , which in
turn are pullbacks of the complex partial derivatives ∂aj /∂zk and ∂aj /∂wk .
Since the aj are bounded holomorphic functions, Cauchy’s Theorem implies
that the derivatives ∂aj /∂zk and ∂aj /∂wk are bounded for k(z, w)k < 1/2;
it follows that the (real) partial derivatives of ai are bounded for kzk < 1/2.
The same applies to all higher order partial derivatives (where the kth order
derivatives get a bound of order k!). This proves (2). ⋄

5.3. Proof of Proposition 5.1. Fix X0 ∈ Teich(S) and local coordinates


ψ = ψX0 as in the previous section. For z ∈ ∆N , let G(z) = GX0 (z) =
(Gij (z)) be the matrix for the pullback metric ψ ∗ gW P , and let Gij (z) =
G(z)−1 ij .
The curvature tensor for G can be calculated in these Euclidean coordi-
nates using the Christoffel symbols:
 
m 1 km ∂ ∂ ∂
Γij = G Gkj + j Gik − k Gij
2 ∂xi ∂x ∂x
and the Riemannian curvature tensor coefficients:
ℓ ∂ ℓ ∂ ℓ
Rijk = Γ − Γ + Γℓjs Γsik − Γℓks Γsij ,
∂xj ik ∂xk ij
where Gij At each z ∈ ∆N the curvature tensor P ℓ Rz for the WP pullback
metric G is defined by setting Rz (ei , ej )ek = ℓ Rijk eℓ . It is a function from
N 3N N
∆ × C to C . Finally, the covariant tensor defined by
X
ν
Ri,j,k,ℓ = Gℓν Rijk ,
ν

satisfies hRz (ei , ej )ek , eℓ i = Rijkℓ .


The formula for the covariant derivative of a covariant tensor Tb1 ...bs is
given by:

(∇c T )b1 ...bs = Tb ...b − Γd b1 c Td...bs − · · · − Γd bs c Tb1 ...bs−1 d .
∂xc 1 s
THE WEIL-PETERSSON GEODESIC FLOW 55

Since kDψk and kDψ −1 k are bounded, the quantities k(∇RW P )ψ(z) k and
k(∇2 RW P )ψ(z) k can therefore be bounded by a (universal) polynomial func-
tion of the quantities |Gij (z)|, |Gij (z)| and

∂ k Gi,j

∂ξ1 · · · ∂ξ (z) ,
k

for k = 1, . . . , 4. But Proposition 5.5 implies that the entries Gij (z) are
O(ℓ(X0 )−1 ) and the entries Gij (z) and their first k derivatives are O(1); the
conclusion of Proposition 5.1 then follows.

6. General criteria for ergodicity of the geodesic flow


Let M be an open, contractible Riemannian manifold, negatively curved,
possibly incomplete. Let Γ be a group that acts freely and properly dis-
continuously on M by isometries, and denote by N the quotient manifold
N = M/Γ. We denote by d both the path metric on M and the quotient
metric on N , which is just the path metric for the induced Riemannian
metric on N . The quotient map p : M → N is a covering map and a local
isometry.
Recall that the completion X of a metric space (X, d) is the set of all
Cauchy sequences hxn i in X modulo the equivalence relation:
hxn i ∼ hyn i ⇐⇒ lim d(xn , yn ) = 0,
n→∞

with the induced metric


d(hxn i, hyn i) = lim d(xn , yn ).
n→∞

Let M be the metric completion of M and let N be the completion of N .


Let ∂N = N \ N . We will use d to denote the metric on all of these spaces.
Consider the following additional assumptions on M and N :
I. M is a geodesically convex: for every p, p′ ∈ M there is a unique
geodesic segment in M connecting p to p′ .
II. N is compact.
III. ∂N is volumetrically cusplike: there exist constants C > 1 and ν > 0
such that:
Vol ({p ∈ N : d(p, ∂N ) < ρ}) ≤ Cρ2+ν ,
for every ρ > 0.
Note that if II.– III. hold, then N has finite volume. In this case, we denote
by m the Riemannian volume on N , normalized so that m(N ) = 1.
A metric space X is CAT (0) if it is a geodesic space and and every
geodesic triangle in X satisfies the CAT (0) inequality with the comparison
Euclidean triangle (see [7, p.159]).
Lemma 6.1. If I. holds, then M and M are both CAT (0) spaces.
56 K. BURNS, H. MASUR AND A. WILKINSON

Proof. The fact that M is CAT (0) follows from [7, Theorem II.1A.6] and
Alexandrov’s Patchwork [7, Proposition II.4.9]. The metric completion of a
CAT (0) space is CAT (0), by [7, Corollary II.3.11]. ⋄

Proposition 6.2 (The flow is a.e. defined for all time). If I.–III. hold, then
for almost every v ∈ T 1 M , there exists an infinite geodesic (necessarily
unique) tangent to v.
Before proving this we state and prove another lemma that will be useful
later as well. Let
Uρ = {v ∈ T 1 N : d(π(v), ∂N ) < ρ},
and let S + (ρ) be the set of all tangent vectors that flow into Uρ in some
forward time 0 ≤ t ≤ 1.
Lemma 6.3. If I.–III. hold, then
m(S + (ρ)) = O(ρ1+η ).
Proof. We look at the “shell” Sk+ (ρ) of vectors v that flow into Uρ in time
between kρ and (k + 1)ρ. Any two vectors in this shell arrive in U2ρ after
time (k + 1)ρ. Volume-preservation of the flow implies that the the volume
of Sk+ (ρ) is at most the volume of U2ρ , which is O(ρ2+η ), by assumption III.
The set S + (ρ) is contained in a union of the shells S0+ (ρ), . . . , Sm
+ (ρ), where
−1 −1
m is O(ρ ). It follows that the volume of S (ρ) is O(ρ ρ ) = O(ρ1+η ).
+ 2+η

Proof of Proposition 6.2. The set of vectors such that the flow is not defined
for some 0 ≤ t ≤ 1 is contained in S + (ρ) for all ρ > 0. By Lemma 6.3 this
set has measure 0. The set En of vectors such that the flow is not defined
for some time 0 ≤ t ≤ n is the union of En−1 and the set of vectors that
flow into En−1 within time 1. Inductively En−1 has measure 0, and since
the flow is measure preserving, so does En . It follows that the set of vectors
for which the flow is defined for all time has full measure. ⋄

The next proposition will be used later to prove ergodicity of the geodesic
flow on T 1 N , under further assumptions on the metric.
Suppose that v ∈ T M determines an infinite geodesic ray γv : [0, ∞) → M
tangent to v at 0. Since M is a CAT (0) space, the functions bsv,t : M → R
defined by
bsv,t (y) = d(y, γv (t)) − t
converge uniformly on compact sets as t → ∞ to a function bsv : M → R,
called a (stable) Busemann function [7, Lemma II.8.18]. For a fixed v, the
Busemann function bsv is clearly Lipschitz continuous, with Lipschitz norm
1. If we assume that I. holds, then we can say more.
THE WEIL-PETERSSON GEODESIC FLOW 57

Proposition 6.4. Assume that I. holds. For any v that determines an


infinite geodesic ray γv , the function bsv is convex and C 1 , and k grad bsv k ≡ 1.
For every y ∈ T 1 M , the unit vector
wvs (y) := − grad bsv (y)
defines an infinite geodesic ray γwvs (y) : [0, ∞) → M tangent to wv (y) at 0
with the property that
d(γv (t), γwvs (y) (t)) ≤ d(γv (0), y),
for all t ≥ 0.
Proof. Since γv is an infinite ray, and M is a geodesically convex Riemann-
ian manifold, the functions bsv,t are convex, C 1 and have the property that
k grad bsv,t (y)k = 1, for every y ∈ M . Since M is nonpositively curved, and
bsv,t converges uniformly on compact sets in M to bsv , the desired proper-
ties of C 1 smoothness of bsv , convexity and k grad bsv k ≡ 1 follow from [4,
Lemma 3.4, and the following Remark]. The final conclusion follows from
[7, Proposition II.8.2]. ⋄

Suppose that v ∈ T 1 M determines a geodesic ray whose projection to


N is forward recurrent. Proposition 6.4 implies that for each t ∈ R, the
set Hvs (t) := (bsv )−1 (t) is a connected, codimension-1, C 1 submanifold of M ,
called a stable horosphere. For a such a v, we define:
W s (v) = {wvs (y) : y ∈ Hvs (0)}.
The set of basepoints π(W s (v)) in M is the horosphere Hvs (0), and W s (v) is
a continuous, codimension-1 submanifold of T 1 M . Similarly, if γv projects
to a backward recurrent geodesic ray in N , we define the unstable Busemann
function
buv (y) = lim d(y, γv (−t)) − t
t→∞
and unstable manifold
W u (v) = {wvu (y) : y ∈ Hvu (0)},
where wvu (y) = − grad buv (y), and Hvu (t) := (buv )−1 (t) is the unstable horo-
sphere at level t determined by v. Our next proposition justifies the termi-
nology “stable and unstable manifolds” for W s (v) and W u (v). The results
stated up to this point all hold true when M is nonpositively curved, but
the following proposition uses the negative curvature assumption on M in
an essential way.
We say that a geodesic ray γ : [0, ∞) → N is (forward) recurrent if
the tangent vector γ̇(0) is an accumulation point for the tangent vectors
{γ̇(t) : t > 0}. We similarly define backward recurrence for a geodesic ray
γ : (−∞, 0] → N . An infinite geodesic is recurrent if it is both forward and
backward recurrent. Under assumptions I.-III., Proposition 6.2 and Poincaré
recurrence imply that almost every v ∈ T 1 N determines an infinite recurrent
geodesic γv : R → N with γ̇v (0) = v.
58 K. BURNS, H. MASUR AND A. WILKINSON

Proposition 6.5 (Contraction of horospheres). Assume I.–III. Let v ∈


Tx M be tangent to an infinite geodesic ray γv whose projection to N is
forward recurrent. Let y ∈ M be any other point, and let w = wvs (y) ∈ Ty M .
Then w is tangent to an infinite geodesic ray γw : [0, ∞) → M , and
lim d(γv (t), γw (t + bsv (y))) = 0;
t→∞
moreover,
lim dSas (ϕt (v), ϕt+bsv (y) (w)) = 0.
t→∞
In particular, if γv projects to a forward recurrent geodesic ray in N , then
for every t > 0, ϕt (W s (v)) = W s (ϕt (v)), and for every w ∈ W s (v), we have
lim dSas (ϕt (v), ϕt (w)) = 0.
t→∞
Similarly, if v is tangent to a backward ray γv : (−∞, 0] → M whose projec-
tion is recurrent, then w = wvu (w) is tangent to a backward ray γw : (−∞, 0] →
M , and
lim dSas (ϕt (v), ϕt+bsv (y) (w)) = 0.
t→−∞
In particular, for every w ∈ W u (v), we have
lim d(ϕt (v), ϕt (w)) = 0.
t→−∞

Before beginning the proof we remark that in [9] a property called non-
refraction was proved for the WP metric. Using that result, a short proof
of the above proposition was given in the WP case in [6]. Since we are in
a general Riemannian setting in this section, we prefer not to make this
additional assuption of nonrefraction.
Proof. Let γv : [0, ∞) → M be an infinite geodesic ray whose projection
to N is recurrent, and let x = γv (0) be the footpoint of v. Suppose that
x′ ∈ M is another point, and let v ′ = wvs (x′ ). Since M is CAT (0), the
distance d(γv (t), γv′ (t)) is a convex function of t; since it is bounded, it must
be nonincreasing, and hence bounded above for all t by d(x, x′ ). We claim
that if d(x, x′ ) < d(x, ∂M ), then the image of γv′ must lie entirely in M .
Since the projection of γv to N is recurrent, there exist sequences gn ∈ Γ
and tn → ∞ such that
d(x, gn γv (tn )) < d(x, ∂M ) − d(x, x′ )
≤ d(x, ∂M ) − d(γv (tn ), γv′ (tn )),
= d(x, ∂M ) − d(gn γv (tn ), gn γv′ (tn )),
which implies that d(x, gn γv′ (tn )) < d(x, ∂M ). Hence gn γv′ (tn ) ∈ M , and
so γv′ (tn ) ∈ M ; geodesic convexity of M implies that γv′ [0, tn ] ⊂ M , for all
n, which proves the claim.
Now a standard ruled surface argument using geodesic convexity and the
negative curvature of M (see e.g. [6, Theorem 4.1], where it is proved in the
WP context) shows that for every γv that projects to a recurrent geodesic
ray in N , and any y ∈ M with the property that γwvs (y) [0, ∞) ⊂ M , the
THE WEIL-PETERSSON GEODESIC FLOW 59

distance d(γwvs (y) (t), γv [0, ∞)) is strictly decreasing in t and tends to 0 as
t → ∞. (Alternately, one can show this using Jacobi fields). What is more,
this convergence takes place in the tangent bundle:
lim dSas (γ̇wvs (y) (t), γ̇v [0, ∞)) = 0.
t→∞

Now suppose that y ∈ M is an arbitrary point. Connect y to x = γv (0)


by a geodesic arc σ in M . Fix ǫ0 > 0 such that d(x, ∂M ) < ǫ0 . We claim
that if x′ is any point on σ that satisfies
lim d(γwvs (x′ ) (t), γv [0, ∞)) = 0,
t→∞

then for any point y ′ on σ such that d(x′ , y ′ ) < ǫ0 /3:


lim d(γwvs (y′ ) (t), γv [0, ∞)) = 0.
t→∞

From the claim it follows that limt→∞ dSas (γ̇wvs (y) (t), γ̇v [0, ∞)) = 0.
To prove the claim, suppose that x′ and y ′ are given. Since the distance
d((γwvs (x′ ) (t), γwvs (y′ ) (t)) is bounded for all t > 0 and convex, it is nonincreas-
ing, and hence bounded above by ǫ0 /3, for all t > 0. If T > 0 is sufficiently
large, then the distance from γwv (x′ ) (t) to γv is less than ǫ0 /3 for all t > T .
Since γv projects to a recurrent ray in N , there exist gn ∈ Γ and tn → ∞
such that d(γv (t), gx) < ǫ0 /3. It follows that γwvs (y′ ) (tn ) ∈ M for all tn > T ,
which implies that γwv (y′ ) [0, ∞) ⊂ M . The conclusion follows.
A simple application of the triangle inequality shows that the property
limt→∞ d(γwvs (y) (t), γv [0, ∞)) = 0 implies that
lim d(γv (t), γwvs (y) (t + bsv (y))) = 0.
t→∞

Since limt→∞ dSas (γ̇wvs (y) (t), γ̇v [0, ∞)) = 0 for every y ∈ M , we conclude
that
lim dSas (ϕt (v), ϕt+bsv (y) (wvs (y)) = 0.
t→∞

Our final set of assumptions will be used to establish ergodicity of the


geodesic flow. We assume there exist constants C > 1 and β > 0 such that:
IV. N has controlled curvature: for all x ∈ N , the curvature tensor R
satisfies:
max{kRx k, k∇Rx k, k∇2 Rx k} ≤ Cd(x, ∂N )−β .
V. N has controlled injectivity radius: for every x ∈ N ,
inj(x) ≥ C −1 d(x, ∂N )β .
VI. The derivative of the geodesic flow is controlled: for every infinite
geodesic γ in N and every t ∈ [0, 1]:
kDγ̇(0) ϕt k ≤ Cd (γ ([−t, t]) , ∂N )−β ;
60 K. BURNS, H. MASUR AND A. WILKINSON

Theorem 6.6. Under assumptions I.-VI., the geodesic flow ϕt on T 1 N is


nonuniformly hyperbolic and ergodic. The entropy h(ϕ1 ) of ϕ1 is positive and
finite, equal to the sum of the positive Lyapunov exponents of ϕt , counted
with multiplicity.

Remark: It seems that Assumption II. (compactness of N ) can be relaxed


to the assumption that N has finite diameter, but we have not verified all
of the details.
The proof of Theorem 6.6 proceeds in several steps. The first is to estab-
lish nonuniform hyperbolicity. To this end, we need the following lemma.
Lemma 6.7. Let ϕ1 be the time-1 map of the geodesic flow, which is defined
m-almost everywhere in T 1 N . Then
Z
log+ kDϕ1 k dm < ∞.
T 1N

Proof. We consider the set S + (1/n) of vectors which flow into the set of
points of distance between 1/(n + 1) and 1/n from S for some 0 ≤ t ≤ 1.
Lemma 6.3 implies that
m(S + (1/n)) = O((1/n)1+η ).
On S + (1/n) we have
log+ kDϕ1 k = O(log n).
Therefore Z
log+ kDϕ1 k dm = O(log n/n1+η ).
S + (1/n)
Summing over n we obtain the conclusion. ⋄

Since log kDϕ1 k is integrable, Oseledec’s theorem implies that for m-


almost every v ∈ T 1 N , there exist real numbers λ1 (v) ≤ λ2 (v) ≤ · · · ≤
L2n−1
λ2n−1 (v) and a splitting Tv T 1 N = i=1 Ei (v) such that for every nonzero
vector ξ ∈ Ei (v):
1
lim log kDv ϕt (ξ)k = λi (v).
t→±∞ t
The numbers λi (v) are called the Lyapunov exponents of ϕt at v, and Ei (v)
the Lyapunov subspaces. The stable and unstable Lyapunov subspaces at v
are E s (v) := ⊕λi (v)<0 Ei (v) and E u (v) := ⊕λi (v)>0 Ei (v), respectively. The
functions v 7→ λi (v) and v 7→ Ei (v) are measurable. Since the generating
vector field ϕ̇ is preserved by Dϕt , it follows that λn (v) = 0. Moreover, since
the orthocomplement ϕ̇⊥ is Dϕt -invariant, and the restriction of Dϕt pre-
serves a natural symplectic form, the Lyapunov exponents of ϕt are paired:
λi (v) = −λ2n−i (v), for i = 1, . . . , n − 1.
Recall that the geodesic flow ϕt is nonuniformly hyperbolic if λn−1 (v) < 0
for almost every v; equivalently, Tv T 1 N = E s (v)⊕Rϕ̇(v)⊕E u (v), for almost
every v. We have:
THE WEIL-PETERSSON GEODESIC FLOW 61

Proposition 6.8 (Nonuniform hyperbolicity). Under assumptions I.-VI.,


the geodesic flow is nonuniformly hyperbolic. That is, there exists a full
measure, ϕt -invariant subset Λ0 ⊂ T 1 N and a measurable, Dϕt -invariant
splitting of the tangent bundle:
TΛ0 (T 1 N ) = E s ⊕ E 0 ⊕ E u
such that, for every v ∈ Λ0 :
(1) E 0 (v) is tangent to the orbits of the flow: E 0 (v) = Rϕ̇(v);
(2) for every ξ u ∈ E u (v), ξ s ∈ E u (v):
1 1
lim log kDv ϕt (ξ u )k > 0, and lim log kDv ϕt (ξ s )k < 0,
t→±∞ t t→±∞ t

and the limits are finite.


Moreover, there exist continuous, disjoint cone fields C u and C s over T 1 N ,
spanning the space of perpendicular Jacobi fields in T T 1 N , such that for
every v ∈ Λ0 and all t > 0:
E u (v) ⊂ C u (v), E s (v) ⊂ C s (v),
Dϕt (C u (v)) ⊂ C u (ϕt (v)), and Dϕ−t (C s (v)) ⊂ C s (ϕ−t (v)).
For each v ∈ Λ0 , E u (v) is spanned by the unstable perpendicular Jacobi
fields at v, and E s (v) is spanned by the stable perpendicular Jacobi fields at
v.
Proof. This proof is standard for compact negatively-curved manifolds (see,
e.g., [16, Section 17.6]). A few modifications adapt the proof to this setting.
Define positive continuous functions k, k : N → R>0 by
p p
k(x) = sup −hR(v, w)w, vi, k(x) = inf −hR(v, w)w, vi,
v,w∈Tx1 N v,w∈Tx1 N

and δ : T 1 N → R>0 by
k 2 (π(v))
δ(v) = 2 .
2(1 + k (π(v)))
For v ∈ T 1 N , we identify Tv T 1 N ∼ = (Tπ (v)N )2 in the standard way, and
let ϕ̇⊥ (v) = {(w0 , w1 ) : hw0 , vi = hw1 , vi = 0} be the orthocomplement of
ϕ̇; it is a 2n − 2-dimensional subspace of Tv T 1 N and is naturally identified
with the space of perpendicular Jacobi fields to γv . We define the conefields
C u , C s by
C u (v) = {(w0 , w1 ) ∈ ϕ̇⊥ (v) : hw0 , w1 i ≥ δ2 (v)k(w0 , w1 )k2Sas }
and
C s (v) = {(w0 , w1 ) ∈ ϕ̇⊥ (v) : hw0 , w1 i ≤ −δ2 (v)k(w0 , w1 )k2Sas }.
By our choice of δ, these conefields satisfy the strict invariance conditions
Dϕt (C u \ {0}) ⊂ int(C u ), and Dϕ−t (C s \ {0}) ⊂ int(C s ),
62 K. BURNS, H. MASUR AND A. WILKINSON

for all t > 0, wherever these objects are defined; the proof follows closely
the proof of [16, Lemma 17.6.1].
These conefields are symplectic in the sense of [41], as we now explain.
The restriction of the symplectic form on T T N defined by
ω((v0 , v1 ), (w0 , w1 )) = hv0 , w1 i − hv1 , w0 i.
to the subbundle ϕ̇⊥ of perpendicular Jacobi fields in T T 1 N is also symplec-
tic, and the derivative of the time-t map Dϕt is a linear symplectomorphism
when restricted to ϕ̇⊥ .
Let α : T 1 N → R>0 be the positive solution to α(1 + α2 )−1 = δ. For
v ∈ T 1 N , let
L1 (v) = {(w0 , α(v)w0 ) : w0 ∈ Tπ(v) N, hv, w0 i = 0},
and let
L2 (v) = {(α(v)w1 , w1 ) : w1 ∈ Tπ(v) N, hv, w1 i = 0}.
Observe that L1 and L2 are transverse Lagrangian subbundles spanning ϕ̇⊥ .
One can uniquely express an arbitrary element (w0 , w1 ) ∈ ϕ̇⊥ as a sum of
elements in L1 and L2 , as follows:
w0 − αw1 w0 − αw1 w1 − αw0 w1 − αw0
(w0 , w1 ) = ( 2
,α 2
) + (α , )
1−α 1−α 1 − α2 1 − α2
Using this decomposition, we define a symplectic quadratic form KL1 ,L2 on
ϕ̇⊥ by:
 
w0 − αw1 w0 − αw1 w1 − αw0 w1 − αw0
KL1 ,L2 (w0 , w1 ) = ω ( , α ), (α , ) .
1 − α2 1 − α2 1 − α2 1 − α2
A straightforward calculation shows that

KL1 ,L2 (w0 , w1 ) = −αkw0 k2 − αkw1 k2 + (1 + α2 )hw0 , w1 i .
Hence KL1 ,L2 (w0 , w1 ) ≥ 0 if and only if
α
hw0 , w1 i ≥ (kw0 k2 + kw1 k2 ) = δk(w0 , w1 )kSas ;
1 + α2
in other words:
C u = {ξ ∈ ϕ̇⊥ : KL1 ,L2 (ξ) ≥ 0}.
We thus have established that Dϕt is a symplectomorphism of ϕ̇⊥ , the loga-
rithm of kDϕt k is integrable, and the conefield C u defined by the symplectic
quadratic form KL1 ,L2 is everywhere strictly invariant under Dϕt , for any
t > 0. It follows immediately from [15, Corollary 2.2] that Dϕt has exactly
n − 1 = dim(N ) − 1 positive Lyapunov exponents, m-almost everywhere.
Similarly, Dϕt has exactly n − 1 negative Lyapunov exponents almost ev-
erywhere. The subbundle E u is defined to be the direct sum of the positive
Lyapunov subbundles, and E s is the sum of the negative Lyapunov subbun-
dles. Invariance of the subbundles and recurrence give that E u ⊂ C u and
E s ⊂ C s almost everywhere. ⋄
THE WEIL-PETERSSON GEODESIC FLOW 63

Proposition 6.9 (Existence and absolute continuity of local stable man-


ifolds). Assume I.-VI. Let n = dim(N ), and let Λ0 ⊂ T 1 N be given by
Proposition 6.8. There exists a full volume, ϕt -invariant subset Λ1 ⊂ Λ0 ,
a measurable function r : Λ1 → R>0 and measurable families of C ∞ , (n −
1)-dimensional embedded disks Wloc s = {W s (v) : v ∈ Λ } and W u =
loc 1 loc
u
{Wloc (v) : v ∈ Λ1 } with the following properties. For each v ∈ Λ1 :
(1) Wlocs (v) is tangent to E s (v) at v, and W u (v) is tangent to E u (v) at
loc
v;
(2) for all t > 0,
s s u u
ϕt (Wloc (v)) ⊂ Wloc (ϕt (v)), and ϕ−t (Wloc (v)) ⊂ Wloc (ϕ−t (v))
s (v) if and only if d(v, w) < r(v) and
(3) w ∈ Wloc
lim dSas (ϕt (v), ϕt (w)) = 0;
t→∞
u (v) if and only if d(v, w) < r(v) and
(4) w ∈ Wloc
lim dSas (ϕt (v), ϕt (w)) = 0.
t→−∞

Moreover, for ∗ ∈ {s, u}, the family Wloc ∗ is absolutely continuous. In par-

ticular:
(5) if Z ⊂ T 1 N has volume m(Z) = 0, then for m-almost every v ∈ Λ1 ,
∗ (v) is a zero set in W ∗ (v) (with respect to the induced
the set Z∩Wloc loc
(n − 1)-dimensional Riemannian volume); and
(6) if D ⊂ T 1 N is any C 1 -embedded, n-dimensional open disk, and B ⊂
D has induced Riemannian volume zero in D, then m(Sat∗loc (B)) =
0, where
[
Sat∗loc (B) := ∗
Wloc (v).
∗ (v)∩B6=∅}
{v∈Λ1 : Wloc

The conclusions of Proposition 6.9 follow from the main results in [17].
To apply these results, it is necessary to verify a list of hypotheses, some
of a technical nature, concerning the C 3 properties of the Sasaki metric
and the geodesic flow. We defer the verification of these properties to the
next subsection and now show how Proposition 6.9 can be used to prove
ergodicity of ϕt . Properties (5) and (6) in Proposition 6.9 are the heart of
the matter in proving ergodicity and account for the technical hurdles we’ve
been jumping. Property (5) is a form of “leafwise absolute continuity” and
(6) is a form of “transverse absolute continuity.” Both are implied by the
condition of absolute continuity in [17].
Let Ω1 be the full measure set of v ∈ T 1 M such that γv projects to
a (forward and backward) recurrent geodesic in T N . For each v ∈ Ω1 ,
there exists a stable manifold W s (v) and an unstable manifold W u (v). For
v ∈ Ω1 and δ > 0, denote by W ∗ (v, δ) the connected component of W ∗ (v) ∩
BT 1 M (v, δ) containing v, where BT 1 M (v, δ) is the Sasaki ball of radius δ
in T 1 M centered at v. For v ′ = Dp(v) ∈ Dp(Ω1 ), and δ < inj(π(v ′ )) we
64 K. BURNS, H. MASUR AND A. WILKINSON

denote by W s (v ′ , δ) the projection Dp(W s (v, δ)); it is an (n−1)-dimensional


embedded disk.
Notice that, for every v ∈ Ω1 , if v ′ = Dp(v) belongs to the full measure
set Λ1 of Proposition 6.9, then the local stable manifold Wloc s (v ′ ) through v ′
s ′ ′
must coincide with W (v , r(v )), where r : Λ1 → R>0 is the function given
by Proposition 6.9. We will use this connection to prove the following.
Proposition 6.10 (Smoothness and absolute continuity of horospherical
laminations). Assume I.-VI. There is a full volume subset Ω2 ⊂ Ω1 such
that for ∗ ∈ {s, u} and for v ∈ Ω2 , the Busemann function b∗v : M → R
is C ∞ . The leaves of the lamination W ∗ = {W ∗ (v) : v ∈ Ω2 } are C ∞
submanifolds of T 1 M diffeomorphic to Rn−1 .
Let Λ2 = Dp(Ω2 ). The family of manifolds
{W ∗ (v, δ) : v ∈ Λ2 , δ < inj(π(v))}
has the following absolute continuity properties.
(1) if Z ⊂ T 1 N has volume m(Z) = 0, then for m-almost every v ∈ Λ2 ,
and every δ < inj(π(v)), the set Z ∩W ∗ (v, δ) is a zero set in W ∗ (v, δ)
(with respect to the induced (n − 1)-dimensional Riemannian vol-
ume); and
(5) if D ⊂ T 1 N is any smoothly embedded, n-dimensional open disk,
and B ⊂ D has induced Riemannian volume zero in D, then for any
δ < 21 inf v∈D inj(π(v)), we have m(Sat∗ (B, δ)) = 0, where
[
Sat∗ (B, δ) := W ∗ (v, δ).
{v∈Λ2 : W(v,δ)∩B6=∅}

Proof. We first show that W s (v) is a C ∞ submanifold of T 1 M , for almost


every v ∈ T 1 M . For any ǫ > 0 there exists a compact set ∆ǫ ⊂ Λ1 of
measure m(∆ǫ ) > 1 − ǫ such that the restriction of the function r from
Proposition 6.9 to ∆ǫ is continuous and bounded from below by a constant
rǫ > 0. Fix ǫ > 0, and let ∆sǫ ⊂ ∆ǫ be the set of vectors v ′ ∈ ∆ǫ such
that ϕkn (v ′ ) ∈ ∆ǫ for a sequence of integers kn → ∞. Poincaré recurrence
implies that m(∆ǫ \ ∆sǫ ) = 0.
Fix v ′ ∈ ∆sǫ ∩ Dp(Ω1 ). Let v ∈ Dp−1 (v ′ ) be an arbitrary lift of v ′ , and
let w ∈ W s (v). We show that W s (v) is C ∞ in a neighborhood of w; as
w is arbitrary, this implies that W s (v) is C ∞ . Since v ′ = Dp(v) ∈ ∆sǫ ,
there exists a sequence kn → ∞ such that ϕkn (v ′ ) ∈ ∆ǫ . At the same time,
Proposition 6.5 implies that
lim dSas (ϕt (v), ϕt (w)) = 0,
t→∞

and so for n sufficiently large, dSas (ϕkn (v), ϕkn (w)) < rǫ /2, where rǫ > 0
is the lower bound on the restriction of r to ∆ǫ . But this implies that
Dp(ϕkn (w)) ∈ Wloc s (ϕ (v ′ )). Since ϕ
kn kn is a diffeomorphism, we conclude
that there is a neighborhood of w in W s (v) that is diffeomorphic to the C ∞
submanifold Wloc s (ϕ (v ′ )). Since w was arbitrary, this implies that W s (v)
kn
THE WEIL-PETERSSON GEODESIC FLOW 65
T
is a C ∞ submanifold of T 1 M . The intersection Λs2 := ǫ>0 ∆sǫ ∩ Dp(Ω1 ) is
a full volume subset of T 1 N , and we have shown that for every v ∈ Ωs2 :=
Dp−1 (Λs2 ), the submanifold W s (v) is C ∞ .
For each v ∈ Ωs2 , consider the map ψ from Hvs × R to M that sends
(y, t) to π(ϕt (wvs (y))), where wvs (y) = − grad bsv (y). Since W s (v) is C ∞ , the
function wvs (y) is C ∞ along Hvs ; it follows that ψ is a diffeomorphism. In the
coordinates on M given by ψ, the Busemann function bsv assigns the value
−t to the point (x, t). It follows that bsv is C ∞ , for every v ∈ Ωs2 . Similarly,
there is a set Ωu2 of full measure such that buv is C ∞ for every v ∈ Ωu2 . Setting
Ω2 = Ωu2 ∩ Ωs2 , we obtain the full measure set where the conclusions of the
proposition will hold.
We establish the absolute continuity properties of W s ; analogous argu-
ments show the properties for W u . The preceding arguments show that for
every v ∈ Λ2 there exists an integer k ≥ 0 such that
(29) ϕk (W s (v, δ)) ⊂ Wloc
s
(ϕk (v)), for every δ < inj(π(v))
For a Sfixed k ≥ 0, denote by Xk the set of v ∈ Λ2 for which (29) holds. Then
Λ2 = k≥0 Xk .
S
Suppose that m(Z) = 0, for some Z ⊂ T 1 N . Then the set Ẑ = k≥0 ϕk (Z)
also has measure 0. It follows from Proposition 6.9 that for almost every
w ∈ Λ1 , the induced Riemannian measure of Ẑ in Wloc s (w) is zero. But this

implies in particular that for every k ≥ 0 and for almost every v ∈ Xk , the
induced Riemannian measure of ϕk (Z) ⊂ Ẑ in ϕk (W s (v, δ)) ⊂ Wloc s (ϕ (v))
k
s
is zero; hence the induced volume of Z in W (v, δ) is 0, for all δ < inj(π(v)).
This establishes (1).
Suppose that D is a C 1 -embedded, n-dimensional disk in T 1 N . Fix δ <
1
2 inf v∈D inj(π(v)). Suppose that B ⊂ D has induced Riemannian volume 0.
Let [
Bk = B ∩ W s (w, δ)
w∈Xk
and note that [
Sats (B, δ) = Sats (Bk , δ);
k≥0
s
hence it suffices to show that m(Sat (Bk , δ)) = 0, for all k ≥ 0.
Fix k ≥ 0. For each w ∈ Bk , there an n-dimensional open ball Dw ⊂ D
S
centered at w in the induced Riemannian metric in D, such that kj=0 ϕj (Dw ) ⊂
T 1 N . Since ϕk is a diffeomorphism, the set ϕk (Bk ∩ Dw ) has induced Rie-
mannian volume zero in the n-dimensional disk ϕk (Dw ). It follows from
Proposition 6.9 that m(Satsloc (ϕk (Bk ∩ Dw ))) = 0, and so
m (ϕ−k (Satsloc (ϕk (Bk ∩ Dw )))) = 0.
But (29) implies that
Sats (Bk ∩ Dw , δ) ⊂ ϕ−k (Satsloc (ϕk (Bk ∩ Dw ))) ,
66 K. BURNS, H. MASUR AND A. WILKINSON

and so m(Sats (Bk ∩Dw , δ)) = 0. Now fix a countable cover {Dwi : wi ∈ Bk }
of Bk in D by such balls (this is possible by the Besicovitch covering theorem,
since D is an embedded C 1 submanifold). Then
[
Sats (Bk , δ) ⊂ Sats (Bk ∩ Dwi , δ),
i

and so m(Sats (Bk , δ)) = 0. Conclusion (2) follows. ⋄

Proof of ergodicity. Assume I.-VI. The proof that ϕt is ergodic is an adapta-


tion of the standard “Hopf Argument,” along the lines of the proof of local
ergodicity in [15]. To prove ergodicity, it suffices to show that for every
continuous function f : T 1 N → R with compact support:
Z Z
1 T
(30) lim f (ϕt (v)) dt = f dm, for m − a.e. v ∈ T 1 N
T →∞ T 0 1
T N

Indeed, if (30) holds for a dense set of functions f in L2 , then by continuity


RT
of the projection f 7→ B(f ) = limT →∞ T1 0 f ◦ϕt dt, (30) will hold for every
f in L2 .
Fix then a continuous function f with compact support and define mea-
surable functions f s and f u by:
Z Z
s 1 T u 1 0
f (v) = lim sup f (ϕt (v)) dt, and f (v) = lim sup f (ϕt (v)) dt.
T →∞ T 0 T →∞ T −T

The Birkhoff Ergodic Theorem implies that there is a set G ⊂ T 1 N of full


measure such that for every v ∈ G, we have f s (v) = f u (v) = B(f )(v). Since
f is continuous with compact support, and the leaves of W s are contracted
by ϕt , it follows that f s is constant along leaves of W s . Similarly, f u is
constant along leaves of W u . Finally, all three functions f s , f u , B(f ) are
invariant under the flow ϕt .
Now fix a arbitrary element v ∈ T 1 N . We will show that there is a
neighborhood Uv of v on which B(f ) is almost everywhere constant. Since
1 1
RT N is connected, R this will imply that B(f ) is constant on T N . Since
T 1 N B(f ) dm = T 1 N f dm, it will then follow that (30) holds, and so ϕ is
ergodic.
Let δ = δ(v) = 41 min{inj(π(v)), d(v, ∂N )}, and let V be the δ-neighborhood
of v in T 1 N . For w ∈ Λ2 ∩ V , consider the set

Nδ (w) = Satu ϕ(−δ,δ) (W s (w, δ)) , δ ;
We claim:
(a) for almost every w ∈ Λ2 ∩ V , B(f ) is almost everywhere constant on
Nδ (w);
(b) there is a neighborhood Uv ⊂ V of v such that for almost every
w ∈ Uv , the set Nδ (w) ∩ Uv has full measure in Uv .
THE WEIL-PETERSSON GEODESIC FLOW 67

Together, these statements imply that there is a neighborhood Uv of v on


which B(f ) is a.e. constant, completing the proof of ergodicity.
We first establish part (a) of this claim. Let G be the full measure subset
of v ∈ Λ2 where the limit (30) exists and f u (v) = f s (v) = B(f )(v). Absolute
continuity property (1) of W s in Proposition 6.10 implies that for almost
every w ∈ V ∩ Λ2 , the intersection G ∩ W s (w, δ) has full volume in W s (w, δ)
(that is, its complement has induced volume 0). Fix such a w. On W s (w, δ),
f s takes a constant value f s ≡ a. On the full volume subset G ∩ W s (w, δ),
f u coincides with f s and therefore also takes the constant value a. Since f u
is ϕt -invariant, and ϕt is a C ∞ flow, f u takes the constant value a on the
full measure subset G′ := ϕ(−δ,δ) (G ∩ W s (w, δ)) of the n-dimensional C ∞
submanifold D = ϕ(−δ,δ) (W s (w, δ)).
But f u is constant along W u manifolds and so takes the constant value
a on Satu (G′ , δ). Since W u satisfies the absolute continuity property (2) in
Proposition 6.10, and G′ has full measure in D, it follows that Satu (G′ , δ)
has full measure in Satu (D, δ) = Nδ (w). Hence f u is constant on a full
measure subset of Nδ (w). Since f u = B(f ), a.e., it follows that B(f ) is
almost everywhere constant on Nδ (w), proving part (a).
We next establish part (b) of the claim. Absolute continuity property
(1) of W ∗ implies that for almost every w′ ∈ Λ2 ∩ V , the disk W ∗ (w′ , δ)
is almost everywhere tangent to E ∗ , which is contained in the continuous
conefield C ∗ . Hence for almost every w′ , the tangent bundle T (W ∗ (w′ , δ)) is
everywhere contained in C ∗ . Invariance of W s under ϕt implies that for any
w ∈ Λ2 ∩ V , the tangent bundle to the disk D(w) = ϕ(−δ,δ) (W s (w, δ)) is
everywhere contained in C s ⊕ E 0 . The conefields C u and C s and the line field
E 0 = Rϕ̇ are continuous (indeed, they are smooth), and so locally the angle
between any two of them is bounded below. It follows that there exists a
neighborhood Uv ⊂ V of v such that for any w, w′ ∈ Λ2 ∩ Uv :
W u (w′ , δ) ∩ D(w) 6= ∅;
in other words, for every w ∈ Λ2 ∩ Uv :
Nδ (w) = Satu (D(w), δ) ⊃ Λ2 ∩ Uv .
This completes the proof of part (b) of the claim, and the proof ergodicity.

6.1. Proof of Proposition 6.9: verifying the conditions of Katok-


Strelcyn-Ledrappier. Let V ⊂ T 1 N be the set of v ∈ T 1 N such that
ϕt (v) ∈ T 1 N , for all t ∈ (−1, 1). Fix t0 ∈ (0, 1) and consider the restriction
of the time-t0 map ϕt0 to V . In this section we verify that the main hypothe-
ses in [17] hold for the map ϕt0 : V → T 1 N . The main results in [17] then
imply the conclusions of Proposition 6.9. To paraphrase [17], the conditions
we will verify ensure that the set of singularities of the map ϕt0 is “thin”
and that the first and second derivatives of ϕt0 grow moderately near this
set. Throughout this subsection, we assume that assumptions I.-VI. hold.
68 K. BURNS, H. MASUR AND A. WILKINSON

In the setup of [17], the background hypotheses are: X is a compact


metric space, and V is an open and dense subset of X carrying a Riemann
structure with controlled singularities near X \ V . In our application, V is
the set described above, endowed with the Sasaki Riemann structure, and
X = T 1 N is the completion of T 1 N in the Sasaki distance metric dSas . We
first verify that X is compact, which establishes condition (A) of [17].
Lemma 6.11. (T 1 N , dSas ) is compact.
Proof. Let hvn,m im be a sequence of elements of T 1 N , where for each m ≥ 1,
hvn,m i is a Sasaki Cauchy sequence in T 1 N . Since dSas (v, w) ≥ d(π(v), π(w)),
it follows that for each m, the sequence hπ(vn,m )i is Cauchy in N ; since N
is compact, by passing to a subsequence in the m’s, we may assume that
hπ(vn,m )im converges to a Cauchy sequence hxn i in N . What this means is
that for every ǫ > 0 there exists an m0 > 0 such that for m ≥ m0 , we have
lim d(π(vn,m ), xn ) < ǫ.
n→∞

Now for each n, consider the collection {v̂n,m | m ≥ 1} ⊂ Tx1n N obtained


by parallel translating each vn,m along a geodesic from Tπ(vn,m ) N to Txn N .
Using compactness of Tx1n N and a diagonal argument, we obtain a subse-
quence mk such that for each n, v̂n,mk converges as k → ∞ to an element
v̂n ∈ Tx1n N , uniformly in n; that is, for every ǫ > 0, there exists k0 > 0 such
that for all k > k0 we have
lim kv̂n,mk − v̂n k < ǫ
n→∞

Since the Sasaki distance dSas (vn,mk , v̂n ) is bounded by d(π(vn,mk ), xn ) +


kv̂n,mk − v̂n k, it follows that for every ǫ > 0 there exists a k1 > 0 such that
for all k ≥ k1 ,
lim dSas (vn,mk , v̂n ) ≤ lim d(π(vn,m ), xn ) + kv̂n,mk − v̂n k < ǫ.
n→∞ n→∞

Hence hvn,mk imk converges as k → ∞ to the Sasaki Cauchy sequence hv̂n i ∈


T 1N . ⋄

Clearly V is an open and dense subset of T 1 N . Let S = T 1 N \ V . The


Sasaki distance from v to the singular set S is bounded above by the distance
from π(v) to ∂N .

6.1.1. More (yet) on the Sasaki metric. Condition (B) in [17], which con-
cerns the Riemann structure on V , has three parts that require verification.
In this subsection, we establish bounds on the derivatives of the Sasaki expo-
nential map exp : T V → V , which we will then use to verify these conditions
as well as later conditions on ϕt0 . To control the Sasaki exponential map,
we will need to control the first three derivatives of the Sasaki metric; these
can be related to the higher order derivatives of the metric on N via the
following lemma.
THE WEIL-PETERSSON GEODESIC FLOW 69

Lemma 6.12. There exists a cubic polynomial C : R3 → R such that for


any Riemannian manifold N and any v ∈ Tx1 N , the Sasaki curvature tensor
RSas satisfies
k(RSas )v k + k∇(RSas )v k ≤ C(kRx k, k∇Rx k, k∇2 Rx k),
where R is the Riemannian curvature tensor on N .
Proof. The sectional curvatures of the Sasaki metric on the unit tangent
bundle can be computed as follows [18]. We use the usual identification
T(x,u) T N ∼
= (Tx N )2 . Let Π be a plane in T(x,u) T 1 N , and choose an or-
thonormal basis {(v1 , w1 ), (v2 , w2 )} for Π satisfying kvi k2 + kwi k2 = 1 for
i = 1, 2 and hv1 , v2 i = hw1 , w2 i = 0. Then the Sasaki sectional curvature of
Π is given by
KSas (Π) = hRx (v1 , v2 )v2 , v1 i + 3hRx (v1 , v2 )w2 , w1 i + kw1 k2 kw2 k2
3 1 1
− kRx (v1 , v2 )uk2 + kRx (u, w2 )v1 k2 + kRx (u, w1 )v2 k2
4 4 4
1
+ hRx (u, w1 )w2 , Rx (u, w2 )v1 i − hRx (u, w1 )v1 , Rx (u, w2 )v2 i
2
+h(∇v1 R)x (u, w2 )v2 , v1 i + h(∇v2 R)x (u, w1 )v1 , v2 i.
The conclusion follows from the Chain Rule and the well-known identities
relating the sectional curvatures with the norm of the curvature tensor. ⋄

The next lemma will be used to bound the derivative of the Sasaki expo-
nential map.
Lemma 6.13. Let Y be a Riemannian manifold, and let J be a Jacobi field
along a geodesic γ : [−δ0 , δ0 ] → Y satisfying J(0) = 0 and kJ ′ (0)k = 1.
Suppose that
sup kRγ(t) k ≤ R0
|t|<δ0
for some R0 > 1. Let ǫ ∈ (0, 1) be given, and let t0 = min{δ0 , ǫ/(3R0 )}.
Then for all |t| ≤ t0 we have
(1 − ǫ)|t| ≤ kJ(t)k ≤ (1 + ǫ)|t|
and
|J ′ (t)| ≤ 1 + ǫ
Proof. Let a(t) = kJ(t)k, and let b(t) = kJ ′ (t)k. Then the Cauchy-Schwarz
inequality implies
|(a2 )′ | = |2aa′ | = |2hJ, J ′ i| ≤ 2ab,
and since |t| < δ0 :
|(b2 )′ | = |2bb′ | = |2hJ ′ , J ′′ i| = |2hJ ′ , R(γ̇, J)γ̇i| ≤ 2R0 ab.
We conclude that wherever |a| and |b| are not zero, we have
|a′ | ≤ b and |b′ | ≤ R0 a.
70 K. BURNS, H. MASUR AND A. WILKINSON

We are assuming that a(0) = 0 and b(0) = 1. Without loss of generality,


assume that |a(t)| > 0 for t > 0 (otherwise, we may replace t = 0 with
a positive value of t in the following argument). From this we obtain the
integral inequality, for t ≥ 0:
Z t Z t
′ ′
(31) |a (t)| ≤ 1 + |b (s)| ds ≤ 1 + R0 a(s) ds.
0 0
Suppose that, for some t1 ∈ (0, t0 ) we have ′
|a (t)|
< 1 + ǫ for all t ∈ [0, t1 )
and |a′ (t1 )| = 1 + ǫ. Then a(t) < (1 + ǫ)t, for all t ∈ [0, t1 ); combined with
(31), this gives that
Z t1
′ R0 (1 + ǫ) 2
|a (t1 )| ≤ 1 + R0 (1 + ǫ)s ds < 1 + t1 < 1 + ǫ,
0 2
since ǫ ∈ (0, 1) implies that
ǫ2 2ǫ
t21 < t20 ≤ < .
9R02 R0 (1 + ǫ)
This contradicts our assumption that |a′ (t1 )| = 1 + ǫ. We conclude that
|a′ (t)| < 1 + ǫ for all t ∈ (0, t0 ); similarly, |a′ (t)| < 1 + ǫ, for all t ∈ (−t0 , 0).
From this we conclude that a(t) ≤ (1 + ǫ)|t| for all |t| ≤ t0 .
We now prove the lower bound. Since b(0) = 1 and |b′ (t)| ≤ R0 a(t), for
|t| ≤ t0 we have
(1 + ǫ)R0 t2
b(t) ≥ 1 − .
2
On the other hand, we know that
(a2 )′′ = 2b2 − 2hR(J, γ̇)γ̇, Ji ≥ 2b2 − 2R0 a2
" 2 #
(1 + ǫ)R0 t2
>2 1− − (1 + ǫ) R0 t > 2[1 − 2(1 + ǫ)2 R0 t2 ],
2 2
2
(using the lower bound for b(t) and upper bound of (1 + ǫ)|t| for a(t)). Now,
since
ǫ2 ǫ2
t2 ≤ t20 ≤ < ,
9R02 2(1 + ǫ)2 R0
we find that
(a2 )′′ (t) > 2[1 − 2(1 + ǫ)2 R0 t2 ] > 2(1 − ǫ2 ).
But then 2a(t)a′ (t) = (a2 )′ (t) > 2(1 − ǫ2 )|t|, and again using the upper
bound on a, we get
(1 − ǫ2 )|t| (1 − ǫ2 )|t|
a′ (t) > > = 1 − ǫ;
a(t) (1 + ǫ)|t|
hence a(t) > (1 − ǫ)|t|.
Finally, since b(0) = 1 and |b′ | ≤ R0 a ≤ R0 (1 + ǫ), it follows that |b(t)| ≤
1 + |t|R0 (1 + ǫ), and so for |t| < |t0 |, we have |b(t)| ≤ 1 + ǫ(1 + ǫ)/3 < 1 + ǫ.
The final conclusion follows. ⋄
THE WEIL-PETERSSON GEODESIC FLOW 71

We apply this lemma to the Sasaki exponential map exp : T V → V to


obtain:
Proposition 6.14. There exist constants δ1 > 0 and k1 > 1 such that for
every v0 ∈ V , if dSas (v0 , S) < δ1 , then for all v ∈ V with dSas (v, v0 ) <
dSas (v0 , S)k1 :
1 − dSas (v0 , S) ≤ kDv exp−1
v0 k
−1
≤ kDξ expv0 k ≤ 1 + dSas (v0 , S),
where ξ = exp−1
v0 (v).

Proof. Let v0 ∈ V and ξ = exp−1 ˆ ξ


v0 (v). Let ξ = kξk be the unit vector in
the direction of ξ. Suppose ξ ′ ∈ Tv10 V is an orthogonal unit vector. Let
a(s, t) = (ξ̂ + sξ ′ )t be the 1-parameter family of rays through the origin in
Tv0 V . Let
α(s, t) = expv0 ◦ a(s, t)
be the 1-parameter family of image geodesics in V . We consider the cor-
responding Jacobi field J(t) along α(0, t) defined by J(t) = ∂α(s, t)/∂s at
s = 0. Clearly J(0) = 0 and J ′ (0) = ξ ′ . Setting t1 = kξk, by the chain rule
we have
kJ(t1 )k = kt1 Dξ expv0 (ξ ′ )k.
Thus we have to bound k J(t 1)
t1 k above and below.
By Lemma 6.12, the sectional curvatures of the Sasaki metric on the unit
tangent bundle are bounded polynomially in terms of the absolute value
of the curvature and the derivative of the curvature of the original metric.
Assumption IV. gives a bound for these latter quantities, and therefore a
polynomial bound on the curvatures in the Sasaki metric, in the reciprocal
of the distance to the singular set S. It follows that there exist k0 > 1 and
δ0 > 0 such that for all v0 ∈ V with dSas (v, S) < δ0 , the Sasaki curvature
tensor RSas satisfies
k(RSas )v k < dSas (v, S)−k0 .
Let k1 = k0 + 2. Then there exists δ1 ∈ (0, 1/3) such that if dSas (v0 , S) <
δ1 and
dSas (v0 , v) ≤ dSas (v0 , S)k1 ,
then the maximum norm R0 of the Sasaki curvature tensor along the geo-
desic joining v0 to v also satisfies R0 < dSas (v0 , S)−k0 . Lemma 6.13 implies
that

J(t1 )
(32) 1−ǫ≤ < 1 + ǫ,
t1
provided ǫ > 3R0 |t1 | = 3R0 dSas (v0 , v). Hence if dSas (v0 , S) < δ1 and
dSas (v, v0 ) ≤ dSas (v0 , S)k1 , then (32) holds for ǫ = dSas (v0 , S), since
3R0 dSas (v0 , v) < 3dSas (v0 , S)k0 · dSas (v0 , S)k1
= 3dSas (v0 , S)2 < dSas (v0 , S) = ǫ.
72 K. BURNS, H. MASUR AND A. WILKINSON

The next proposition gives bounds on the second derivative of exp, which
we will later use to verify condition (1.3) of [17].
Proposition 6.15. There exist constants δ2 > 0 and k2 > 1 such that for
every v0 ∈ V , if dSas (v0 , S) < δ2 , then for all v ∈ V with dSas (v, v0 ) <
dSas (v0 , S)k2 :
max{kDξ2 expv0 k, kDv2 exp−1
v0 k} < dSas (v0 , S)
−k2

where ξ = exp−1
v0 (v).

Proof. Fix v0 ∈ V and v ∈ V in a neighborhood of v0 . Let ξ = exp−1 v0 (v).


ˆ ξ ′
Let ξ = kξk be the unit vector in the direction of ξ, and suppose ξ ∈ Tv10 V is
an orthogonal unit vector. As in the proof of Proposition 6.14, we consider
the variation of geodesics
α(s, t) = expv0 ◦ a(s, t),
where a(s, t) = (ξ̂ + sξ ′ )t. Define Z, J and Q by
d d d2 d
Z(s, t) =α(s, t); J(s, t) = α(s, t); Q(s, t) = 2 α(s, t) = J(s, t).
dt ds ds ds
The chain rule implies that
2
Q(0, t) = Da(0,t) expv0 (tξ ′ , tξ ′ ),
and so
1
kDξ2 expv0 (ξ ′ , ξ ′ )k = kQ(0, kξk)k,
kξk2
since ξ = a(0, kξk).
Observe that for a fixed s, J(s, ·) is a Jacobi field down the geodesic α(s, ·)
and so satisfies the Jacobi equation
d
J = −RSas (Z, J)Z.
dt2
From this, the symmetry of the Levi-Civita connection, and the definition
of Q it follows that
d d d d
Q= J = − RSas (Z, J)Z
dt2 ds dt2 ds
  
d ′ ′
=− RSas (Z, J)Z + RSas (J , J)Z + RSas (Z, Q)Z + RSas (Z, J)J ,
ds
where ′ denotes the derivative with respect to t. Then kQ′′ (0, t)k ≤ C1 (t) +
kQ(0, t)kC2 (t), where
C1 (t) = k (∇RSas )exp k kJ(0, t)k+2k (RSas )exp kkJ(0, t)k kJ ′ (0, t)k,
v0 (tξ̂) v0 (tξ̂)

and
C2 (t) = k (RSas )exp k.
v0 (tξ̂)
THE WEIL-PETERSSON GEODESIC FLOW 73

Assumption IV. and Lemma 6.12 imply that there exists k0 > 1 such that
max{k (RSas )v0 k, k (∇RSas )v0 k} < dSas (v0 , S)−k0 .
Fix δ2 ∈ (0, 1/10) such that if dSas (v0 , S) < δ2 , then
sup max{k (RSas )expv (tξ) k, k (∇RSas )expv0 (tξ) k} < dSas (v0 , S)−k0 −1 .
0
|t|≤dSas (v0 ,S)k0 +1

Assume that dSas (v0 , S) < δ2 . Lemma 6.13 implies that for |t| < dSas (v0 , S)k0 +2 ,
both kJ(0, t)k and kJ ′ (0, t)k are bounded by 2, and so
C1 (t) ≤ 10dSas (v0 , S)−k0 −1 < dSas (v0 , S)−k0 −2 ,
and
C2 (t) ≤ dSas (v0 , S)−k0 −1 < dSas (v0 , S)−k0 −2 .
Let q(t) = kQ(0, t)k and let r(t) = kQ′ (0, t)k. As in the proof of Lemma 6.13,
we have that
|qq ′ | = |hQ, Q′ i| ≤ qr, and |rr ′ | ≤ |hQ′ , Q′′ i| ≤ r(C1 + qC2 ).
Note that q(0) = r(0) = 0. An analysis similar to that in the proof of
Lemma 6.13 (whose details we omit) shows that for |t| < dSas (v0 , S)k0 +2 ,
we have
q(t) ≤ t2 dSas (v0 , S)−k0 −2 .
Hence, if kξk ≤ dSas (v0 , S)k0 +2 , then
q(kξk)
kDξ2 expv0 (ξ ′ , ξ ′ )k = ≤ dSas (v0 , S)−k0 −2 .
kξk2
Hence the upper bound for kDξ2 expv0 k in the conclusion of the proposition
holds, with the exponent k2 = k0 + 2. An upper bound for kDv2 exp−1 v0 k on
−k follows from this upper bound on kDv exp−1
2
the order of dSas (v0 , S) v0 k,
2

the upper bound on kDv exp−1 v0 k given by Proposition 6.14, and the fact
that
  
D 2 exp−1 = − D exp−1 D 2 exp D exp−1 .

6.1.2. Verifying condition (B) in [17]. For v ∈ V , let inj(v) denote the
radius of injectivity of the Sasaki exponential map expv : Tv V → V . Since
dSas (v, w) ≥ d(π(v), π(w)), the controlled injectivity assumption V. implies
that
inj(v) ≥ inj(π(v)) ≥ Cd(π(v), ∂N )β ≥ CdSas (v, S)β .
This implies condition (Ba) of [17]. Conditions (Bb) and (Bc) in [17] follow
in a straightforward way from Proposition 6.14.
74 K. BURNS, H. MASUR AND A. WILKINSON

6.1.3. Verifying conditions (1.1) – (1.4) of [17]. Conditions (1.1) – (1.4) of


[17] concern the volume of the singular set S and the behavior of ϕt0 near
S. Condition (1.1) of [17], which concerns the volume of a neighborhood of
S, follows directly from Lemma 6.3. Condition (1.2) of [17], concerning the
integrability of log+ kDϕt0 k, follows immediately from Lemma 6.7. Condi-
tion (1.4) of [17] requires a bound on kDv0 ϕt0 k on the order of dSas (v0 , S)−β ,
for some β > 0. This follows in a straightforward way from assumption VI.
This leaves condition (1.3).
Fix v0 ∈ V , and let Φ = Φv0 : Tv0 V → Tϕt0 (v0 ) V be defined by
Φ = exp−1
ϕt (v0 ) ◦ ϕt0 ◦ expv0 .
0

Condition (1.3) of [17] requires a bound on the second derivative of Φ as


an inverse power of the distance to the singular set, which follows from the
next proposition.
Proposition 6.16. There exist constants δ3 > 0 and k3 > 1 such that for
every v0 ∈ V , if dSas (v0 , S) < δ3 , then for all v ∈ V with dSas (v, v0 ) <
dSas (v0 , S)k3 :
kDξ2 Φv0 k < dSas (v0 , S)−k3 ,
where ξ = exp−1
v0 (v).

Proof. Choose constants k2 > 1 and δ2 > 0 satisfying the conclusions of


Proposition 6.15 and such that if dSas (v0 , S) < δ2 , then for every |t| ≤ t0
sup max{k (RSas )v k, k (∇RSas )v k} < dSas (v0 , S)−k2 .
dSas (v,ϕt (v))≤dSas (ϕt (v),S)k2

By assumption VI., there exist δ3 < min{δ2 , 1/(8n)} and k2′ > k2 such that
for dSas (v0 , S) < δ3 , and all |t| ≤ t0 :

max{kDv0 ϕt k, kDv0 ϕ−t k} ≤ dSas (v0 , S)−k2 .
Proposition 4.2 implies that if ξ, ξ ′ ∈ T V satisfy dSas (πV (ξ), πV (ξ ′ )) <
dSas (vt , S)k2 then

dSas (Dϕt (ξ), Dϕt (ξ ′ )) ≤ (8n) dSas (v0 , S)−4k2 −3k2 dSas (ξ0 , ξ1 )

≤ dSas (v0 , S)−7k2 . dSas (ξ0 , ξ1 ).
To bound the norm kDξ2 Φv0 k it suffices to bound the Lipschitz constant of
the map Dξ Φv0 in a small neighborhood of ξ. This in turn is bounded by the
product of the Lipschitz constants of the three factors D exp−1 ϕt0 (v0 ) , Dϕt0
and D expv0 in the composition defining Dξ Φv0 .
The Lipschitz constants for D expv0 and D exp−1 ϕt (v0 ) are bounded for
0
nearby points by the norms kD 2 expv0 k and kD 2 exp−1
ϕt (v0 ) k, which by 0
Proposition 6.15 are both bounded on the order of dSas (v0 , S)−k2 . We have
just shown that the Lipschitz constant for Dϕt0 is bounded on the order of

dSas (v0 , S)−7k2 . Hence the Lipschitz constant of Dξ Φv0 is bounded on the
THE WEIL-PETERSSON GEODESIC FLOW 75

order of dSas (v0 , S)−k4 , for k4 = 2k2 + 7k2′ . The details of this argument are
left to the reader. ⋄

This completes the verification of the hypotheses in [17] implying the


conclusions of Proposition 6.9.
6.2. Additional conditions in [17] implying finite, positive entropy.
The final conclusion of Theorem 6.6 that remains to be proved concerns
the entropy of ϕ. The positivity of the entropy follows from [17] and the
hypotheses we have just verified. Finitude of the entropy requires that an
additional hypothesis – Condition (C) – hold. As stated in [17], condition
(C) is the requirement that the capacity of the space X = T 1 N be finite. In
fact, a slightly weaker condition is required, which is given by the following
proposition. Recall that Uρ , for ρ > 0, denotes the set of v ∈ T 1 N such that
d(π(v), ∂N ) < ρ.
Proposition 6.17. There exists q > 1 such that if ρ0 > 0 is sufficiently
small, then for any ρ < ρ0 there is a cover of T 1 N \ Uρ0 by open balls of
radius ρ, whose cardinality does not exceed ρ−q .
Proof. Proposition 6.14 implies that there exist δ > 0 and k > 1 such that
for ρ0 < δ and all v ∈ T 1 N \ Uρ0 , the derivative of the Sasaki exponential
map Dξ expv and its inverse have norm bounded by 2, for all kξk < ρk0 .
Hence on a ball of radius ρk0 in T 1 N \ Uρ0 , the Sasaki metric is uniformly
comparable to Euclidean; in particular, the volume of a ball of radius ρ ≤ ρk0
is bounded below by c−1 ρn , where c > 1 is a universal constant, and the
local geometry of covers by balls is the same as Euclidean.
For ρ < ρ0 , the Besicovitch covering lemma then implies that T 1 N \ Uρ0
p(ρ)
can be covered by balls {Bℓ (ρ)}ℓ=1 of radius ρk+1 with universally bounded
overlap C. Hence
 
  X p(ρ) p(ρ)
[
p(ρ) c−1 ρn(k+1) ≤ m(Bℓ (ρ)) ≤ Cm  Bℓ  ≤ Cm(T 1 N ) = C
ℓ=1 ℓ=1

and so
p(ρ) ≤ Ccρ−n(k+1) ,
for all ρ < ρ0 . This implies the conclusion of the proposition, with q =
n(k + 1) + 1.

7. Ergodicity and finite entropy of the WP geodesic flow


Fix a Riemann surface S, and let T = Teich(S), MCG = MCG(S) and
M = M(S). We describe here first how the results of Section 6 can be
applied to obtain ergodicity and finite entropy of the geodesic flow on the
quotient M1 = T 1 T /MCG. Note that the results in Section 6 cannot be
76 K. BURNS, H. MASUR AND A. WILKINSON

applied directly with M = T and Γ = MCG, since MCG does not act freely
on T . Our strategy is to prove ergodicity first for a finite branched cover
T 1 T /MCG[3]. Here MCG[k] is the level k congruence subgroup:
MCG[k] = {φ ∈ MCG : φ∗ = 0 acting on H1 (S, Z/kZ)},
which is clearly a finite index subgroup of MCG. If is a well-known fact that
for k ≥ 3, MCG[k] is torsion-free and so acts freely and properly discontin-
uously by isometries on T . The quotient T 1 T /MCG[k] has finite volume.
We obtain ergodicity for the flow on T 1 T /MCG[k] for any k ≥ 3 using the
setup of the previous section. This result has independent interest.

7.1. Ergodicity of the flow on T 1 (T /MCG[k]). Fix k ≥ 3. To es-


tablish ergodicity and finite metric entropy of the W P geodesic flow on
T 1 (T /MCG[k]), we show that the assumptions I.-VI. of the previous sec-
tion are satisfied for M = T , Γ = MCG[k] and the WP metric.
The fact that in the Weil-Petersson metric T is geodesically convex was
proved by Wolpert [38]. Since the completion M of M is compact, and
T /MCG[k]) is a finite branched cover of M, it follows that the completion
of T /MCG[k] is compact as well. Hence assumptions I. and II. hold true.
The curvature bound in assumption IV. is due to Wolpert and was stated
as Proposition 3.8. The bounds on k∇RW P k and k∇2 RW P k in assumption
IV. is the content of Theorem 5.1. Assumption VI. was proved in The-
orem 3.1. It remains to prove Assumptions III. and V. For X ∈ T , we
continue to denote by ρ0 (X) the WP distance from X to ∂T .
Verifying assumption III.: ∂ (T /MCG[k]) is volumetrically cusplike.
Given ρ, let
Eρ = {X : ρ0 (X) ≤ ρ}.
Lemma 7.1. We have Vol(Eρ ) = O(ρ4 )
Proof. Cover Eρ with a finite number of open sets in which we can use
the plumbing coordinates (s, t) introduced in Section 5.1.1. Fix one such
neighborhood U with p ≥ 1 short curves. The matrix (gi,j ) of the metric is
the inverse of the matrix (hφi , φj i).
By the estimates in Section 5.1 the determinant of the matrix (hφi , φj i)
is
Y p
≍ −|ti |2 (log |ti |)3 .
i=1
The volume element is then the determinant of the inverse and the volume
is then found by integration;
Yp Z |ti | p
Y
|dti | 1
Vol(U ) = 2 (log |t |)3
= .
0 −|t i | i (− log |ti |)2
i=1 i=1

Thus Vol(Eρ ) = O(ρ4p ). ⋄


THE WEIL-PETERSSON GEODESIC FLOW 77

Verifying assumption IV.: T /MCG[k] has controlled injectivity ra-


dius.
For α ∈ C, denote by τα ∈ MCG the Dehn twist about the curve α. Given
a simplex σ = {α1 , . . . , αp } ∈ C(S), let Γ(σ) =< τ1 , . . . , τp > be the abelian
group generated by the Dehn twists about the αi .
Lemma 7.2. For each ǫ > 0 there exists c > 0 and j ≤ (3g − 3 + n)! such
that if φ ∈ MCG[k] and d(X, φ(X)) < c, then there exists σ ∈ C(S) such
that X ∈ Ω(σ, ǫ) and φj ∈ Γ(σ).
Proof. Let ǫ > 0 be given. Since MCG[k] acts properly discontinuously
without fixed points, the first conclusion must be true: there exists c0 > 0
such that if φ ∈ MCG[k] and d(X, φ(X)) < c0 , then X ∈ Ω(σ, ǫ), for some
σ ∈ C(S). Now suppose the second statement is not true; i.e., suppose that
φj ∈/ Γ(σ) for any j smaller than the order of the permutation group of σ.
Then there is a sequence Xm ∈ Ω(σ, ǫ) and φm ∈ MCG such that for every
j, we have: φjm ∈ MCG[k] \ Γ(σ) and
dW P (Xm , φm (Xm )) → 0.
Passing to a subsequence and applying an element ψm ∈ Γ(σ) if necessary,
we can assume Xm converges to a noded surface Xσ . There is a constant
c1 = c1 (δ) such that if there is a curve β and a pair of surfaces Z1 and Z2
such that ℓβ (Z1 ) ≤ δ/2 and ℓβ (Z2 ) ≥ δ, then
dW P (Z1 , Z2 ) ≥ c1 .
This implies that φm (σ) = σ for m sufficiently large, and so for some j, the
mapping class φjm preserves the individual curves of σ.
We are assuming that φjm ∈ / Γ(σ). The classification of elements of MCG
implies that the restriction of φjm to each piece of Xσ is either pseudo-Anosov
or finite order. Since φjm ∈ MCG[k], it cannot have finite order on every
piece of Xσ and so must be pseudo-Anosov on some piece. But then there
is a uniform lower bound [9, Theorem 7.6] for dW P (Xσ , φjm (Xσ )) and thus a
lower bound for dW P (Xm , φ(Xm )) for m sufficiently large, a contradiction.

Lemma 7.3. There is a constant c > 0 such that for any X ∈ T /MCG[k]:
inj(X) ≥ cρ0 (X)3 .
Proof. By Proposition 15 of [36] there is a positive constant c > 0 such that
for X ∈ Ω(σ, ǫ), dW P (X, Γ(σ)(X)) ≥ cρ0 (X)3 . This bounds the injectivity
radius from below. ⋄

Applying Theorem 6.6, we have now proved


Theorem 7.4. The Weil-Petersson flow is ergodic and has finite entropy
on T 1 T /MCG[k].
78 K. BURNS, H. MASUR AND A. WILKINSON

7.2. Ergodicity of the flow on M1 (S): Proof of Theorem 1. The


manifold T /MCG[k] is a finite branched cover over M. Let h : X → X a
conformal automorphism of finite order. Then the action h : T → T has a
fixed point set F (h). It is known [30] that if S is compact, unless h is the
hyperelliptic involution in genus 2, then F (h) has complex dimension at most
3g − 5. In fact F (h) is the Teichmüller space of the quotient orbifold X/h.
In genus 2 the hyperelliptic involution fixes every point. In the noncompact
case where S has punctures, the complex dimension of F (h) is at most 3g−4.
Let F denote the union of the fixed point sets of the action of all finite order
elements of MCG(S), excluding the genus 2 hyperellipic case. This is a
countable union of lower dimensional Teichmüller spaces.
Lemma 7.5. F is a closed subset of T , of codimension at least 2.
Proof. We have already seen that each fixed point set has real codimension
at least 2 so we need only check that the union is locally finite. Fix a
compact set K ⊂ T . By the proper discontinuity of the action of MCG on
T , there cannot be an infinite set of finite order elements each with a fixed
point in K. Thus K is intersected by only finitely many of the fixed point
sets F (h), and so the union of these sets is closed. ⋄

We now finish the proof of ergodicity. Since the fixed point set has codi-
mension at least 2, the flow is defined almost everywhere on the quotient
M1 . If one has a positive measure invariant set in E ⊂ M1 then the lift
of E is a positive measure invariant measure set in T 1 T /MCG[k], which by
ergodicity must have 0 or full measure. The same is then true for E. Hence
the geodesic flow on M1 is ergodic.
Since the geodesic flow on T 1 T /MCG[k] covers the geodesic flow on a full
measure subset of M1 , it follows that the entropy of the flow on M1 is also
finite. This completes the proof of Theorem 1.

References
[1] D. V. Anosov, Geodesic flows on closed Riemannian manifolds of negative curvature,
Proc. Steklov Math. Inst. 90 (1967), 1–235.
[2] D. V. Anosov and Ya. G. Sinai, Certain smooth ergodic systems, Russian Math.
Surveys 22 (1967), 103–167.
[3] W. Ballman, M. Brin and K. Burns, On the differentiability of horocycles and horo-
cycle foliations. J. Diff. Geom. 26 (1987), 337–347.
[4] W. Ballman, M. Gromov and V. Schroeder, Manifolds of nonpositive curvature.
Birkhäuser, Boston 1985.
[5] M. Bridgeman Hausdorff dimension and the Weil-Petersson extension to quasifuch-
sian space arXiv:1002.1900
[6] J. Brock, H. Masur and Y. Minsky Asymptotics of Weil-Petersson geodesics, lami-
nations, recurrence, and flows, to appear GAFA
[7] M. Bridson and A. Haefliger. Metric spaces of non-positive curvature. Springer-
Verlag, Berlin, 1999.
[8] J. Cheeger and D. Ebin, Comparison theorems in Riemannian geometry. Revised
reprint of the 1975 original. AMS Chelsea Publishing, Providence, RI, (2008).
THE WEIL-PETERSSON GEODESIC FLOW 79

[9] G. Daskalopoulos, R.Wentworth,Classification of Weil-Petersson Isometries Amer-


ican J. Math 123 2003 941-975
[10] D. Dolgopyat, H.Hu and Ya. Pesin, An example of a smooth hyperbolic measure
with countably many ergodic components,Smooth ergodic theory and its applications
(Seattle, WA, 1999), Proc. Sympos. Pure Math., 69, Amer. Math. Soc., (2001).
[11] P. Eberlein, Geometry of nonpositively curved manifolds Chicago Lecture Notes in
Mathematics, University of Chicago Press (1996)
[12] U. Hamenstädt, Dynamical properties of the Weil-Petersson metric. Arxiv 0901.4301
[13] E. Hopf, Statistik der geodätischen Linien in Mannigfaltigkeiten negativer
Krümmung, Ber. Verh. Sächs. Akad. Wiss. Leipzig 91 (1939), 261–304.
[14] J. H. Hubbard, The monodromy of projective structures. Riemann surfaces and re-
lated topics: Proceedings of the 1978 Stony Brook Conference (State Univ. New
York, Stony Brook, N.Y., 1978), pp. 257–275, Ann. of Math. Stud., 97, Princeton
Univ. Press, Princeton, N.J., 1981.
[15] A. Katok, Infinitesimal Lyapunov functions, invariant cone families and stochastic
properties of smooth dynamical systems. With the collaboration of Keith Burns.
Ergodic Th. Dynam. Syst. 14 (1994), no. 4, 757–785.
[16] A. Katok and B. Hasselblatt, Introduction to the modern theory of dynamical sys-
tems. Encyclopedia of Mathematics and its Applications, 54. Cambridge University
Press, Cambridge, 1995. xviii+802 pp.
[17] A. Katok and J.-M. Strelcyn, Invariant manifolds, entropy and billiards; smooth
maps with singularities Lecture Notes in Mathematics 1222 Springer-Verlag (1980)
[18] O. Kowalski and M. Sekizawa, On tangent sphere bundles with small or large con-
stant radius, Ann. Global Anal. Geom., 18 (2000), 207-219.
[19] M. Mirzakhani, Simple closed geodesics on hyperbolic manifolds Annals of Math
(2008) 97-125
[20] C. T. McMullen, “Riemann surfaces, dynamics and geometry.” Course notes
[21] C. T. McMullen, The moduli space of Riemann surfaces is Kähler hyperbolic. Ann.
of Math. (2) 151 (2000), no. 1, 327–357.
[22] C. T. McMullen, Thermodynamics, dimension and the Weil-Petersson metric. In-
vent. Math. 173 (2008), no. 2, 365–425.
[23] H. Masur, Extension of the Weil-Petersson metric to the boundary of Teichmüller
space Duke Math J. 43 (1976) 623-635
[24] H. Masur and M. Wolf. The Weil-Petersson isometry group. Geom. Dedicata 93
(2002), 177-190.
[25] S. Nag, The complex analytic theory of Teichmüller Spaces. Canadian Math. Soc.
Series. Wiley-Interscience Publication. J. Wiley & Sons, New York. 1988.
[26] Ja. Pesin, Characteristic Ljapunov exponents, and ergodic properties of smooth dy-
namical systems with invariant measure. (Russian) Dokl. Akad. Nauk SSSR 226
(1976).
[27] M. Pollicott, H. Weiss and S. A. Wolpert. Topological dynamics of the Weil-Petersson
geodesic flow. [ps] arXiv:0711.3221
[28] M. Pollicott and H. Weiss, Ergodicity of geodesic flow on noncomplete negatively
curves surfaces, to appear, Asian Journal of Mathematics
[29] C. C. Pugh, The C 1+α hypothesis in Pesin theory. Inst. Hautes Études Sci. Publ.
Math. 59 (1984), 143–161.
[30] H. E. Rauch, A transcendental view of the space of algebraic Riemann surfaces. Bull.
Amer. Math. Soc. 71 (1965) 1–39.
[31] A. Sasaki, On the differential geometry of tangent bundles of Riemannian manifolds.
Tôhoku Math. J. 10 (1958), 338–354.
[32] Ya. Sinai, Dynamical systems with elastic reflections: ergodic properties of scattering
billiards, Russian Math. Surveys, 25 (1970), 137–189.
80 K. BURNS, H. MASUR AND A. WILKINSON

[33] S. Wolpert, Noncompleteness of the Weil-Petersson metric for Teichmüller space.


Pacific J. Math. 61 (1975), 573?577.
[34] A. Wolpert, The Fenchel-Nielsen Deformation, Annals of Math. 115 (1982) 501-528
[35] S. A. Wolpert, Extension of the Weil-Petersson connection. Duke Math. J. 146
(2009), no. 2, 281–303.
[36] S. A. Wolpert, Geometry of the Weil-Petersson completion of Teichmüller space.
Surveys in differential geometry, Vol. VIII (Boston, MA, 2002), 357–393, Surv. Dif-
fer. Geom., VIII, Int. Press, Somerville, MA, 2003.
[37] S. A. Wolpert, Understanding Weil-Petersson curvature. arXiv:0809.3699
[38] S. A. Wolpert, Behavior of geodesic-length functions on Teichmüller space. J. Dif-
ferential Geom. 79 (2008), no. 2, 277–334.
[39] S. A. Wolpert, The Weil-Petersson Hessian of length on Teichmüller space
arXiv:0902.0203
[40] S. A. Wolpert, Families of Riemann surfaces and Weil-Petersson geometry, CBMS
Regional Conference Series in Mathematics, Conference Board of the Mathematical
Sciences, Washington, DC
[41] M. Wojtkowski, Invariant families of cones and Lyapunov exponents, Ergod. Th.
Dynam. Syst. f 5 (1985), 145–161.
¯
[42] A. Zorich, Flat surfaces, Frontiers in Number Theory, Physics, and Geometry, P.
Cartier et al., ed. Springer-Verlag (2006).

You might also like