• Aucun résultat trouvé

On the Weil–Petersson Curvature of the Moduli Space of Riemann Surfaces of Large Genus

N/A
N/A
Protected

Academic year: 2022

Partager "On the Weil–Petersson Curvature of the Moduli Space of Riemann Surfaces of Large Genus"

Copied!
37
0
0

Texte intégral

(1)

Advance Access publication May 17, 2016 doi: 10.1093/imrn/rnw053

On the Weil–Petersson Curvature of the Moduli Space of Riemann Surfaces of Large Genus

Yunhui Wu

Department of Mathematics, Rice University, Houston, TX, 77005-1892, USA

Correspondence to be sent to: e-mail: yw22@rice.edu

Let Sg be a closed surface of genusgandMg be the moduli space ofSg endowed with the Weil–Petersson metric. In this article we investigate the Weil–Petersson curva- tures of Mg for large genusg. First, we study the asymptotic behavior of the extremal Weil–Petersson holomorphic sectional curvatures at certain thick surfaces in Mg as g → ∞. Then we prove two curvature properties on the whole space Mg as g → ∞ in a probabilistic way.

1 Introduction

Let Sg be a closed surface of genusg with g > 1, andMg be the moduli space of Sg. Endowed with the Weil–Petersson metric, the moduli space Mg is Kähler [1], incom- plete [6,31] and geodesically complete [33]. One can refer to the book [35] for the recent developments on Weil–Petersson geometry.

Tromba [30] and Wolpert [32] found a formula for the curvature tensor of the Weil–Petersson metric, which has been applied to study a variety of curvature properties of Mg over the past several decades. For example, the moduli space Mg has negative sectional curvature [30, 32], strongly negative curvature in the sense of Siu [27], dual Nakano negative curvature [16], and nonpositive definite Riemannian curvature operator

Received November 10, 2015; Revised February 18, 2016; Accepted March 8, 2016 Communicated by Prof. Maryam Mirzakhani

© The Author(s) 2016. Published by Oxford University Press. All rights reserved. For permissions, please e-mail: journals.permission@oup.com.

Downloaded from https://academic.oup.com/imrn/article-abstract/2017/4/1066/3048819 by Tsinghua University user on 22 November 2018

(2)

[39]. One can also refer to [4,10–12,14,15,17,29,34–36,38,40] for other aspects of the curvatures ofMg.

The subject of the asymptotic geometry ofMg asgtends to infinity has recently become quite active: see, for example, Mirzakhani [20–23] for the volume of Mg, Cavendish–Parlier [7] for the diameter ofMg, and Bromberg–Brock [3] for the least Weil–

Petersson translation length of pseudo-Anosov mapping classes. In terms of curvature bounds, by combining the results in Wolpert [32] and Teo [29], we may see that, restricted on the thick part of the moduli space, the scalar curvature is comparable to−gasggoes to infinity. The negative scalar curvature can be viewed as the1-norm of the Riemann- ian Weil–Petersson curvature operator. Thep(1p∞)-norm of the Weil–Petersson curvature operator was studied in [40] as gtends to infinity. For other related topics, one can also refer to [8,9,18,24,26,28,42] for more details.

We focus in this article on the asymptotic behavior for the Weil–Petersson sec- tional curvatures as the genusgtends to infinity. Tromba [30] and Wolpert [32] deduced from their formula that the Weil–Petersson holomorphic sectional curvature of Mg is bounded above by the constant2π(g−1)1 , which confirmed a conjecture of Royden in [25]. If one carefully checks their proofs, this upper bound2π(g11)can never be obtained: other- wise, there exists a harmonic Beltrami differential on a closed hyperbolic surface whose magnitude along the surface is a positive constant, which is impossible. As far as we know, the explicit optimal upper bound for the Weil–Petersson holomorphic sectional curvature is not known yet. The aim of this article is to study the Weil–Petersson curva- tures for large genus. Our first result tells that the rate−1g, lying in Tromba–Wolpert’s upper bound for Weil–Petersson holomorphic sectional curvature, is optimal asgtends to infinity. Before stating the theorem we make the following notation throughout this article.

Notation.We say

f1(g)f2(g)

if there exists a universal constantC>0, independent ofg, such that f2(g)

C f1(g)Cf2(g). Theorem 1.1. Given a constant

0>2 ln(3+2√ 2).

Downloaded from https://academic.oup.com/imrn/article-abstract/2017/4/1066/3048819 by Tsinghua University user on 22 November 2018

(3)

LetXg∈Mgbe a hyperbolic surface satisfying that the injectivity radius inj(Xg)0.

Then, the Weil–Petersson holomorphic sectional curvature HolK atXgsatisfies that

ν∈maxHBD(Xg)HolK(ν) −1 g,

where HBD(Xg)is the set of harmonic Beltrami differentials onXg. Buser and Sarnak proved in [5] that there exists a universal constantC>0 such that for all genusg2 there exists a hyperbolic surfaceYg∈Mgsuch that the injectivity radius inj(Yg)ofYg satisfies that

inj(Yg)Clng.

The following corollary is an immediate consequence of Theorem 1.1, Buser–

Sarnak’s above result, and Tromba–Wolpert’s upper bound for Weil–Petersson holo- morphic sectional curvature.

Corollary 1.2. The supremum of the Weil–Petersson holomorphic sectional curvature of the moduli spaceMg satisfies that

sup

XgMg max

ν∈HBD(Xg)HolK(ν) −1

g.

Theorem 1.8 in [40] says that the minimal Weil–Petersson holomorphic sectional curvature of a sufficiently thick hyperbolic surface (sufficiently thick means large injec- tivity radius) is comparable to−1, which answered a question of Mirzakhani. Combine Theorem 1.1with a refinement of the argument for the proof of Theorem 1.8 in [40], we get

Theorem 1.3. Given a constant

0>2 ln(3+2√ 2).

LetXg∈Mgbe a hyperbolic surface satisfying that the injectivity radius inj(Xg)0.

Downloaded from https://academic.oup.com/imrn/article-abstract/2017/4/1066/3048819 by Tsinghua University user on 22 November 2018

(4)

Then, the ratio of the minimal Weil–Petersson holomorphic sectional curvature over the maximal Weil–Petersson holomorphic sectional curvature atXg satisfies that

minν∈HBD(Xg)HolK(ν)

maxν∈HBD(Xg)HolK(ν) g.

There are recent suggestions that as the genus ggrows large, some regions in the moduli spaceMg should become increasingly flat. It was shown in [40] that this is not true from the view point of Riemannian curvature operator. Actually we showed in [40] that the-norm of the Riemannian Weil–Petersson curvature operator at every point inMgis uniformly bounded below away from zero. It is not known whether this phenomenon still holds for the-norm of the Riemannian Weil–Petersson sectional curvature.

Let Xg ∈ Mg and TXgMg be the tangent space of Mg at Xg. For sure TXgMg is identified with HBD(Xg)which is the set of harmonic Beltrami differentials onXg. Since the rest part of the introduction is on real Riemannian sectional curvatures, with abuse of notation we useTXgMginstead of HBD(Xg). For any two-dimensional planePTXgMg

(maybe not holomorphic), we denote byK(P)the Riemannian Weil–Petersson sectional curvature of the planeP. The following result (The author is grateful to Hugo Parlier for bringing to my attention the Weil–Petersson curvatures on random surfaces.) tells that, from the view point of Riemannian sectional curvature we also have that no region in the moduli spaceMg becomes increasingly flat asgtends to infinity. More precisely, Theorem 1.4. There exists a universal constant C0 > 0 such that the probability satisfies that

g→∞limProb{Xg∈Mg; min

P⊂TXgMgK(P)−C0<0} =1.

The proof of the theorem above requires a result due to Mirzakhani in [23], which says that a random Riemann surface will contain an arbitrarily large embedded hyperbolic geodesic ball asgtends to infinity.

Since Mg has negative sectional curvature [30,32], the following functionhis well defined.

h(Xg):= minPTXgMgK(P)

maxPTXgMgK(P), ∀Xg∈Mg.

Downloaded from https://academic.oup.com/imrn/article-abstract/2017/4/1066/3048819 by Tsinghua University user on 22 November 2018

(5)

The functionhabove is also well defined in any Riemannian manifold of negative (or positive) Riemannian sectional curvature. Recall that Zheng–Yau in [41] proved that a compact Kähler manifold with weakly four-pinched Riemannian sectional curvature (the range of his in [1, 4)) has nonpositive definite Riemannian curvature operator if the sectional curvature is negative. It is known that the Weil–Petersson metric of Mg

has negative sectional curvature [30,32] and nonpositive definite Riemannian curvature operator [39]. So it is interesting to study this functionhonMg.

It is clear that h(Xg) 1 for all Xg ∈ Mg. The results in [10, 34] tell that supXgMgh(Xg) = ∞. Indeed, one may choose a separating curve αSg and consider the direction along which the lengthαpinches to zero. Then the Weil–Petersson holo- morphic sectional curvature along the pinching direction will blow up as α → 0 (see [11, 34]). On the other hand, since α is separating, there exists arbitrary flat planes (see [10, 19]) near the stratum whose nodes have vanishing α-lengths. Thus, h is unbounded near certain part of the boundary of Mg. However, it is not clear about the range of h in the thick part of the moduli space. Our next result is that in a probabilistic way h is unbounded globally on Mg as g tends to infinity. More precisely,

Theorem 1.5. For anyL>0, then the probability satisfies that

glim→∞Prob{Xg ∈Mg;h(Xg)L} =1.

Contrast with Zheng–Yau’s result in [41], for large enoughg, almost no point in the moduli spaceMghas weakly four-pinched Riemannian sectional curvature although the Riemannian curvature operator ofMgis nonpositive definite [39].

Let X be a Riemann surface and μ be a harmonic Beltrami differen- tial on (X,σ (z)|dz|2). Recall that the -norm |μ|(X) of μ over X is given by

|μ|(X) := supp∈X|μ(p)| and the Weil–Petersson norm ||μ||WP is given by ||μ||WP :=

X|μ(z)|2σ (z)|dz|2. For the proofs of Theorems 1.1, 1.3, 1.4, and 1.5, the main idea is to construct harmonic Beltrami differentials on a Riemann surface such that the domain, on which the magnitudes of the harmonic Beltrami differentials are roughly their-norms, either has small area or has area which is comparable to the genusg.

The following result is crucial in the proofs of all the results above. It is also interesting on itself.

Downloaded from https://academic.oup.com/imrn/article-abstract/2017/4/1066/3048819 by Tsinghua University user on 22 November 2018

(6)

Theorem 1.6. Given a positive integern∈Z+and a constant 0>2 ln(3+2√

2).

Let Xg ∈ Mg be a hyperbolic surface. Assume that there exists a set of finite points {pi}ni=1Xgsatisfying that

(1) inj(pi)20, ∀1in.

(2) dist(pi,pj)0, ∀1i=j n, where dist(·,·)is the distance function on Xg.

Then, there exists a harmonic Beltrami differentialμ∈HBD(Xg)such that (a) |μ(pi)| |μ|(Xg)1, ∀1in.

(b) ||μ||2WP n.

Remark 1.7. Whenn=1 andXghas large injectivity radius, Theorem1.6was obtained in [40]. I am kindly told by S. Wolpert that the method in Section 2 of Chapter 8 in his book [35] can also lead to the existence of such a harmonic Beltrami differential for this special case thatn=1 andXghas large injectivity radius.

Plan of the paper.Section2provides some necessary background and the basic prop- erties of the Weil–Petersson metric that we will need. In Section 3 we construct the harmonic Beltrami differentials which hold for Theorem 1.6. We establish Theorem 1.6in Sections4 and5. Then we apply Theorem1.6to prove Theorems1.1and1.3in Section6. In Section7we will prove Theorems1.4and1.5.

2 Notations and Preliminaries

In this section we will set our notations and provide some necessary background material on surface theory and Weil–Petersson metric.

2.1 Hyperbolic disk

LetDbe the unit disk in the plane endowed with the hyperbolic metricρ(z)|dz|2, where ρ(z)= 4

(1− |z|2)2.

Downloaded from https://academic.oup.com/imrn/article-abstract/2017/4/1066/3048819 by Tsinghua University user on 22 November 2018

(7)

The distance to the origin is

distD(0,z)=ln1+ |z| 1− |z|.

For allr 0, letB(0;r)= {z ∈ D; distD(0,z) <r}andBeu(0;r)= {z ∈ D;|z| <r}. Then, the relation between the hyperbolic geodesic ball and Euclidean geodesic ball is given by the following equation.

B(0;r)=Beu

0;er−1 er+1

.

Let Aut(D)be the automorphism group ofD. For anyγ ∈Aut(D)there exist two constantsa∈Dandθ∈ [0, 2π)such that

γ (z)=exp(iθ) za 1−az.

The transitivity of the action of Aut(D)onDtells that for allz∈Dandγ ∈Aut(D), ρ(γ (z))|γ(z)|2 =ρ(z).

2.2 Bergman projection

In this subsection we briefly review the formula for the Bergman projection, which is a classical tool to construct harmonic Beltrami differentials on Riemann surfaces. One may refer to [1] for more details.

Let Xg be a hyperbolic surface and g be its associated Fuchsian group. A complex-valued function u on Dis called a measurable automorphic form of weight

−4 with respect togonDif it is a measurable function onDand satisfies that u(γz)γ(z)2 =u(z), ∀z∈D, γg.

If we allow a measurable automorphic formuof weight−4 to be holomorphic on D, then we call uis a holomorphic automorphic form of weight −4. We denote by A2(D,g)the complex vector space of all holomorphic automorphic functions of weight

−4 with respect tog, which is a(6g−6)-dimensional linear space.

LetBL2 (D,g)be the set of all measurable Beltrami automorphic forms of weight

−4 with respect togonDwith

||f||=esssupzD|f(z)|<∞,

Downloaded from https://academic.oup.com/imrn/article-abstract/2017/4/1066/3048819 by Tsinghua University user on 22 November 2018

(8)

wheref(z)= uρ(z)(z)for some measurable automorphic formu(z)of weight−4 with respect togonD. One may refer to Chapter 7 of [13] for more details onBL2 (D,g).

Recall the Bergman Kernel functionK(z,ξ)of the unit diskDis given by K(z,ξ)= 12

π(1zξ)4 = n=0

2

π(n+1)(n+2)(n+3)(zξ)n, (2.1) wherezandξ are arbitrary inD.

A direct computation gives that

K(γz,γξ)γ(z)2γ(ξ)2=K(z,ξ) (2.2) for allγ ∈Aut(D).

The Bergman projectionβ2ofBL2 (D,g)on toA2(D,g)is given by the following theorem.

Theorem 2.1([1], Formula (1.18)). For anyfBL2(D,g). Letξ =x+yi∈Dand set 2f)(z)=

D

f(ξ)K(z,ξ)dxdy,z∈D.

Then we have

β2fA2(D,g).

Proof. One can also see Theorem 7.3 in [13].

2.3 Surfaces and Weil–Petersson metric

Let Sg be a closed surface of genus g 2 and Tg be the Teichmüller space of Sg. The tangent space at a pointXg =(Sg,σ(z)|dz|2)is identified with the space of harmonic Bel- trami differentials onXgwhich are forms ofμ=ψσ, whereψis a holomorphic quadratic differential onXg. Let dA(z)=σ (z)dxdybe the volume form ofXg =(Sg,σ(z)|dz|2), where z = x+yi. TheWeil–Petersson metric is the Hermitian metric on Tg arising from the Petersson scalar product

< ϕ,ψ >WP=

S

ϕ(z) σ (z)

ψ(z) σ(z)dA(z)

via duality. We will concern ourselves primarily with its Riemannian part gWP. Let Teich(Sg)denote the Teichmüller space endowed with the Weil–Petersson metric. The

Downloaded from https://academic.oup.com/imrn/article-abstract/2017/4/1066/3048819 by Tsinghua University user on 22 November 2018

(9)

mapping class group Mod(Sg)acts properly discontinuously on Teich(Sg)by isometries.

The moduli spaceMgof Riemann surfaces, endowed with the Weil–Petersson metric, is defined as

Mg:=Teich(Sg)/Mod(Sg).

The following proposition has been proved in a lot of literature. For example, one can refer to [12,29,38]. We use the following form which is proven by Teo through using the Taylor series expansion for a holomorphic function.

Proposition 2.2 ([29], Proposition 3.1). Let Xg ∈ Mg and μTXgMg be a harmonic Beltrami differential ofXg. Then, for anypXg and 0<rinj(p),

|μ(p)|2 C1(r)

B(p;r)|μ(z)|2dA(z),

where the constantC1(r)=(43π(1((1+4eerr)2)3)))1 andB(p;r)Xgis the geodesic ball of

radiusrcentered atp.

Proof. One can also see Proposition 2.10 in [40].

One may refer to [13,35] for more details on the Weil–Petersson metric.

2.4 Riemannian tensor of the Weil–Petersson metric

The Weil–Petersson curvature tensor is given by the following. Letμα,μβbe two elements in the tangent space atXg, and

gαβ =

Xg

μα·μβdA.

For the inverse of(gij), we use the convention gijgkj=δik. The curvature tensor is given by

Rijkl = 2

∂tk∂tlgijgst

∂tkgit

∂tlgsj.

Let D = −2( −2)1, where is the Beltrami–Laplace operator on Xg = (Sg,σ (z)|dz|2). The following curvature formula was established by Tromba and Wolpert

Downloaded from https://academic.oup.com/imrn/article-abstract/2017/4/1066/3048819 by Tsinghua University user on 22 November 2018

(10)

independently in [30,32], which has been applied to study various curvature properties of the Weil–Petersson metric in the past 30 years.

Theorem 2.3(Tromba–Wolpert). The curvature tensor satisfies Rijkl=

Xg

D(μiμj)·kμl)dA+

Xg

D(μiμl)·kμj)dA.

Recall that a holomorphic sectional curvature is a Riemannian sectional curva- ture along a holomorphic plane. Thus, Theorem2.3gives that for everyμTXgMg, the Weil–Petersson holomorphic sectional curvature HolK(μ)along the holomorphic plane spanned byμis

HolK(μ)=−2

XgD(|μ|2)· |μ|2dA

||μ||4WP

.

We end this section by the following proposition, whose proof relies on Proposi- tion2.2, Lemma 5.1 in [37] and the Cauchy–Schwartz inequality. This proposition will be applied several times in this article. The statement is slightly different from Proposition 2.11 in [40].

Proposition 2.4. Let Xg ∈ Mg and μTXgMg be a harmonic Beltrami differential of Xg. Then, the Weil–Petersson holomorphic sectional curvature HolK(μ)satisfies that for anypXg,

−2 supz∈X|μ(z)|2

||μ||2WP

HolK(μ)−C2(inj(p))|μ(p)|4

||μ||4WP

,

where the constantC2(inj(p)) >0 only depends on the injectivity radius inj(p)atp.

Proof. It follows from the same argument as the proof of Proposition 2.11 in [40]. We

leave it as an exercise.

3 Construction of the Objective Harmonic Beltrami differentials

In this section we will construct the harmonic Beltrami differentials which hold for Theorem1.6.

First we deal with the casen =1 in Theorem1.6. LetXg ∈ Mg be a hyperbolic surface, pXg and inj(p) be the injectivity radius of Xg at p. For any constant r

Downloaded from https://academic.oup.com/imrn/article-abstract/2017/4/1066/3048819 by Tsinghua University user on 22 November 2018

(11)

(0, inj(p)], we consider the characteristic function

ν0(z):=

⎧⎨

1, ∀z∈B(p;r).

0, otherwise.

WhereB(p;r)Xgis the geodesic ball of radiusrcentered atp.

Consider the covering mapπ : D→ Xg. Up to a conjugation, we lift pto 0 ∈ D and letg denote its associated Fuchsian group. Then, it is not hard to see thatν0 can be lifted toν0BL2 (D,g)satisfying that for allγg,

ν0(z):=

⎧⎨

γ−1z)

γ−1z), ∀zγB(0;r). 0, otherwise.

(3.1)

We apply the Bergman projectionβ2toν0.

Lemma 3.1. Letν0BL2 (D,g)given in equation (3.1) . Then, we have 2ν0)(z)=12

er−1 er+1

2

γg

γ(z)2.

Proof. The proof is a direct computation.

Since 0<rinj(p), we have

γ1B(0;r)γ2B(0;r)= ∅, ∀γ1 =γ2g.

Letξ =x+yi∈D. Theorem2.1gives that, for allz∈D, 2ν0)(z)=

Dν0(ξ)K(z,ξ)dxdy

=

γg

γ◦B(0;r)

γ−1ξ)

γ1ξ)K(z,ξ)dxdy

=

γg

B(0;r)

γ(ξ)

γ(ξ)K(z,γξ)|γ(ξ)|2dxdy.

Equation (2.2) tells that

K(z,γξ)= K(γ1z,ξ) γ1z))2γ(ξ)2.

Downloaded from https://academic.oup.com/imrn/article-abstract/2017/4/1066/3048819 by Tsinghua University user on 22 November 2018

(12)

Recall that

K(z,ξ)= n=0

2

π(n+1)(n+2)(n+3)(zξ)n. Hence,

2ν0)(z)=

γg

B(0;r)γ(ξ)2K(γ1z,ξ) 1

1z))2γ(ξ)2dxdy

=

γg

n=0

2

π(n+1)(n+2)(n+3)

×

B(0;r)

1

1z))2((γ1z)ξ)ndxdy

=12 π

γg

1 1z))2

Beu(0;er−1 er+1)

dxdy

=12

er−1 er+1

2

γ∈g

γ(z)2,

where the last equality applies the fact that 1z))2= 1

1)(z)2, ∀γ ∈g.

Remark 3.2. When the surface has big enough injectivity radius, it was shown in [40]

that the Weil–Petersson holomorphic sectional curvature along the holomorphic plane spanned by the holomorphic quadratic differential

γ∈gγ(z)2 is comparable to the maximal Weil–Petersson holomorphic sectional curvature of the moduli space at this

surface. Moreover, it is comparable to−1.

Let p,qXg be two points with dist(p,q) 2r > 0, where r is a constant satisfying that

0<rmin{inj(p), inj(q)}. (3.2) We liftpandqto 0 andqinD, respectively, which satisfies that

distD(0,q)=dist(p,q)2r. (3.3)

Downloaded from https://academic.oup.com/imrn/article-abstract/2017/4/1066/3048819 by Tsinghua University user on 22 November 2018

(13)

Letσq∈Aut(D)withσq(0)=q. Actually one may choose σq(z)= z+q

1+qz, ∀z∈D.

We define a functionν1BL2 (D,g)as follows. For allγg,

ν1(z):=

⎧⎪

⎪⎪

⎪⎨

⎪⎪

⎪⎪

γ−1z)

γ−1◦z), ∀z∈γB(0;r).

γ−1z)

γ−1z)×σσq((γ◦σq)−1z)

q((γ◦σq)−1z), ∀zγB(q;r). 0, otherwise.

(3.4)

Equations (3.2) and (3.3) tells thatν1(z)is well defined onD. Lemma 3.3. For anyz∈D, we have

γ∈g

γB(q;r)

γ1ξ) γ1ξ)

σq((γσq)−1ξ)

σq((γσq)1ξ)K(z,ξ)dxdy

=12

er−1 er+1

2

γg

q−1γ )(z)2.

Proof. Sinceg⊂Aut(D),

γB(q;r)=γσqB(0;r) ∀γ ∈g. Letξ =x+yi∈D. Then, for allz∈Dwe have

γg

γB(q;r)

γ−1ξ)

γ1ξ)×σq((γσq)1ξ)

σq((γσq)1ξ)×K(z,ξ)dxdy

=

γ∈g

γ◦σq◦B(0;r)

γ1ξ)

γ1ξ)×σq((γσq)−1ξ)

σq((γσq)1ξ)×K(z,ξ)dxdy

=

γ∈g

B(0;r)

γqξ)

γqξ)×σq(ξ)

σq(ξ) ×K(z,γσqξ)× |(γ ◦σq)(ξ)|2dxdy

=

γ∈g

B(0;r)

γqξ)

γqξ)×σq(ξ) σq(ξ)

× K((γσq)1z,ξ)

((γσq)((γσq)1z))2·((γσq)(ξ))2 × |(γ ◦σq)(ξ)|2dxdy

Downloaded from https://academic.oup.com/imrn/article-abstract/2017/4/1066/3048819 by Tsinghua University user on 22 November 2018

(14)

=

γ∈g

1

((γσq)((γσq)1z))2

B(0;r)

K((γσq)−1z,ξ)dxdy

=

γ∈g

1

((γσq)((γσq)1z))2

×

n=0

2

π(n+1)(n+2)(n+3)

B(0;r)((γσq)1z·ξ)ndxdy

=

γ∈g

1

((γσq)((γσq)1z))2 ×12 π

Beu(0;eerr−1+1)

dxdy

=

γ∈g

1

((γσq)((γσq)1z))2 ×12

er−1 er+1

2

=12

er−1 er+1

2

γg

((γσq)1)(z)2

=12

er−1 er+1

2

γg

q1γ )(z)2.

Now we apply the Bergman projectionβ2toν1(z).

First from our assumptions on equations (3.2) and (3.3) we know that the balls in{γ ◦B(0;r),γB(q;r)}γg are pairwisely disjoint. Thus, Lemmas3.1and3.3tell that

Lemma 3.4. For allz∈D, we have

2ν1)(z)=12

er−1 er+1

2

(

γg

q1γ )(z)2+

γ∈g

γ(z)2).

Similarly we generalize the construction above for any finite subset inXg, which is the remaining part of this section.

Definition 3.5. Given two constantsn∈Z+and >0, a finite set of points{pi}ni=1Xg

is said to be(,n)-separated if

dist(pi,pj), ∀1i=jn. (3.5)

Downloaded from https://academic.oup.com/imrn/article-abstract/2017/4/1066/3048819 by Tsinghua University user on 22 November 2018

(15)

A finite set of points {pi}ni=1Xg is called an-net of Xg if the set of points {pi}ni=1Xgare(,n)-separated and

ni=1B(pi;)=Xg. (3.6) Letr>0 be a constant and{pi}ni=1Xgbe a(2r,n)-separated finite set of points satisfying that

1minin{inj(pi)}r. (3.7) We lift p1 to the originp1 = 0 ∈ D. Let g be its associated Fuchsian group andF be the Dirichlet fundamental domain centered at 0 w.r.tg. We also lift{pi}ni=2to {pi}ni=2F, respectively. Thus, for all 1i,jn,

distD(pi,pj)dist(pi,pj). (3.8) For 1in, letσpi ∈Aut(D)withσpi(0)=pi. For sure one may choose

σpi(z)= z+pi

1+piz, ∀z∈D.

In particularσp1 is the identity map. That is,σp1(z)=zfor allz∈D.

Similar to equation (3.4) we define a functionνnBL2 (D,g). More precisely, for allγgand 1in,

νn(z):=

⎧⎪

⎪⎩

γ−1◦z)

γ−1◦z)×σσpi ((γ◦σpi)−1z)

pi((γ◦σpi)−1z), ∀z∈γB(pi;r).

0, otherwise.

(3.9)

Proposition 3.6. For anyz∈D, we have 2νn)(z)=12

er−1 er+1

2n i=1

γg

pi1γ )(z)2.

Proof. Since{pi}1in are(2r,n)-separated, equation (3.7) tells that γ1B(pi;r)γ2B(pj;r)= ∅, ∀γ1 =γ2g, or i=j∈ [1,n].

Then, the conclusion follows from the same computation as the proof of Lemma3.3.

Downloaded from https://academic.oup.com/imrn/article-abstract/2017/4/1066/3048819 by Tsinghua University user on 22 November 2018

(16)

In the following two sections, we will prove that the harmonic Beltrami differential

n i=1

γ∈gpi−1◦γ )(z)2 ρ(z) dz

dz holds for Theorem1.6.

4 Two Bounds

In this section, we use the same notations in Section3.

For each positive integeri∈ [1,n], we define

μi(z):=

γ∈gpi1γ )(z)2

ρ(z) , ∀z∈D. (4.1)

Whereρ(z)= (1−|4z|2)2 is the scalar function of the hyperbolic metric on the unit disk.

The following computation follows from the idea of Ahlfors in [2] (one can also see [40] for an English version).

Ahlfors’ method:From the triangle inequality we know that

i(z)|

γg

|(σpi1γ )(z)|2

ρ(z) . (4.2)

Then sinceρ(γ (z))|γ(z)|2=ρ(z)for anyγ ∈Aut(D), andρ(ζ )=4(1− |ζ|2)−2, we have

γ∈g

|(σp−1iγ )(z)|2

ρ(z) = 1

4

γ∈g

(1− |(σp−1iγ )(z)|2)2. (4.3)

The inequalities above yields that for allz∈D,

i(z)| 1 4

γg

(1− |(σpi1γ )(z)|2)2. (4.4)

Letbe the (Euclidean) Laplace operator on the (Euclidean) disk. Then a direct computation shows that

(4.5)

(

γ∈g

(1− |(σpi1γ )(z)|2)2=8·

γ∈g

(2|(σpi1γ (z)|2−1)|(σpi1γ )(z)|2.

Note that the terms on the right side are non-negative when |σpi1γ (z)|212. With that in mind, recall thatBeu(0;1

2) := {z ∈ D; |z| < 12}is the ball of Euclidean radius1

2, letVi:= ∪γ∈gγ1σpiBeu(0;1

2)be the pullbacks of this ballBeu(0;1 2). The

Downloaded from https://academic.oup.com/imrn/article-abstract/2017/4/1066/3048819 by Tsinghua University user on 22 November 2018

(17)

equation above gives that

γg(1− |(σpi1γ )(z)|2)2is subharmonic inD−Vi. Since both

γ∈g(1− |(σpi1γ )(z)|2)2andViareg-invariant, andg is cocompact, we find supz∈D

γg

(1− |(σpi1γ )(z)|2)2 =sup

z∈Vi

γ∈g

(1− |(σpi1γ )(z)|2)2

= sup

z∈σpiBeu(0;1 2)

γ∈g

(1− |(σpi1γ )(z)|2)2 (4.6)

which in particular is bounded above by a constant depending ong andpi.

Recall the relation between the Euclidean distance and the hyperbolic distance is distD(0,z)=ln1+ |z|

1− |z|.

Sinceσpi∈Aut(D),σpiBeu(0;12)is the hyperbolic geodesic ballB(pi; ln(3+2√ 2)) of radius ln(3+2√

2)centered atpi. Hence, equation (4.6) is equivalent to

(4.7) sup

zD

γ∈g

(1− |(σpi1γ )(z)|2)2= sup

zB(pi;ln(3+2 2))

γg

(1− |(σpi1γ )(z)|2)2.

4.1 An upper bound function

Set

μ(z)= n

i=1

μi(z)= n

i=1

γg

pi1γ )(z)2

ρ(z) , ∀z∈D. (4.8)

Where{μi}1in are given in equation (4.1).

Similar as equation (4.4) we have

|μ(z)|1 4

n i=1

γ∈g

(1− |(σpi1γ )(z)|2)2, ∀z∈D. (4.9)

Define the right side function to be f(z):=1

4 n

i=1

γ∈g

(1− |(σpi1γ )(z)|2)2, ∀z∈D. (4.10)

From the definition we know that f is a g-invariant function in D, which descends into a function on the hyperbolic surface Xg = D/ g. The following propo- sition is crucial in this article. We will apply it to prove Theorem1.6. The proof follows from a similar argument for the proof of equation (4.7).

Downloaded from https://academic.oup.com/imrn/article-abstract/2017/4/1066/3048819 by Tsinghua University user on 22 November 2018

(18)

Proposition 4.1. The functionf satisfies that sup

zD f(z)= sup

z∈∪ni=1B(pi;ln(3+2 2))

f(z).

Proof. For 1inandz∈D, set fi(z)=

γ∈g

(1− |(σpi1γ )(z)|2)2.

Equation (4.5) tells that the function fi is subharmonic in the complement (∪γgγ1B(pi; ln(3+2√

2)))cof(∪γ∈gγ1B(pi, ln(3+2√

2)))inD. Sincef =n

i=1fi, we have

f(z)0, ∀z∈ ∩ni=1

γ∈gγ1B pi; ln

3+2√ 2c

. That is,

f(z)0, ∀z∈

γgni=1γ1B pi; ln

3+2√

2 c

. Sincef isg-invariant, it follows from the Maximal-Principle that

sup

z

γ∈gni=1γ−1B(pi;ln(3+22))c

f(z)= sup

z∈∪ni=1S(pi;ln(3+2 2))

f(z),

whereS(pi; ln(3+2√

2))= {q∈D; distD(q,pi)=ln(3+2√ 2)}. Sincef isg-invariant,

sup

z

γ∈gni=1γ−1B(pi;ln(3+22))

f(z)= sup

z∈∪ni=1B(pi;ln(3+22))f(z).

Therefore,

sup

zD

f(z)= sup

z∈∪ni=1B(pi;ln(3+22))f(z). 4.2 Bounds forf when0>2 ln(3+2√

2)

Given a positive constant0with

0>2 ln(3+2√ 2).

Downloaded from https://academic.oup.com/imrn/article-abstract/2017/4/1066/3048819 by Tsinghua University user on 22 November 2018

Références

Documents relatifs

We define the intersection i(g, p) of the Riemannian metric of negative curvature g with the geodesic current ^ as the total mass of SM for the measure ^p(^).. The following theorem

Toute utilisa- tion commerciale ou impression systématique est constitutive d’une in- fraction pénale.. Toute copie ou impression de ce fichier doit conte- nir la présente mention

relation between Enriques surfaces, plane quintic curves with two nodes and plane quintic curves with a cusp. Also a plane quintic curve with a cusp is

Theorem 1.6 says that the Weil–Petersson volume of any Weil–Petersson geodesic ball in the thick part of moduli space will decay to 0 as g tends to infinity.. It would be

We call a K¨ ahler orbifold to be of Hodge type, if the K¨ ahler form is the Chern form of a positive orbifold line bundle.A theorem of Baily [2] states that in this case the

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des

We investigate some differential geometric aspects of the space of Bridgeland stability conditions on a Calabi–Yau triangulated category The aim is to give a provisional definition

We give an application of the main theorem to the com- pletions of Teichm¨ uller space with respect to a class of metrics that generalize the Weil-Petersson