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

Optimal lower bounds for first eigenvalues of Riemann surfaces for large genus

Yunhui Wu and Yuhao Xue Yau Mathematical Sciences Center, Tsinghua University, Haidian District, Beijing 100084, China [email protected] [email protected]
Abstract.

In this article we study the first eigenvalues of closed Riemann surfaces for large genus. We show that for every closed Riemann surface XgX_{g} of genus gg (g2)(g\geq 2), the first eigenvalue of XgX_{g} is greater than 1(Xg)g2\frac{\mathcal{L}_{1}(X_{g})}{g^{2}} up to a uniform positive constant multiplication. Where 1(Xg)\mathcal{L}_{1}(X_{g}) is the shortest length of multi closed curves separating XgX_{g}. Moreover,we also show that this new lower bound is optimal as gg\to\infty.

1. Introduction

For any integer g2g\geq 2, let g\mathcal{M}_{g} be the moduli space of closed Riemann surfaces of genus gg and XggX_{g}\in\mathcal{M}_{g} be a closed hyperbolic surface of genus gg. The spectrum of the Laplacian on XgX_{g} is a fascinating topic in several mathematical fields including analysis, geometry, number theory, topology and so on for a long time. The spectrum of XgX_{g} is a discrete closed subset in 0\mathbb{R}^{\geq 0} and consists of eigenvalues with finite multiplicity. We enumerate them, counted with multiplicity, in the following increasing order

0=λ0(Xg)<λ1(Xg)λ2(Xg).0=\lambda_{0}(X_{g})<\lambda_{1}(X_{g})\leq\lambda_{2}(X_{g})\leq\cdots.

Buser [4] showed that for any constant ε>0\varepsilon>0, there exists a hyperbolic surface 𝒳g\mathcal{X}_{g} of genus gg such that λ2g3(𝒳g)<ε\lambda_{2g-3}(\mathcal{X}_{g})<\varepsilon and λn(𝒳g)<14+ε\lambda_{n}(\mathcal{X}_{g})<\frac{1}{4}+\varepsilon for any n(2g2)n\geq(2g-2). Recently Otal-Rosas [14] showed that λ2g2(Xg)>14\lambda_{2g-2}(X_{g})>\frac{1}{4} for any XggX_{g}\in\mathcal{M}_{g}. One may also see Ballmann-Matthiesen-Mondal [1, 2] and Mondal [12] for more general statements on λ2g2(Xg)\lambda_{2g-2}(X_{g}).

Definition.

For any XggX_{g}\in\mathcal{M}_{g} and integer i[1,2g3]i\in[1,2g-3], we define a positive quantity i(Xg)\mathcal{L}_{i}(X_{g}) of XgX_{g} to be minimal possible sum of the lengths of simple closed geodesics in XgX_{g} which cut XgX_{g} into i+1i+1 pieces.

The quantity i(Xg)\mathcal{L}_{i}(X_{g}) can be arbitrarily closed to 0 for certain XggX_{g}\in\mathcal{M}_{g}. Schoen-Wolpert-Yau [15] showed that the ii-th eigenvalue of XgX_{g} is comparable to the quantity i(Xg)\mathcal{L}_{i}(X_{g}) of XgX_{g} above. More precisely, they showed that for any integer i[1,2g3]i\in[1,2g-3], there exists two constants αi(g)>0\alpha_{i}(g)>0 and βi(g)>0\beta_{i}(g)>0, depending on gg, such that for any XggX_{g}\in\mathcal{M}_{g},

(1) αi(g)λi(Xg)i(Xg)βi(g).\displaystyle\alpha_{i}(g)\leq\frac{\lambda_{i}(X_{g})}{\mathcal{L}_{i}(X_{g})}\leq\beta_{i}(g).

One may see Dodziuk-Randol [10] for a different proof of Schoen-Wolpert-Yau’s theorem, and see Dodziuk-Pignataro-Randol-Sullivan [9] on similar results for Riemann surfaces with punctures. The upper bounds in (1) follow by suitable choices of test functions on certain collars, whose central closed geodesics are part of a pants decomposition of XgX_{g} whose boundary curves have bounded lengths in terms of gg, which is due to Bers [3]. For proving the lower bounds in (1), one essential step is to show that λ1(Xg)α1(g)1(Xg)\lambda_{1}(X_{g})\geq\alpha_{1}(g)\cdot\mathcal{L}_{1}(X_{g}). Which was applied in [10] to obtain the lower bounds in (1) for other eigenvalues λi(Xg)(2i2g3)\lambda_{i}(X_{g})\ (2\leq i\leq 2g-3), together by using a mini-max principle.

In this paper we study the asymptotic behavior of the constant α1(g)\alpha_{1}(g) in (1) for large genus. The method in this article is motivated by [10] of Dodziuk-Randol. We prove

Theorem 1.

For every g2g\geq 2, there exists a uniform constant K1>0K_{1}>0 independent of gg such that for any hyperbolic surface XggX_{g}\in\mathcal{M}_{g}, the first eigenvalue λ1(Xg)\lambda_{1}(X_{g}) of XgX_{g} satisfies that

λ1(Xg)K11(Xg)g2.\lambda_{1}(X_{g})\geq K_{1}\cdot\frac{\mathcal{L}_{1}(X_{g})}{g^{2}}.

By [15] we know that Theorem 1 is optimal as 1(Xg)0\mathcal{L}_{1}(X_{g})\to 0. Actually Theorem 1 is also optimal as gg\to\infty: in [17] we constructed a hyperbolic surface 𝒳gg\mathcal{X}_{g}\in\mathcal{M}_{g} for all g2g\geq 2 such that the first eigenvalue satisfies that

λ1(𝒳g)K21(𝒳g)g2\lambda_{1}(\mathcal{X}_{g})\leq K_{2}\cdot\frac{\mathcal{L}_{1}(\mathcal{X}_{g})}{g^{2}}

where K2>0K_{2}>0 is a uniform constant independent of gg. In Section 5 we will discuss the details to see that Theorem 1 is optimal simultaneously as 1(Xg)0\mathcal{L}_{1}(X_{g})\to 0 and gg\to\infty.

One open question

It is interesting to study the optimal asymptotic behaviors of the other constants αi(g)(2i2g3)\alpha_{i}(g)\ (2\leq i\leq 2g-3) in Schoen-Wolpert-Yau’s theorem as gg\to\infty. The answer to the following question is unknown.

Question.

For every g2g\geq 2, does there exist a uniform constant K>0K>0 independent of gg such that for any hyperbolic surface XggX_{g}\in\mathcal{M}_{g} and i[1,2g3]i\in[1,2g-3], the ii-th eigenvalue λi(Xg)\lambda_{i}(X_{g}) of XgX_{g} satisfies that

λi(Xg)Kii(Xg)g2?\lambda_{i}(X_{g})\geq K\cdot\frac{i\cdot\mathcal{L}_{i}(X_{g})}{g^{2}}?
Remark.

Theorem 1 affirmatively answers the question above for case i=1i=1.

Plan of the paper.

The paper is organized as follows. In Section 2 we collect preliminaries for closed hyperbolic surfaces and provide several necessary technical lemmas. In Section 3 we follow [10] to prove that if the first eigenvalue λ1(Xg)\lambda_{1}(X_{g}) is small, then there exists a uniform gap in terms of gg between the oscillation of the first eigenfunction on certain modified thick part and the total oscillations of the first eigenfunction on all components of this modified thick part, which is Proposition 15. For this reason, the first eigenfunction has large oscillation on the modified thin part consisting of certain collars. By definition the first eigenvalue is greater than the energy of its eigenfunction on the modified thin part. Then in Section 4 by using some combinatorial discussion and estimations in previous sections, we finish the proof of Theorem 1. In the last Section we discuss an example to see that the lower bound in Theorem 1 is optimal as gg goes to \infty.

Acknowledgements.

The authors would like to thank Long Jin for helpful discussions. The first named author is partially supported by a grant from Tsinghua University. We are grateful to one referee to point out a mistake in our original statement of Lemma 13. We are also grateful to another referee for useful comments and suggestions which are helpful.

2. Preliminaries

In this section we will set up the notations and provide some necessary background on two-dimensional hyperbolic geometry and spectrum theory of hyperbolic surfaces.

2.1. Hyperbolic surfaces

Let XgX_{g} be a closed hyperbolic surface of genus g2g\geq 2 and γXg\gamma\subset X_{g} be a non-trivial loop. There always exists a unique closed geodesic, still denoted by γ\gamma, representing this loop. The Collar Lemma says that it has a tubular neighborhood which is a topological cylinder with a standard hyperbolic metric. And the width of this cylinder, only depending on the length of γ\gamma, goes to infinity as the length of γ\gamma goes to 0. We use the following version [5, Theorem 4.1.1] of the Collar Lemma.

Lemma 2 (Collar lemma).

Let γ1,γ2,,γm\gamma_{1},\gamma_{2},...,\gamma_{m} be disjoint simple closed geodesics on a closed hyperbolic Riemann surface XgX_{g}, and (γi)\ell(\gamma_{i}) be the length of γi\gamma_{i}. Then m3g3m\leq 3g-3 and we can define the collar of γi\gamma_{i} by

T(γi)={xXg;dist(x,γi)w(γi)}T(\gamma_{i})=\{x\in X_{g};\ \mathop{\rm dist}(x,\gamma_{i})\leq w(\gamma_{i})\}

where

w(γi)=arcsinh1sinh12(γi)w(\gamma_{i})=\mathop{\rm arcsinh}\frac{1}{\sinh\frac{1}{2}\ell(\gamma_{i})}

is the half width of the collar.

Then the collars are pairwise disjoint for i=1,,mi=1,...,m. Each T(γi)T(\gamma_{i}) is isomorphic to a cylinder (ρ,t)[w(γi),w(γi)]×𝕊1(\rho,t)\in[-w(\gamma_{i}),w(\gamma_{i})]\times\mathbb{S}^{1}, where 𝕊1=/\mathbb{S}^{1}=\mathbb{R}/\mathbb{Z}, with the metric

(2) ds2=dρ2+(γi)2cosh2ρdt2.\displaystyle ds^{2}=d\rho^{2}+\ell(\gamma_{i})^{2}\cosh^{2}\rho dt^{2}.

And for a point (ρ,t)(\rho,t), the point (0,t)(0,t) is its projection on the geodesic γi\gamma_{i}, |ρ|\lvert\rho\rvert is the distance to γi\gamma_{i}, tt is the coordinate on γi𝕊1\gamma_{i}\cong\mathbb{S}^{1}.

As the length (γ)\ell(\gamma) of the central closed geodesic goes to 0, the width

(3) ew(γ)4(γ)\displaystyle e^{w(\gamma)}\sim\frac{4}{\ell(\gamma)}

which tends to infinity. In this paper, we mainly deal with the case that (γ)\ell(\gamma) is small and so w(γ)w(\gamma) is large.

Note that a collar is homeomorphic to a cylinder which may have arbitrary large width. When saying a collar in this paper, we will not always assume it has the maximal width given in Lemma 2. Actually we will use a subcollar with a slightly shorter width. Now we make certain elementary computations on a collar. Assume γXg\gamma\subset X_{g} is a closed geodesic of length >0\ell>0. There are two types of coordinates on a collar of γ\gamma. One is (ρ,t)(\rho,t) given in Lemma 2. The other one is the polar coordinate (r,θ)(r,\theta) on the upper half plane \mathbb{H}. Which may be written as

(4) {ρ=logtanθ2t=1logr.\begin{cases}\rho&=-\log\tan\frac{\theta}{2}\\ t&=\frac{1}{\ell}\log r.\end{cases}

For any two points zz and ww\in\mathbb{H}, the hyperbolic distance dist(z,w)\mathop{\rm dist}_{\mathbb{H}}(z,w) satisfies that

(5) coshdist(z,w)=1+|zw|22ImzImw.\cosh\text{dist}_{\mathbb{H}}(z,w)=1+\frac{|z-w|^{2}}{2\mathop{\rm Im}z\mathop{\rm Im}w}.

For any two points (ρ,0)(\rho,0) and (ρ,t)(\rho,t) in the coordinate given in Lemma 2 which have the same distance to the center closed geodesic γ\gamma, the curve s(ρ,s)s\mapsto(\rho,s) where 0st0\leq s\leq t is not a geodesic. However one may compute the length L((ρ,0),(ρ,t))L((\rho,0),(\rho,t)) of the geodesic homotopic to that curve (see figure 1) by using (5). Actually we have

coshL((ρ,0),(ρ,t))=coshdist(eiθ,reiθ)=1+|r1|22rsin2θ\cosh L((\rho,0),(\rho,t))=\cosh\text{dist}_{\mathbb{H}}(e^{i\theta},re^{i\theta})=1+\frac{|r-1|^{2}}{2r\sin^{2}\theta}

where (r,θ)(r,\theta) is given in (4).

ρ\rhott(ρ,0)(\rho,0)(ρ,t)(\rho,t)geodesicthe curve s(ρ,s)s\mapsto(\rho,s)
Figure 1.

By applying (4) we have

(6) sinhL((ρ,0),(ρ,t))2=sinht2coshρ.\sinh\frac{L((\rho,0),(\rho,t))}{2}=\sinh\frac{t\ell}{2}\cosh\rho.

In particular, the injectivity radius inj(ρ,0))\mathop{\rm inj}(\rho,0)) at the point (ρ,0)(\rho,0) satisfies

(7) sinhinj((ρ,0))=sinhL((ρ,0),(ρ,1))2=sinh2coshρ.\sinh\mathop{\rm inj}((\rho,0))=\sinh\frac{L((\rho,0),(\rho,1))}{2}=\sinh\frac{\ell}{2}\cosh\rho.

Let [w,w]×𝕊1[-w,w]\times\mathbb{S}^{1} endowed with the hyperbolic metric given in (2). We assume that the center closed geodesic has length of \ell and the width is ww which may not be the maximal width given in Lemma 2. Then the hyperbolic volume of the collar [w,w]×𝕊1[-w,w]\times\mathbb{S}^{1} is

(8) Vol([w,w]×𝕊1)\displaystyle\mathop{\rm Vol}([-w,w]\times\mathbb{S}^{1}) =01wwcoshρdρdt\displaystyle=\int_{0}^{1}\int_{-w}^{w}\ell\cosh\rho d\rho dt
=2sinhw.\displaystyle=2\ell\sinh w.

One may refer to [5] for more details in this subsection.

2.2. Thick-thin decomposition

In this subsection, we make a thick-thin part decomposition of XggX_{g}\in\mathcal{M}_{g}. For any pXgp\in X_{g} we let inj(p)\mathop{\rm inj}(p) denote the injectivity radius of XgX_{g} at pp. For any given constant ε>0\varepsilon>0, we define

Xgε:={pXg;inj(p)ε}X_{g}^{\geq\varepsilon}:=\{p\in X_{g};\ \mathop{\rm inj}(p)\geq\varepsilon\}

to be the ε\varepsilon-thick part of XgX_{g}, and its complement

Xg<ε:={pXg;inj(p)<ε}X_{g}^{<\varepsilon}:=\{p\in X_{g};\ \mathop{\rm inj}(p)<\varepsilon\}

to be the ε\varepsilon-thin part of XgX_{g}. By the Collar Lemma 2, we know that for a small ε>0\varepsilon>0, the thin part Xg<εX_{g}^{<\varepsilon} consists of certain disjoint collars (or empty).

In this paper we use a modified thick-thin decomposition. More precisely, for a small enough given constant ε>0\varepsilon>0 (given in Lemma 4), we consider all closed geodesics of length less than 2ε2\varepsilon which are denoted by γ1,,γm\gamma_{1},...,\gamma_{m} for some m0m\geq 0\in\mathbb{Z}. Consider the collar TiT_{i} of each γi\gamma_{i} defined by

(9) Ti:={xXg;dist(x,γi)wi}T_{i}:=\{x\in X_{g};\ \mathop{\rm dist}(x,\gamma_{i})\leq w_{i}\}

where i\ell_{i} is the length of γi\gamma_{i} and wiw_{i} is the width of TiT_{i} given by

(10) 0wi=max{0,arcsinh1sinh12i2}.0\leq w_{i}=\max\{0,\mathop{\rm arcsinh}\frac{1}{\sinh\frac{1}{2}\ell_{i}}-2\}.

Here the width wiw_{i} is less that the maximal width w(γi)w(\gamma_{i}) in Lemma 2. Thus, for small enough ε>0\varepsilon>0, {Ti}1im\{T_{i}\}_{1\leq i\leq m} are also pairwisely disjoint.

Definition 3.

We define

(11) :=i=1mTi.\mathcal{B}:=\bigcup_{i=1}^{m}T_{i}.

which is called the ε\varepsilon-modified thin part of XgX_{g}. And we define

(12) 𝒜:=Xg¯\mathcal{A}:=\overline{X_{g}\setminus\mathcal{B}}

which is called the ε\varepsilon-modified thick part of XgX_{g}. Furthermore, for each γi\gamma_{i}, we define

(13) Si:={xXg;widist(x,γi)wi+1}S_{i}:=\{x\in X_{g};\ w_{i}\leq\mathop{\rm dist}(x,\gamma_{i})\leq w_{i}+1\}

which is called the shell of TiT_{i}.

For the modified ε\varepsilon-thick-thin decomposition above, we have the following properties which will be applied in the proof of Proposition 12 and 14 to deal with some technical details. They just follow from some elementary computations on collars.

Lemma 4.

There exists a uniform constant ε>0\varepsilon>0 independent of gg such that for every hyperbolic surface XgX_{g} of genus gg, the modified ε\varepsilon-thick-thin decomposition Xg=𝒜X_{g}=\mathcal{A}\cup\mathcal{B} satisfies the following properties.

  1. (1)

    The width of a collar given by (10) satisfies

    wi1>ε.w_{i}\geq 1>\varepsilon.
  2. (2)

    The closed geodesics γ1,,γm\gamma_{1},...,\gamma_{m}\in\mathcal{B} are disjoint. The corresponding T1S1,,TmSmT_{1}\cup S_{1},...,T_{m}\cup S_{m} are also disjoint. Moreover, for each i[1,m]i\in[1,m], Si𝒜S_{i}\subset\mathcal{A}.

  3. (3)

    The volumes of all collars TiT_{i} and shells SiS_{i} are bounded. More precisely,

    12Vol(Ti)4and12Vol(Si)4.\frac{1}{2}\leq\mathop{\rm Vol}(T_{i})\leq 4\ \ \ \text{and}\ \ \ \frac{1}{2}\leq\mathop{\rm Vol}(S_{i})\leq 4.
  4. (4)

    For each point p𝒜p\in\mathcal{A},

    inj(p)ε.\mathop{\rm inj}(p)\geq\varepsilon.
  5. (5)

    For any η(0,ε]\eta\in(0,\varepsilon], if two points pp and qq in one component of 𝒜\mathcal{A} have distance dist(p,q)=η\mathop{\rm dist}(p,q)=\eta in XgX_{g}, then there exists a path connecting pp and qq which is contained in the component such that it has length less than 5η5\eta.

Proof.

(i) Recall that the length i\ell_{i} of the closed geodesic γi\gamma_{i} is less than 2ε2\varepsilon. The width wiw_{i} is given by (10). So we only need ε>0\varepsilon>0 to be small enough such that

(14) arcsinh1sinhε21>ε.\mathop{\rm arcsinh}\frac{1}{\sinh\varepsilon}-2\geq 1>\varepsilon.

(ii) By Lemma 2, for ε<arcsinh1\varepsilon<\mathop{\rm arcsinh}1 the closed geodesics {γ1,,γm}\{\gamma_{1},...,\gamma_{m}\} are disjoint. By definition we know that for each 1im1\leq i\leq m,

TiSi={xXg;dist(x,γi)arcsinh(1sinh12i)1}T_{i}\cup S_{i}=\{x\in X_{g};\ \mathop{\rm dist}(x,\gamma_{i})\leq\mathop{\rm arcsinh}(\frac{1}{\sinh\frac{1}{2}\ell_{i}})-1\}

which is contained in the collar of center closed geodesic γi\gamma_{i} with maximal width arcsinh(1sinh12i)\mathop{\rm arcsinh}(\frac{1}{\sinh\frac{1}{2}\ell_{i}}). By Lemma 2 we know that {T1S1,,TmSm}\{T_{1}\cup S_{1},...,T_{m}\cup S_{m}\} are also disjoint. By definition we clearly have that Si𝒜S_{i}\subset\mathcal{A} for each i[1,m]i\in[1,m].

(iii) By (8), we have

Vol(Ti)=2isinhwi\mathop{\rm Vol}(T_{i})=2\ell_{i}\sinh w_{i}

and

Vol(Si)=2isinh(wi+1)2isinhwi.\mathop{\rm Vol}(S_{i})=2\ell_{i}\sinh(w_{i}+1)-2\ell_{i}\sinh w_{i}.

So we have

Vol(Ti)+Vol(Si)=2isinh(wi+1)2isinh12i4.\mathop{\rm Vol}(T_{i})+\mathop{\rm Vol}(S_{i})=2\ell_{i}\sinh(w_{i}+1)\leq\frac{2\ell_{i}}{\sinh\frac{1}{2}\ell_{i}}\leq 4.

As functions of i\ell_{i}, by using (3) we have

limi0Vol(Ti)=limi02isinh(log(4i)2)=4e2,\lim_{\ell_{i}\to 0}\mathop{\rm Vol}(T_{i})=\lim_{\ell_{i}\to 0}2\ell_{i}\cdot\sinh(\log(\frac{4}{\ell_{i}})-2)=\frac{4}{e^{2}},

and

limi0Vol(Si)=4(e1)e2.\lim_{\ell_{i}\to 0}\mathop{\rm Vol}(S_{i})=\frac{4(e-1)}{e^{2}}.

Thus, for a small enough constant ε>0\varepsilon>0 we have

(15) 12Vol(Ti)4and12Vol(Si)4.\frac{1}{2}\leq\mathop{\rm Vol}(T_{i})\leq 4\ \ \text{and}\ \ \frac{1}{2}\leq\mathop{\rm Vol}(S_{i})\leq 4.

(iv) For a point p𝒜p\in\mathcal{A}, suppose that inj(p)<ε\mathop{\rm inj}(p)<\varepsilon. Then there exists a geodesic loop α\alpha based at pp with length (α)<2ε\ell(\alpha)<2\varepsilon. This closed curve α\alpha should be homotopic to a closed geodesic with length less than 2ε2\varepsilon, so it is one of the γ1,,γm\gamma_{1},...,\gamma_{m} denoted by γi\gamma_{i}. In the collar of such a geodesic γi\gamma_{i}, one may assume the coordinate of pp is (ρ,0)(\rho,0) with ρwi\rho\geq w_{i}. For small enough ε>0\varepsilon>0 satisfying

sinh(ε)2ε,\sinh(\varepsilon)\leq 2\varepsilon,

by (7) we have

2ε\displaystyle 2\varepsilon >\displaystyle> (α)sinh(12(α))sinh(inj(p))sinhi2coshρ\displaystyle\ell(\alpha)\geq\sinh(\frac{1}{2}\ell(\alpha))\geq\sinh(\mathop{\rm inj}(p))\geq\sinh\frac{\ell_{i}}{2}\cosh\rho
\displaystyle\geq sinhi2coshwisinhi2cosh(arcsinh(1sinh12i))e2\displaystyle\sinh\frac{\ell_{i}}{2}\cosh w_{i}\geq\sinh\frac{\ell_{i}}{2}\cosh(\mathop{\rm arcsinh}(\frac{1}{\sinh\frac{1}{2}\ell_{i}}))\cdot e^{-2}
\displaystyle\geq 1e2\displaystyle\frac{1}{e^{2}}

where we apply the inequality cosh(xy)cosh(x)ey\cosh(x-y)\geq\cosh(x)e^{-y} for all x,y0x,y\geq 0. So if

(17) ε<12e2,\varepsilon<\frac{1}{2e^{2}},

then inj(p)ε\mathop{\rm inj}(p)\geq\varepsilon for all p𝒜p\in\mathcal{A}. Remark here it is easy to see that sinh(ε)2ε\sinh(\varepsilon)\leq 2\varepsilon if 0<ε<12e20<\varepsilon<\frac{1}{2e^{2}}.

(v) Assume that pp and qq are in one component of 𝒜\mathcal{A} which have distance

dist(p,q)=ηε.\mathop{\rm dist}(p,q)=\eta\leq\varepsilon.

If the shortest geodesic connecting pp and qq lies in 𝒜\mathcal{A}, then the distance from pp to qq in 𝒜\mathcal{A} is exactly η\eta. We are done in this case. So one may assume that the shortest geodesic α\alpha connecting pp and qq crosses the boundaries of 𝒜\mathcal{A}. Now we assume ε>0\varepsilon>0 satisfies (i)(iv)(i)-(iv). In particular, ε<12e2\varepsilon<\frac{1}{2e^{2}}. Since dist(p,q)ε<1\mathop{\rm dist}(p,q)\leq\varepsilon<1, these two points pp and qq must lie in one certain shell SiS_{i} for some i[1,m]i\in[1,m]. Assume in the coordinate on collar p=(ρ1,t1)p=(\rho_{1},t_{1}), q=(ρ2,t2)q=(\rho_{2},t_{2}) and 0<ρ1ρ20<\rho_{1}\leq\rho_{2}. Let r=(ρ1,t2)Si𝒜r=(\rho_{1},t_{2})\in S_{i}\subset\mathcal{A}. We Consider the curve

α2:s(s,t2)\alpha_{2}:s\mapsto(s,t_{2})

connecting rr and qq in 𝒜\mathcal{A} and the curve

β:s(ρ1,s)\beta:s\mapsto(\rho_{1},s)

connecting pp and rr in 𝒜\mathcal{A} such that βα2\beta\cup\alpha_{2} is homotopic to α\alpha. By the structure of collar and shell, α2\alpha_{2} is a shortest geodesic but β\beta is not a geodesic. Consider the shortest geodesic α1\alpha_{1} connecting pp and rr which is homotopic to β\beta.

ρ=wi\rho=w_{i}ρ=wi+1\rho=w_{i}+1pprrqqβ\betaα1\alpha_{1}α2\alpha_{2}α\alpha
Figure 2.

The geodesic α2\alpha_{2} achieves the distance between {(ρ,t)Si;ρ=ρ1}\{(\rho,t)\in S_{i};\ \rho=\rho_{1}\} and{(ρ,t)Si;ρ=ρ2}\{(\rho,t)\in S_{i};\ \rho=\rho_{2}\}. So we have

(α2)dist(p,q)=η.\ell(\alpha_{2})\leq\mathop{\rm dist}(p,q)=\eta.

Since α1\alpha_{1} is homotopic to αα21\alpha\cup\alpha_{2}^{-1} and α1\alpha_{1} is a shortest geodesic, by the triangle inequality we have

(α1)(α)+(α2)2η.\ell(\alpha_{1})\leq\ell(\alpha)+\ell(\alpha_{2})\leq 2\eta.

By (6),

sinh(α1)2=sinh|t1t2|i2coshρ1.\sinh\frac{\ell(\alpha_{1})}{2}=\sinh\frac{|t_{1}-t_{2}|\ell_{i}}{2}\cosh\rho_{1}.

If ε>0\varepsilon>0 is small enough and satisfies

sinhε2ε,\sinh\varepsilon\leq 2\varepsilon,

then

(α1)|t1t2|i2coshρ1\ell(\alpha_{1})\geq\frac{|t_{1}-t_{2}|\ell_{i}}{2}\cosh\rho_{1}

because (α1)2ηε\frac{\ell(\alpha_{1})}{2}\leq\eta\leq\varepsilon and |t1t2|i2i2ε\frac{|t_{1}-t_{2}|\ell_{i}}{2}\leq\frac{\ell_{i}}{2}\leq\varepsilon. On the other hand,

(β)\displaystyle\ell(\beta) =|t1t2icoshρ1ds|\displaystyle=|\int_{t_{1}}^{t_{2}}\ell_{i}\cosh\rho_{1}ds|
=|t1t2|icoshρ1.\displaystyle=|t_{1}-t_{2}|\ell_{i}\cosh\rho_{1}.

So we have

(β)2(α1).\ell(\beta)\leq 2\ell(\alpha_{1}).

Clearly βα2𝒜\beta\cup\alpha_{2}\subset\mathcal{A} is a path connecting pp and qq whose length satisfies that

(β)+(α2)2(α1)+η5η.\displaystyle\ell(\beta)+\ell(\alpha_{2})\leq 2\ell(\alpha_{1})+\eta\leq 5\eta.

Which completes the proof. ∎

Remark.
  1. (1)

    The uniform constant ε>0\varepsilon>0 only needs to satisfy (14), (15) and (17). For example, one may choose

    ε=0.05.\varepsilon=0.05.
  2. (2)

    The bounds in this lemma are not optimal. But they are good enough to be applied to prove our later propositions.

2.3. An upper bound for i(Xg)\mathcal{L}_{i}(X_{g}) (1i(2g3))(1\leq i\leq(2g-3))

Recall that for all g2g\geq 2 the Bers’ constant LgL_{g} is the best possible constant such that for each hyperbolic surface XgX_{g} of genus gg there exists a pants decomposition whose boundary curves have length less than LgL_{g}. One may see [5, Chapter 5] for more details. The following result is due to Bers.

Theorem 5 (Bers, [3]).

For each g2g\geq 2,

Lg26(g1).L_{g}\leq 26(g-1).

In this subsection we apply Theorem 5 to get an upper bound for 1(Xg)\mathcal{L}_{1}(X_{g}) which will be applied to prove Theorem 1. More precisely, we show

Lemma 6.

For every hyperbolic surface XgX_{g} of genus gg and i[1,2g3]i\in[1,2g-3],

i(Xg)78i(g1).\mathcal{L}_{i}(X_{g})\leq 78\cdot i\cdot(g-1).
Proof.

Let {𝒫j}1j(2g2)\{\mathcal{P}_{j}\}_{1\leq j\leq(2g-2)} be a pants decomposition of XgX_{g}, where each 𝒫j\mathcal{P}_{j} is a pair of pants, such that

(18) max1j(2g2)maxα𝒫jα(Xg)Lg.\displaystyle\max_{1\leq j\leq(2g-2)}\max_{\alpha\in\partial\mathcal{P}_{j}}\ell_{\alpha}(X_{g})\leq L_{g}.

For each 1i(2g3)1\leq i\leq(2g-3), we set

S:={α;α𝒫jfor all 1ji}.S:=\{\alpha;\ \alpha\in\partial\mathcal{P}_{j}\ \emph{for all $1\leq j\leq i$}\}.

It is not hard to see that XgSX_{g}\setminus S has components containing all the 𝒫j\mathcal{P}_{j}’s where 1ji1\leq j\leq i and their complement Xg(1ji𝒫j)X_{g}\setminus\left(\cup_{1\leq j\leq i}\mathcal{P}_{j}\right). Thus, SS divides XgX_{g} into at least (i+1)(i+1) components. By construction we have

(19) #S3i.\displaystyle\#S\leq 3\cdot i.

By the definition of i(Xg)\mathcal{L}_{i}(X_{g}) we know that

(20) i(Xg)αSα(Xg).\displaystyle\mathcal{L}_{i}(X_{g})\leq\sum_{\alpha\in S}\ell_{\alpha}(X_{g}).

Then it follows by (18), (19), (20) and Theorem 5 that

(21) i(Xg)3iLg78i(g1).\displaystyle\mathcal{L}_{i}(X_{g})\leq 3\cdot i\cdot L_{g}\leq 78\cdot i\cdot(g-1).

Which completes the proof. ∎

Remark.
  1. (1)

    It would be interesting to study the asymptotic behavior of the quantity supXggi(Xg)\sup_{X_{g}\in\mathcal{M}_{g}}\mathcal{L}_{i}(X_{g}), where 1i(2g3)1\leq i\leq(2g-3), as gg\to\infty. To our best knowledge, it is even unkown for supXgg1(Xg)\sup_{X_{g}\in\mathcal{M}_{g}}\mathcal{L}_{1}(X_{g}) as gg\to\infty. After this article was submitted, recently we show in [13] that supXgg1(Xg)Cln(g)\sup_{X_{g}\in\mathcal{M}_{g}}\mathcal{L}_{1}(X_{g})\leq C\ln(g) for all g2g\geq 2 and some universal constant C>0C>0 independent of gg.

  2. (2)

    It is known [5, Theorem 5.1.4] that the Bers constant Lg6g2L_{g}\geq\sqrt{6g}-2. As gg\to\infty, the asymptotic behavior of LgL_{g} is still unknown. The upper bounds in (1) of Schoen-Yau-Wolpert depend on LgL_{g}.

2.4. Eigenvalues

Let XgX_{g} be a closed Riemann surface of genus g2g\geq 2 which can also be viewed as a hyperbolic metric on XgX_{g}. Let Δ\Delta be the Laplacian with respect to this metric. A number λ\lambda is called an eigenvalue if Δf+λf=0\Delta f+\lambda\cdot f=0 on XgX_{g} for some non-zero function ff on XgX_{g}. And the corresponding function ff is called an eigenfunction. It is known that the set of eigenvalues is an infinite sequence of non-negative numbers

0=λ0(Xg)<λ1(Xg)λ2(Xg).0=\lambda_{0}(X_{g})<\lambda_{1}(X_{g})\leq\lambda_{2}(X_{g})\leq\cdots.

Let {fi}i0\{f_{i}\}_{i\geq 0} be its corresponding orthonormal sequence of eigenfunctions. Clearly f0f_{0} is the constant function 14π(g1)\frac{1}{\sqrt{4\pi(g-1)}}. The mini-max principle tells that for any integer k0k\geq 0,

λk(Xg)=inf{Xg|f|2Xgf2; 0fH1(Xg)andXgffi=0i[0,k1]}\displaystyle\lambda_{k}(X_{g})=\inf\{\frac{\int_{X_{g}}|\nabla f|^{2}}{\int_{X_{g}}f^{2}};\ 0\neq f\in H^{1}(X_{g})\ \text{and}\ \int_{X_{g}}f\cdot f_{i}=0\ \forall i\in[0,k-1]\}

where H1(Xg)H^{1}(X_{g}) is the completion under the H1H^{1}-norm of the space of smooth functions on XgX_{g}. One may see [6] for more details.

2.5. Energy bounds on collars

In this subsection we recall several useful energy bounds in [10] of certain functions on collars. And we will make a little modification for one of them.

The first one is a special case of [10, Lemma 1] for n=2n=2.

Lemma 7.

[10, Lemma 1] Let XggX_{g}\in\mathcal{M}_{g} be a hyperbolic surface and T=[w,w]×𝕊1T=[-w,w]\times\mathbb{S}^{1} be a collar of XgX_{g}. Let λ1(T)\lambda_{1}(T) be the first Dirichlet eigenvalue for TT. Then

λ1(T)>14.\lambda_{1}(T)>\frac{1}{4}.

The second one is as follows.

Lemma 8.

[10, Lemma 3] Let XggX_{g}\in\mathcal{M}_{g} be a hyperbolic surface, and T=[w,w]×𝕊1T=[-w,w]\times\mathbb{S}^{1} be a collar of XgX_{g} with center closed geodesic of length \ell. Denote by Γ1\Gamma_{1} and Γ2\Gamma_{2} the two boundary components of TT, which are topologically circles. Suppose a smooth function ff on TT satisfies

min(x,x)Γ1×Γ2|f(x)f(x)|=c0\min_{(x,x^{*})\in\Gamma_{1}\times\Gamma_{2}}|f(x)-f(x^{*})|=c\geq 0

where xΓ2x^{*}\in\Gamma_{2} is the reflection of xΓ1x\in\Gamma_{1} through the center closed geodesic. Then

T|f|2c24.\int_{T}|\nabla f|^{2}\geq\frac{c^{2}}{4}\ell.

The third one is a slightly different version of [10, Lemma 2].

Lemma 9.

Let XggX_{g}\in\mathcal{M}_{g} be a hyperbolic surface, and T=[w,w]×𝕊1T=[-w,w]\times\mathbb{S}^{1} be a collar of XgX_{g} with shell S=[w1,w]×𝕊1[w,w+1]×𝕊1S=[-w-1,-w]\times\mathbb{S}^{1}\cup[w,w+1]\times\mathbb{S}^{1}. Let δ\delta be a constant with 0<δ<1160<\delta<\frac{1}{16}, c>0c>0 be a constant and ff be a smooth function on TST\cup S satisfying:

  1. (1)

    T|f|2c>0\int_{T}|f|^{2}\geq c>0,

  2. (2)

    S|f|2δc\int_{S}|f|^{2}\leq\delta c,

  3. (3)

    S|f|2δc\int_{S}|\nabla f|^{2}\leq\delta c.

Then

T|f|2116δ4c.\int_{T}|\nabla f|^{2}\geq\frac{1-16\delta}{4}c.
Proof.

We follow the argument in the proof of [10, Lemma 2]. First recall that the Collar Lemma 2 tells that one may assume (ρ,t)[w1,w+1]×𝕊1(\rho,t)\in[-w-1,w+1]\times\mathbb{S}^{1} is a coordinate on TST\cup S. And the hyperbolic metric on TST\cup S is

ds2=dρ2+2cosh2ρdt2ds^{2}=d\rho^{2}+\ell^{2}\cosh^{2}\rho dt^{2}

where \ell is the length of center closed geodesic. In this coordinate we define a function FF on TST\cup S as

F(ρ,t):={f(ρ,t)if|ρ|w,(w+1|ρ|)f(ρ,t)if|ρ|w.F(\rho,t):=\begin{cases}f(\rho,t)&\text{if}\ |\rho|\leq w,\\ (w+1-|\rho|)f(\rho,t)&\text{if}\ |\rho|\geq w.\end{cases}

By definition F|(TS)=0F|_{\partial(T\cup S)}=0. Then it follows by Lemma 7 that

(22) TS|F|2>14TSF2.\int_{T\cup S}|\nabla F|^{2}>\frac{1}{4}\int_{T\cup S}F^{2}.

It is clear that |(w+1|ρ|)|2=||ρ||2=1|\nabla(w+1-|\rho|)|^{2}=|\nabla|\rho||^{2}=1 on SS. So we have

S|F|2\displaystyle\int_{S}|\nabla F|^{2} =\displaystyle= S|(w+1|ρ|)f(ρ,t)+(w+1|ρ|)f(ρ,t)|2\displaystyle\int_{S}|\nabla(w+1-|\rho|)\cdot f(\rho,t)+(w+1-|\rho|)\cdot\nabla f(\rho,t)|^{2}
\displaystyle\leq S(|f(ρ,t)|+(w+1|ρ|)|f(ρ,t)|)2\displaystyle\int_{S}(|f(\rho,t)|+(w+1-|\rho|)\cdot|\nabla f(\rho,t)|)^{2}
\displaystyle\leq 2Sf2+2S|f|2(by Cauchy-Schwarz inequality)\displaystyle 2\int_{S}f^{2}+2\int_{S}|\nabla f|^{2}\quad\emph{(by Cauchy-Schwarz inequality)}
\displaystyle\leq 4δc.\displaystyle 4\delta c.

Where assumption (2)(2) and (3)(3) are applied in the last inequality. Thus, it follows by (22) and (2.5) that

T|f|2\displaystyle\int_{T}|\nabla f|^{2} =\displaystyle= T|F|2\displaystyle\int_{T}|\nabla F|^{2}
\displaystyle\geq 14TSF2S|F|2\displaystyle\frac{1}{4}\int_{T\cup S}F^{2}-\int_{S}|\nabla F|^{2}
\displaystyle\geq 14c4δc(by assumption (1))\displaystyle\frac{1}{4}c-4\delta c\quad\quad\emph{(by assumption $(1)$)}
=\displaystyle= 116δ4c.\displaystyle\frac{1-16\delta}{4}c.

Which completes the proof. ∎

3. Uniform gaps for eigenfunctions

Recall that the Cheeger isoperimetric constant h(Xg)h(X_{g}) is defined as

h(Xg):=inflength(Γ)min{Vol(A1),Vol(A2)}h(X_{g}):=\inf\frac{\mathop{\rm length}(\Gamma)}{\min\{\mathop{\rm Vol}(A_{1}),\mathop{\rm Vol}(A_{2})\}}

where the infimum is taken over all smooth curves Γ\Gamma which divide XgX_{g} into two pieces A1A_{1} and A2A_{2}.

Lemma 10 (Cheeger inequality, [7]).

Then

λ1(Xg)14h2(Xg).\lambda_{1}(X_{g})\geq\frac{1}{4}h^{2}(X_{g}).

Let Γ\Gamma be a set of smooth curves dividing XgX_{g} into two disjoint pieces A1A_{1} and A2A_{2}. Then Γ\Gamma must be one of the following three cases:

(a)(a). Γ\Gamma contains a simple closed curve bounding a disk DD in XgX_{g}.

(b)(b). Γ\Gamma contains two simple closed curves τ\tau and γ\gamma which bounds a cylinder TT in XgX_{g}.

(c)(c). Γ\Gamma is not of type (a) and (b). That is, no two pairwise simple closed curves in Γ\Gamma are homotopic, and no simple closed curve in Γ\Gamma is homotopically trivial. In particular, length(Γ)1(Xg)\mathop{\rm length}(\Gamma)\geq\mathcal{L}_{1}(X_{g}).

For cases (a)(a) and (b)(b), it follows by elementary isoperimetric inequalities that length(Γ)Vol(D)\mathop{\rm length}(\Gamma)\geq\mathop{\rm Vol}(D) and length(Γ)Vol(T)\mathop{\rm length}(\Gamma)\geq\mathop{\rm Vol}(T). For case (c)(c), we have length(Γ)min{Vol(A1),Vol(A2)}1(Xg)Vol(Xg)\frac{\mathop{\rm length}(\Gamma)}{\min\{\mathop{\rm Vol}(A_{1}),\mathop{\rm Vol}(A_{2})\}}\geq\frac{\mathcal{L}_{1}(X_{g})}{\mathop{\rm Vol}(X_{g})}. Thus it follows by the Cheeger inequality that

(25) λ1(Xg)min{14,1(Xg)24Vol(Xg)2}.\lambda_{1}(X_{g})\geq\min\{\frac{1}{4},\frac{\mathcal{L}_{1}(X_{g})^{2}}{4\mathop{\rm Vol}(X_{g})^{2}}\}.

One may see the proof of [17, Proposition 9] for more details on (25).

In light of (25), to prove λ1(Xg)min{14,c1(Xg)Vol(Xg)2}\lambda_{1}(X_{g})\geq\min\{\frac{1}{4},c\frac{\mathcal{L}_{1}(X_{g})}{\mathop{\rm Vol}(X_{g})^{2}}\} we only need to consider the case that 1(Xg)\mathcal{L}_{1}(X_{g}) is small. The method in this article is motivated by [10] of Dodziuk-Randol, which gave a different proof on the main results of Schoen-Yau-Wolpert in [15].

For fixed small enough constant ε>0\varepsilon>0 given by Lemma 4 (for example ε=0.05\varepsilon=0.05), we always assume

(26) 1(Xg)ε.\displaystyle\mathcal{L}_{1}(X_{g})\leq\varepsilon.

In particular the modified thin part \mathcal{B} defined in Section 2.2 is non-empty.

Let φ\varphi be an eigenfunction with respect to the first eigenvalue λ1\lambda_{1} on XgX_{g} such that

Xgφ2=1.\int_{X_{g}}\varphi^{2}=1.

Denote the components of the modified thick part 𝒜Xg\mathcal{A}\subset X_{g} by M1,,MmM_{1},...,M_{m} where m>0m>0\in\mathbb{Z}. For each integer i[1,m]i\in[1,m] we set

Mi^={pXg;dist(p,Mi)ε}.\widehat{M_{i}}=\{p\in X_{g};\ \mathop{\rm dist}(p,M_{i})\leq\varepsilon\}.

When ε\varepsilon is small enough, these subsets M1^,,Mm^\widehat{M_{1}},...,\widehat{M_{m}} are still pairwise disjoint.

Definition 11.

The oscillation of φ\varphi on each component MiM_{i}, denoted by Osc(i)\mathop{\rm Osc}(i), is defined as

Osc(i):=maxxMiφ(x)minxMiφ(x).\mathop{\rm Osc}(i):=\max_{x\in M_{i}}\varphi(x)-\min_{x\in M_{i}}\varphi(x).

We have the following bound for i=1mOsc(i)\sum_{i=1}^{m}\mathop{\rm Osc}(i) when λ1(Xg)\lambda_{1}(X_{g}) is small.

Proposition 12.

Let XggX_{g}\in\mathcal{M}_{g} be a hyperbolic surface with λ1(Xg)14\lambda_{1}(X_{g})\leq\frac{1}{4}, and φ\varphi be an eigenfunction with respect to λ1(Xg)\lambda_{1}(X_{g}) with Xgφ2=1\int_{X_{g}}\varphi^{2}=1. Then there exists a constant c=c(ε)>0c=c(\varepsilon)>0 only depending on ε\varepsilon such that

i=1mOsc(i)cVol(Xg)λ1(Xg).\sum_{i=1}^{m}\mathop{\rm Osc}(i)\leq c\sqrt{\mathop{\rm Vol}(X_{g})\lambda_{1}(X_{g})}.
Proof.

On every component Mi0M_{i_{0}} of 𝒜\mathcal{A}, we assume p,qMi0p,q\in M_{i_{0}} with

φ(p)=maxxMi0φ(x)andφ(q)=minxMi0φ(x).\varphi(p)=\max_{x\in M_{i_{0}}}\varphi(x)\quad\emph{and}\quad\varphi(q)=\min_{x\in M_{i_{0}}}\varphi(x).

Consider a shortest path γ:[0,l]Mi0\gamma:[0,l]\to M_{i_{0}} in Mi0M_{i_{0}} connecting pp and qq with arc-parameter, where l>0l>0 is the length of γ\gamma. Let 0=t1<t2<<tk=l0=t_{1}<t_{2}<...<t_{k}=l be a partition of [0,l][0,l] where k>0k>0\in\mathbb{Z} satisfying ti+1ti=12εt_{i+1}-t_{i}=\frac{1}{2}\varepsilon for all 1ik21\leq i\leq k-2 and tktk112εt_{k}-t_{k-1}\leq\frac{1}{2}\varepsilon. Let pi=γ(ti)p_{i}=\gamma(t_{i}) and consider the embedding balls B(pi;ε)B(p_{i};\varepsilon) centered at pip_{i} of radius ε\varepsilon. By standard Sobolev embeddings [16] we know that

φL(B(pi;ε2))c(ε)j=0NΔj(dφ)L2(B(pi;ε))||\nabla\varphi||_{L^{\infty}(B(p_{i};\frac{\varepsilon}{2}))}\leq c(\varepsilon)\sum_{j=0}^{N}||\Delta^{j}(d\varphi)||_{L^{2}(B(p_{i};\varepsilon))}

for some integer N>0N>0 (we remark here that the Sobolev embedding holds with a uniform constant because it follows by by Lemma 4 that the injectivity radius satisfies that inj(pi)ε\mathop{\rm inj}(p_{i})\geq\varepsilon). Since φ\varphi is an eigenfunction and λ1(Xg)<14\lambda_{1}(X_{g})<\frac{1}{4},

(27) φL(B(pi;ε2))\displaystyle||\nabla\varphi||_{L^{\infty}(B(p_{i};\frac{\varepsilon}{2}))} c(ε)j=0NΔj(dφ)L2(B(pi;ε))\displaystyle\leq c(\varepsilon)\sum_{j=0}^{N}||\Delta^{j}(d\varphi)||_{L^{2}(B(p_{i};\varepsilon))}
=c(ε)j=0Nλ1jdφL2(B(pi;ε))\displaystyle=c(\varepsilon)\sum_{j=0}^{N}\lambda_{1}^{j}||d\varphi||_{L^{2}(B(p_{i};\varepsilon))}
43c(ε)φL2(B(pi;ε)).\displaystyle\leq\frac{4}{3}c(\varepsilon)||\nabla\varphi||_{L^{2}(B(p_{i};\varepsilon))}.

Such an estimation (27) is standard. One may see [11, Proposition 2.2] for general estimations. Since pip_{i} and pi+1p_{i+1} are both contained in the embedding ball B(pi;ε2)¯\overline{B(p_{i};\frac{\varepsilon}{2})}, by (27) we have

(28) |φ(pi)φ(pi+1)|23εc(ε)φL2(B(pi;ε)).|\varphi(p_{i})-\varphi(p_{i+1})|\leq\frac{2}{3}\varepsilon c(\varepsilon)||\nabla\varphi||_{L^{2}(B(p_{i};\varepsilon))}.

Recall that Mi0^={pXg;dist(p,Mi0)ε}\widehat{M_{i_{0}}}=\{p\in X_{g};\ \mathop{\rm dist}(p,M_{i_{0}})\leq\varepsilon\} and so

B(pi;ε)Mi0^.B(p_{i};\varepsilon)\subset\widehat{M_{i_{0}}}.

Suppose

B(pi;ε)B(pj;ε)B(p_{i};\varepsilon)\cap B(p_{j};\varepsilon)\neq\emptyset

for some iji\neq j. By the triangle inequality we have

dist(pi,pj)<2ε.\mathop{\rm dist}(p_{i},p_{j})<2\varepsilon.

Then it follows by Lemma 4 that there exists a path γMi0\gamma^{\prime}\subset M_{i_{0}} connecting pip_{i} and pjp_{j} of length

(γ)10ε.\ell(\gamma^{\prime})\leq 10\varepsilon.

Since γ\gamma is a shortest path and ti+1ti=12εt_{i+1}-t_{i}=\frac{1}{2}\varepsilon, we have

|ij|21.|i-j|\leq 21.

Where since the last two values satisfies that tktk1ε2t_{k}-t_{k-1}\leq\frac{\varepsilon}{2}, we use 2121 instead of 2020 in the inequality above. So we have

(29) B(pi;ε)B(pi+r;ε)=,for allr22.\displaystyle B(p_{i};\varepsilon)\cap B(p_{i+r};\varepsilon)=\emptyset,\ \ \emph{for all}\ r\geq 22.

Which implies that each point in Mi0^\widehat{M_{i_{0}}} can be only contained in at most 2121 embedding balls in {B(pi;ε)}1ik\{B(p_{i};\varepsilon)\}_{1\leq i\leq k}. By Lemma 4 we know that the injectivity radius satisfies that inj(x)ε\mathop{\rm inj}(x)\geq\varepsilon for all xMi0x\in M_{i_{0}}. Thus, we have that number kk satisfies

(30) k21Vol(B(ε))Vol(Mi0^)\displaystyle k\leq\frac{21}{\mathop{\rm Vol}(B(\varepsilon))}\mathop{\rm Vol}(\widehat{M_{i_{0}}})

where Vol(B(ε))\mathop{\rm Vol}(B(\varepsilon)), only depending on ε\varepsilon, is the hyperbolic volume of a geodesic ball B(ε)B(\varepsilon)\subset\mathbb{H} with radius ε\varepsilon. Then it follow by (28), (29), (30) and the Cauchy-Schwarz inequality that

(31) Osc(i0)\displaystyle\mathop{\rm Osc}(i_{0}) =|φ(p)φ(q)|\displaystyle=|\varphi(p)-\varphi(q)|
i=1k1|φ(pi)φ(pi+1)|\displaystyle\leq\sum_{i=1}^{k-1}|\varphi(p_{i})-\varphi(p_{i+1})|
23εc(ε)i=1k1φL2(B(pi;ε))\displaystyle\leq\frac{2}{3}\varepsilon c(\varepsilon)\sum_{i=1}^{k-1}||\nabla\varphi||_{L^{2}(B(p_{i};\varepsilon))}
=23εc(ε)i=1k1B(pi;ε)|φ|2\displaystyle=\frac{2}{3}\varepsilon c(\varepsilon)\sum_{i=1}^{k-1}\sqrt{\int_{B(p_{i};\varepsilon)}|\nabla\varphi|^{2}}
23εc(ε)k1i=1k1B(pi;ε)|φ|2\displaystyle\leq\frac{2}{3}\varepsilon c(\varepsilon)\sqrt{k-1}\sqrt{\sum_{i=1}^{k-1}\int_{B(p_{i};\varepsilon)}|\nabla\varphi|^{2}}
14εc(ε)Vol(B(ε))Vol(Mi0^)Mi0^|φ|2.\displaystyle\leq\frac{14\varepsilon c(\varepsilon)}{\sqrt{\mathop{\rm Vol}(B(\varepsilon))}}\sqrt{\mathop{\rm Vol}(\widehat{M_{i_{0}}})}\sqrt{\int_{\widehat{M_{i_{0}}}}|\nabla\varphi|^{2}}.

Since i0[1,m]i_{0}\in[1,m] is arbitrary, by taking a summation we get

(32) i=1mOsc(i)\displaystyle\sum_{i=1}^{m}\mathop{\rm Osc}(i) i=1m14εc(ε)Vol(B(ε))Vol(Mi^)Mi^|φ|2\displaystyle\leq\sum_{i=1}^{m}\frac{14\varepsilon c(\varepsilon)}{\sqrt{\mathop{\rm Vol}(B(\varepsilon))}}\sqrt{\mathop{\rm Vol}(\widehat{M_{i}})}\sqrt{\int_{\widehat{M_{i}}}|\nabla\varphi|^{2}}
14εc(ε)Vol(B(ε))i=1mVol(Mi^)i=1mMi^|φ|2\displaystyle\leq\frac{14\varepsilon c(\varepsilon)}{\sqrt{\mathop{\rm Vol}(B(\varepsilon))}}\sqrt{\sum_{i=1}^{m}\mathop{\rm Vol}(\widehat{M_{i}})}\sqrt{\sum_{i=1}^{m}\int_{\widehat{M_{i}}}|\nabla\varphi|^{2}}
14εc(ε)Vol(B(ε))Vol(Xg)Xg|φ|2\displaystyle\leq\frac{14\varepsilon c(\varepsilon)}{\sqrt{\mathop{\rm Vol}(B(\varepsilon))}}\sqrt{\mathop{\rm Vol}(X_{g})}\sqrt{\int_{X_{g}}|\nabla\varphi|^{2}}
=14εc(ε)Vol(B(ε))Vol(Xg)λ1(Xg).\displaystyle=\frac{14\varepsilon c(\varepsilon)}{\sqrt{\mathop{\rm Vol}(B(\varepsilon))}}\sqrt{\mathop{\rm Vol}(X_{g})\lambda_{1}(X_{g})}.

Then the conclusion follows by setting c(ε)=14εc(ε)Vol(B(ε))c(\varepsilon)=\frac{14\varepsilon c(\varepsilon)}{\sqrt{\mathop{\rm Vol}(B(\varepsilon))}}. ∎

Proposition 12 tells that the total oscillations of φ\varphi over components of 𝒜\mathcal{A} is small. However, the following two results will tell that the oscillation of φ\varphi on 𝒜\mathcal{A} is big if λ1(Xg)\lambda_{1}(X_{g}) is small enough. This roughly tells that the total oscillations of φ\varphi over components of Xg𝒜X_{g}\setminus\mathcal{A} is big in some sense.

Lemma 13.

Let XggX_{g}\in\mathcal{M}_{g} be a hyperbolic surface with λ1(Xg)14\lambda_{1}(X_{g})\leq\frac{1}{4}, and φ\varphi be the eigenfunction with Δφ+λ1(Xg)φ=0\Delta\varphi+\lambda_{1}(X_{g})\cdot\varphi=0. Then there exist two points p1p2𝒜p_{1}\neq p_{2}\in\mathcal{A}, where 𝒜\mathcal{A} is defined in (12), such that

φ(p1)φ(p2)0.\varphi(p_{1})\cdot\varphi(p_{2})\leq 0.
Proof.

Suppose for contradiction that φ<0\varphi<0 or φ>0\varphi>0 (globally) on 𝒜\mathcal{A}. Without loss of generality we assume that φ>0\varphi>0 on 𝒜\mathcal{A}; otherwise one may replace φ-\varphi by φ\varphi. Since Xgφ=0\int_{X_{g}}\varphi=0, there exists a collar TXgT\subset X_{g} such that minpTφ(p)<0\min_{p\in T}\varphi(p)<0. By assumption we know that φ>0\varphi>0 on the boundary of TT. Thus, the nodal set {φ=0}\{\varphi=0\} bounds at least one non-empty subset TT^{\prime} of TXgT\subset X_{g}. By the analyticity of eigenfunction, one may assume that the boundary T\partial T^{\prime} of TT^{\prime} is smooth (e.g. see [8]). Then by the Stokes’ Theorem we have

T|φ|2=λ1(Xg)Tφ2.\int_{T^{\prime}}|\nabla\varphi|^{2}=\lambda_{1}(X_{g})\cdot\int_{T^{\prime}}\varphi^{2}.

Set

φ~:={φonT,0onTT.\widetilde{\varphi}:=\begin{cases}\varphi&\text{on}\ T^{\prime},\\ 0&\text{on}\ T\setminus T^{\prime}.\end{cases}

As φ~\widetilde{\varphi} vanishes on T\partial T, it follows by Lemma 7 that

14<λ1(T)T|φ~|2T|φ~|2=λ1(Xg).\frac{1}{4}<\lambda_{1}(T)\leq\frac{\int_{T}|\nabla\widetilde{\varphi}|^{2}}{\int_{T}|\widetilde{\varphi}|^{2}}=\lambda_{1}(X_{g}).

Which is a contradiction. ∎

Remark.

Lemma 13 implies that

maxx𝒜φ(x)minx𝒜φ(x)supx𝒜|φ(x)|.\max_{x\in\mathcal{A}}\varphi(x)-\min_{x\in\mathcal{A}}\varphi(x)\geq\sup_{x\in\mathcal{A}}|\varphi(x)|.

We next show that if λ1(Xg)\lambda_{1}(X_{g}) is small enough, then the magnitude of the eigenfunction on 𝒜\mathcal{A} has a uniform positive lower bound in term of the genus. More precisely,

Proposition 14.

Let XggX_{g}\in\mathcal{M}_{g} be a hyperbolic surface with

λ1(Xg)110001Vol(Xg).\lambda_{1}(X_{g})\leq\frac{1}{1000}\frac{1}{\mathop{\rm Vol}(X_{g})}.

And let φ\varphi be eigenfunction with respect to λ1(Xg)\lambda_{1}(X_{g}) with Xgφ2=1\int_{X_{g}}\varphi^{2}=1. Then

supx𝒜|φ(x)|132Vol(Xg).\sup_{x\in\mathcal{A}}|\varphi(x)|\geq\frac{1}{32\sqrt{\mathop{\rm Vol}(X_{g})}}.
Proof.

Suppose for contradiction that

supx𝒜|φ(x)|<s1Vol(Xg)\sup_{x\in\mathcal{A}}|\varphi(x)|<s\frac{1}{\sqrt{\mathop{\rm Vol}(X_{g})}}

where

s=132.s=\frac{1}{32}.

Then we have

𝒜φ2s2Vol(𝒜)Vol(Xg).\int_{\mathcal{A}}\varphi^{2}\leq s^{2}\frac{\mathop{\rm Vol}(\mathcal{A})}{\mathop{\rm Vol}(X_{g})}.

Since Xgφ2=1\int_{X_{g}}\varphi^{2}=1, XgX_{g} has non-empty modified thin part \mathcal{B} defined in (11). Thus one may assume that \mathcal{B} consists of disjoint collars T1,,TkT_{1},...,T_{k} where k>0k>0\in\mathbb{Z}. Then we have

i=1kTiφ2\displaystyle\sum_{i=1}^{k}\int_{T_{i}}\varphi^{2} =\displaystyle= 1𝒜φ2\displaystyle 1-\int_{\mathcal{A}}\varphi^{2}
\displaystyle\geq 1s2\displaystyle 1-s^{2}
\displaystyle\geq (1s2)i=1kVol(Ti)Vol(Xg).\displaystyle(1-s^{2})\sum_{i=1}^{k}\frac{\mathop{\rm Vol}(T_{i})}{\mathop{\rm Vol}(X_{g})}.

Thus for certain component TT of \mathcal{B}, we have

Tφ2(1s2)Vol(T)1Vol(Xg).\int_{T}\varphi^{2}\geq(1-s^{2})\mathop{\rm Vol}(T)\frac{1}{\mathop{\rm Vol}(X_{g})}.

Recall that the shell SS of collar TT is defined as

S={pXg; 0<dist(p,T)1}𝒜S=\{p\in X_{g};\ 0<\mathop{\rm dist}(p,T)\leq 1\}\subset\mathcal{A}

and Lemma 4 tells that

12Vol(T)4and12Vol(S)4.\frac{1}{2}\leq\mathop{\rm Vol}(T)\leq 4\quad\emph{and}\quad\frac{1}{2}\leq\mathop{\rm Vol}(S)\leq 4.

So we have

Tφ2\displaystyle\int_{T}\varphi^{2} \displaystyle\geq 1s221Vol(Xg),\displaystyle\frac{1-s^{2}}{2}\frac{1}{\mathop{\rm Vol}(X_{g})},
Sφ2\displaystyle\int_{S}\varphi^{2} \displaystyle\leq s2Vol(S)Vol(Xg)4s21Vol(Xg),\displaystyle s^{2}\frac{\mathop{\rm Vol}(S)}{\mathop{\rm Vol}(X_{g})}\leq 4s^{2}\frac{1}{\mathop{\rm Vol}(X_{g})},
S|φ|2\displaystyle\int_{S}|\nabla\varphi|^{2} \displaystyle\leq Xg|φ|2=λ1(Xg)110001Vol(Xg).\displaystyle\int_{X_{g}}|\nabla\varphi|^{2}=\lambda_{1}(X_{g})\leq\frac{1}{1000}\frac{1}{\mathop{\rm Vol}(X_{g})}.

Note that s2=11024s^{2}=\frac{1}{1024}. We apply Lemma 9 for the case that c=1s221Vol(Xg)c=\frac{1-s^{2}}{2}\frac{1}{\mathop{\rm Vol}(X_{g})} and δ=164\delta=\frac{1}{64} to get

T|φ|2\displaystyle\int_{T}|\nabla\varphi|^{2} \displaystyle\geq 31611102421Vol(Xg)\displaystyle\frac{3}{16}\frac{1-\frac{1}{1024}}{2}\frac{1}{\mathop{\rm Vol}(X_{g})}
\displaystyle\geq 1161Vol(Xg).\displaystyle\frac{1}{16}\frac{1}{\mathop{\rm Vol}(X_{g})}.

Then we have

λ1(Xg)=Xg|φ|21161Vol(Xg)\lambda_{1}(X_{g})=\int_{X_{g}}|\nabla\varphi|^{2}\geq\frac{1}{16}\frac{1}{\mathop{\rm Vol}(X_{g})}

which contradicts our assumption

λ1(Xg)110001Vol(Xg).\lambda_{1}(X_{g})\leq\frac{1}{1000}\frac{1}{\mathop{\rm Vol}(X_{g})}.

The proof is complete. ∎

Recall (25) says that

λ1(Xg)min{14,1(Xg)24Vol(Xg)2}\lambda_{1}(X_{g})\geq\min\{\frac{1}{4},\frac{\mathcal{L}_{1}(X_{g})^{2}}{4\mathop{\rm Vol}(X_{g})^{2}}\}

In order to prove Theorem 1, i.e.,

λ1(Xg)K1(Xg)Vol(Xg)2\lambda_{1}(X_{g})\geq K\frac{\mathcal{L}_{1}(X_{g})}{\mathop{\rm Vol}(X_{g})^{2}}

where K>0K>0 is a uniform constant independent of gg. In light of (25) it suffices to consider the case that 1(Xg)\mathcal{L}_{1}(X_{g}) is small. Now we assume that

(33) 1(Xg)ε\displaystyle\mathcal{L}_{1}(X_{g})\leq\varepsilon

and

(34) λ1(Xg)11000ε1(Xg)Vol(Xg)2.\lambda_{1}(X_{g})\leq\frac{1}{1000\varepsilon}\frac{\mathcal{L}_{1}(X_{g})}{\mathop{\rm Vol}(X_{g})^{2}}.

Since Vol(Xg)=4π(g1)\mathop{\rm Vol}(X_{g})=4\pi(g-1), by (34) we know that

λ1(Xg)14.\lambda_{1}(X_{g})\leq\frac{1}{4}.

So first it follows by Proposition 12 that

i=1mOsc(i)\displaystyle\sum_{i=1}^{m}\mathop{\rm Osc}(i) \displaystyle\leq c(ε)Vol(Xg)λ1(Xg)\displaystyle c(\varepsilon)\sqrt{\mathop{\rm Vol}(X_{g})\lambda_{1}(X_{g})}
\displaystyle\leq c(ε)1(Xg)1000ε1Vol(Xg).\displaystyle c(\varepsilon)\sqrt{\frac{\mathcal{L}_{1}(X_{g})}{1000\varepsilon}}\frac{1}{\sqrt{\mathop{\rm Vol}(X_{g})}}.

Since 1(Xg)ε\mathcal{L}_{1}(X_{g})\leq\varepsilon and Vol(Xg)=4π(g1)>1\mathop{\rm Vol}(X_{g})=4\pi(g-1)>1, by (34) we know that

λ1(Xg)min{14,110001Vol(Xg)}.\lambda_{1}(X_{g})\leq\min\{\frac{1}{4},\frac{1}{1000}\frac{1}{\mathop{\rm Vol}(X_{g})}\}.

Then it follows by the remark following Lemma 13 and Proposition 14 that

(36) maxx𝒜φ(x)minx𝒜φ(x)132Vol(Xg).\max_{x\in\mathcal{A}}\varphi(x)-\min_{x\in\mathcal{A}}\varphi(x)\geq\frac{1}{32\sqrt{\mathop{\rm Vol}(X_{g})}}.

Now we have the following result.

Proposition 15.

Let ε>0\varepsilon>0 in Lemma 4 and XgX_{g} be a hyperbolic surface of genus gg with

c(ε)1(Xg)1000ε<164andλ1(Xg)11000ε1(Xg)Vol(Xg)2and1(Xg)ε.c(\varepsilon)\sqrt{\frac{\mathcal{L}_{1}(X_{g})}{1000\varepsilon}}<\frac{1}{64}\quad\emph{and}\quad\lambda_{1}(X_{g})\leq\frac{1}{1000\varepsilon}\frac{\mathcal{L}_{1}(X_{g})}{\mathop{\rm Vol}(X_{g})^{2}}\quad\emph{and}\quad\mathcal{L}_{1}(X_{g})\leq\varepsilon.

Then we have

|maxx𝒜φ(x)minx𝒜φ(x)|i=1mOsc(i)164Vol(Xg).|\max_{x\in\mathcal{A}}\varphi(x)-\min_{x\in\mathcal{A}}\varphi(x)|-\sum_{i=1}^{m}\mathop{\rm Osc}(i)\geq\frac{1}{64\sqrt{\mathop{\rm Vol}(X_{g})}}.
Proof.

It clearly follows by (3) and (36). ∎

Proposition 15 in particular implies that the complement Xg𝒜X_{g}-\mathcal{A} of 𝒜\mathcal{A} is non-empty if the assumptions in the proposition hold.

4. Proof of Theorem 1

In this section we prove Theorem 1. We first make some necessary preparations.

Recall that {Mi}1im\{M_{i}\}_{1\leq i\leq m} are the components of 𝒜\mathcal{A}. For each i[1,m]i\in[1,m] we let ai,bia_{i},b_{i}\in\mathbb{R} such that

ai=minxMiφ(x)andbi=maxxMiφ(x).a_{i}=\min_{x\in M_{i}}\varphi(x)\quad\emph{and}\quad b_{i}=\max_{x\in M_{i}}\varphi(x).

That is, Im(φ|Mi)=[ai,bi]Im(\varphi|_{M_{i}})=[a_{i},b_{i}]. So for each i[1,m]i\in[1,m],

Osc(i)=biai.\mathop{\rm Osc}(i)=b_{i}-a_{i}.

For each component of Xg𝒜X_{g}\setminus\mathcal{A}, it is a collar whose two boundary curves are contained in two components denoted by MiM_{i} and MjM_{j} of 𝒜\mathcal{A} (MiM_{i} may be the same as MjM_{j}). There may exist multiple collars bounded by MiM_{i} and MjM_{j}. We denote these collars by Tij1,,TijθijT_{ij}^{1},...,T_{ij}^{\theta_{ij}}. Set

δijθ=min(x,x)Γ1×Γ2|φ(x)φ(x)|\delta_{ij}^{\theta}=\min_{(x,x^{*})\in\Gamma_{1}\times\Gamma_{2}}|\varphi(x)-\varphi(x^{*})|

where Γ1,Γ2\Gamma_{1},\Gamma_{2} are the two boundaries of TijθT_{ij}^{\theta} and xΓ2x^{*}\in\Gamma_{2} is the reflection of xΓ1x\in\Gamma_{1} through the center closed geodesic γijθ\gamma_{ij}^{\theta}. By Lemma 8 we have

Tijθ|φ|2rijθ4(δijθ)2\int_{T_{ij}^{\theta}}|\nabla\varphi|^{2}\geq\frac{r_{ij}^{\theta}}{4}(\delta_{ij}^{\theta})^{2}

where rijθr_{ij}^{\theta} is the length of center closed geodesic of collar TijθT_{ij}^{\theta}. Set

(37) δij={0if[ai,bi][aj,bj],dist([ai,bi],[aj,bj])if[ai,bi][aj,bj]=.\delta_{ij}=\begin{cases}0&\text{if}\ [a_{i},b_{i}]\cap[a_{j},b_{j}]\neq\emptyset,\\ \mathop{\rm dist}([a_{i},b_{i}],[a_{j},b_{j}])&\text{if}\ [a_{i},b_{i}]\cap[a_{j},b_{j}]=\emptyset.\end{cases}

Then

(38) δijθδij.\displaystyle\delta_{ij}^{\theta}\geq\delta_{ij}.

Let {Tijθ}\{T_{ij}^{\theta}\} be all the components of Xg𝒜X_{g}\setminus\mathcal{A}. Putting the inequalities above together we get

(39) λ1(Xg)\displaystyle\lambda_{1}(X_{g}) =Xg|φ|2\displaystyle=\int_{X_{g}}|\nabla\varphi|^{2}
TijθTijθ|φ|2\displaystyle\geq\sum_{T_{ij}^{\theta}}\int_{T_{ij}^{\theta}}|\nabla\varphi|^{2}
Tijθrijθ4(δij)2\displaystyle\geq\sum_{T_{ij}^{\theta}}\frac{r_{ij}^{\theta}}{4}(\delta_{ij})^{2}
141(components ofXg𝒜)(Tijθrijθδij)2\displaystyle\geq\frac{1}{4}\frac{1}{\sharp(\emph{components of}\ X_{g}\setminus\mathcal{A})}\left(\sum_{T_{ij}^{\theta}}\sqrt{r_{ij}^{\theta}}\delta_{ij}\right)^{2}
112(g1)(Tijθrijθδij)2.\displaystyle\geq\frac{1}{12(g-1)}\left(\sum_{T_{ij}^{\theta}}\sqrt{r_{ij}^{\theta}}\delta_{ij}\right)^{2}.

Next we will bound the quantity Tijθrijθδij\sum_{T_{ij}^{\theta}}\sqrt{r_{ij}^{\theta}}\delta_{ij} from below in terms of 1(Xg)\mathcal{L}_{1}(X_{g}) and Vol(Xg)\mathop{\rm Vol}(X_{g}). As a function, the summation Tijθrijθδij\sum_{T_{ij}^{\theta}}\sqrt{r_{ij}^{\theta}}\delta_{ij} is linear with respect to the intervals [ai,bi][a_{i},b_{i}]. So it achieves its minimum at some extremal case, which may be related to 1(Xg)\mathcal{L}_{1}(X_{g}). We first prove the following elementary property.

Lemma 16.

Let n>0n>0\in\mathbb{Z} and assume {Ii=[ai,bi]}1in\{I_{i}=[a_{i},b_{i}]\}_{1\leq i\leq n} are nn closed intervals with increasing order

(40) a1b1a2b2a3<anbn.a_{1}\leq b_{1}\leq a_{2}\leq b_{2}\leq a_{3}...<a_{n}\leq b_{n}.

Let δ=i=1n|biai|\delta=\sum_{i=1}^{n}|b_{i}-a_{i}| be the total lengths of all the intervals. Then for any collection of nonnegative numbers {αij}1i<jn\{\alpha_{ij}\}_{1\leq i<j\leq n}, there exists an integer K0K_{0} with 1K0<n1\leq K_{0}<n such that

(41) i<jαijdist(Ii,Ij)(bna1i=1n|biai|)1iK0<jnαij.\sum_{i<j}\alpha_{ij}\mathop{\rm dist}(I_{i},I_{j})\geq(b_{n}-a_{1}-\sum_{i=1}^{n}|b_{i}-a_{i}|)\sum_{1\leq i\leq K_{0}<j\leq n}\alpha_{ij}.
Proof.

We prove it by induction on nn.

If n=1n=1, both sides of (41) are equal to 0.

If n=2n=2, (41) holds by letting K0=1K_{0}=1.

Assume that (41) holds for k(n1)k\leq(n-1).

Now for k=nk=n, we first regard

Ln(a1,b1,,an,bn)=i<jαijdist(Ii,Ij)=i<jαij(ajbi)L_{n}(a_{1},b_{1},...,a_{n},b_{n})=\sum_{i<j}\alpha_{ij}\mathop{\rm dist}(I_{i},I_{j})=\sum_{i<j}\alpha_{ij}(a_{j}-b_{i})

as a function (a1,b1,,an,bn)(a_{1},b_{1},...,a_{n},b_{n}). Set

L(t)=Ln(a1,b1,a2+t,b2+t,a3,b3,,an,bn).L(t)=L_{n}(a_{1},b_{1},a_{2}+t,b_{2}+t,a_{3},b_{3},...,a_{n},b_{n}).

Clearly L(t)L(t) is linear with respect to tt on its domain which is the one preserving the relation (40). So L(t)L(t) takes its minimum on the boundaries, i.e., when

a2+t=b1orb2+t=a3.a_{2}+t=b_{1}\ \ \text{or}\ \ b_{2}+t=a_{3}.

So we have

(42) Ln(a1,b1;)min{Ln(a1,b1,b1,(b1+b2a2),a3,b3,,an,bn),Ln(a1,b1,(a3+a2b2),a3,a3,b3,,an,bn)}.L_{n}(a_{1},b_{1};...)\geq\min\left\{\begin{array}[]{c}L_{n}(a_{1},b_{1},b_{1},(b_{1}+b_{2}-a_{2}),a_{3},b_{3},...,a_{n},b_{n}),\\ L_{n}(a_{1},b_{1},(a_{3}+a_{2}-b_{2}),a_{3},a_{3},b_{3},...,a_{n},b_{n})\\ \end{array}\right\}.

Without loss of generality one may assume that Ln(a1,b1,b1,(b1+b2a2),,an,bn)L_{n}(a_{1},b_{1},b_{1},(b_{1}+b_{2}-a_{2}),...,a_{n},b_{n}) is the smaller one. Then we have

Ln(a1,b1,b1,(b1+b2a2),,an,bn)=i3α1i(aib1)\displaystyle L_{n}(a_{1},b_{1},b_{1},(b_{1}+b_{2}-a_{2}),...,a_{n},b_{n})=\sum_{i\geq 3}\alpha_{1i}(a_{i}-b_{1})
+i3α2i(ai(b1+b2a2))+3i<jnαij(ajbi)\displaystyle+\sum_{i\geq 3}\alpha_{2i}(a_{i}-(b_{1}+b_{2}-a_{2}))+\sum_{3\leq i<j\leq n}\alpha_{ij}(a_{j}-b_{i})
\displaystyle\geq i3(α1i+α2i)(ai(b1+b2a2))+3i<jnαij(ajbi)\displaystyle\sum_{i\geq 3}(\alpha_{1i}+\alpha_{2i})(a_{i}-(b_{1}+b_{2}-a_{2}))+\sum_{3\leq i<j\leq n}\alpha_{ij}(a_{j}-b_{i})

Set

I1=[a1,b1+b2a2]andIi=[ai+1,bi+1](2i(n1))I_{1}^{\prime}=[a_{1},b_{1}+b_{2}-a_{2}]\quad\emph{and}\quad\ I_{i}^{\prime}=[a_{i+1},b_{i+1}]\ (2\leq i\leq(n-1))

and

α1i=α1(i+1)+α2(i+1)andαij=α(i+1)(j+1)(2i<j(n1)).\alpha_{1i}^{\prime}=\alpha_{1(i+1)}+\alpha_{2(i+1)}\quad\emph{and}\quad\alpha_{ij}^{\prime}=\alpha_{(i+1)(j+1)}\ (2\leq i<j\leq(n-1)).

Then we have

i3(α1i+α2i)(aib1b2+a2)+3i<jnαij(ajbi)=1i<jn1αijdist(Ii,Ij).\displaystyle\sum_{i\geq 3}(\alpha_{1i}+\alpha_{2i})(a_{i}-b_{1}-b_{2}+a_{2})+\sum_{3\leq i<j\leq n}\alpha_{ij}(a_{j}-b_{i})=\sum_{1\leq i<j\leq n-1}\alpha^{\prime}_{ij}\mathop{\rm dist}(I_{i}^{\prime},I_{j}^{\prime}).

By our induction assumption on k=(n1)k=(n-1), we know that there exists an integer K1[1,n1]K_{1}\in[1,n-1] such that

1i<j(n1)αijdist(Ii,Ij)\displaystyle\sum_{1\leq i<j\leq(n-1)}\alpha^{\prime}_{ij}\mathop{\rm dist}(I_{i}^{\prime},I_{j}^{\prime}) \displaystyle\geq (bna1i=1n|biai|)1iK1<j(n1)αij\displaystyle(b_{n}-a_{1}-\sum_{i=1}^{n}|b_{i}-a_{i}|)\sum_{1\leq i\leq K_{1}<j\leq(n-1)}\alpha^{\prime}_{ij}
=\displaystyle= (bna1i=1n|biai|)1i(K1+1)<jnαij.\displaystyle(b_{n}-a_{1}-\sum_{i=1}^{n}|b_{i}-a_{i}|)\sum_{1\leq i\leq(K_{1}+1)<j\leq n}\alpha_{ij}.

Set K0=K1+1K_{0}=K_{1}+1. Then the conclusion follows by (42), (4) and (4). ∎

Now we return to estimate the quantity Tijθrijθδij\sum_{T_{ij}^{\theta}}\sqrt{r_{ij}^{\theta}}\delta_{ij} in (39).

Proposition 17.
Tijθrijθδij\displaystyle\sum_{T_{ij}^{\theta}}\sqrt{r_{ij}^{\theta}}\delta_{ij} 1(Xg)(maxi,j|biaj|i=1m|biai|)\displaystyle\geq\sqrt{\mathcal{L}_{1}(X_{g})}\left(\max_{i,j}|b_{i}-a_{j}|-\sum_{i=1}^{m}|b_{i}-a_{i}|\right)
=1(Xg)(|maxx𝒜φ(x)minx𝒜φ(x)|i=1mOsc(i))\displaystyle=\sqrt{\mathcal{L}_{1}(X_{g})}\left(|\max_{x\in\mathcal{A}}\varphi(x)-\min_{x\in\mathcal{A}}\varphi(x)|-\sum_{i=1}^{m}\mathop{\rm Osc}(i)\right)

where [ai,bi]=Im(φ|Mi)[a_{i},b_{i}]=Im(\varphi|_{M_{i}}).

Proof.

Assume

i=1m[ai,bi]=k=1N[ek,fk]\bigcup_{i=1}^{m}[a_{i},b_{i}]=\bigcup_{k=1}^{N}[e_{k},f_{k}]

where the [ek,fk][e_{k},f_{k}] are disjoint and

minx𝒜φ(x)=e1f1<e2f2<<ekfk=maxx𝒜φ(x).\min_{x\in\mathcal{A}}\varphi(x)=e_{1}\leq f_{1}<e_{2}\leq f_{2}<...<e_{k}\leq f_{k}=\max_{x\in\mathcal{A}}\varphi(x).

Since the [ek,fk][e_{k},f_{k}] are disjoint, for each i[1,m]i\in[1,m] there exists a unique ki[1,N]k_{i}\in[1,N] such that

Im(φ|Mi)=[ai,bi][eki,fki].Im(\varphi|_{M_{i}})=[a_{i},b_{i}]\subset[e_{k_{i}},f_{k_{i}}].

For iji\neq j, we set

δij:=dist([eki,fki],[ekj,fkj]).\delta^{\prime}_{ij}:=\mathop{\rm dist}([e_{k_{i}},f_{k_{i}}],[e_{k_{j}},f_{k_{j}}]).

Let δij\delta_{ij} be the constants given in (37). Clearly we have

(45) δijδijfor all ij.\delta_{ij}\geq\delta^{\prime}_{ij}\quad\emph{for all $i\neq j$}.

Which implies that

(46) TijθrijθδijTijθrijθδij.\displaystyle\sum_{T_{ij}^{\theta}}\sqrt{r_{ij}^{\theta}}\delta_{ij}\geq\sum_{T_{ij}^{\theta}}\sqrt{r_{ij}^{\theta}}\delta^{\prime}_{ij}.

We rewrite the quantity Tijθrijθδij\sum_{T_{ij}^{\theta}}\sqrt{r_{ij}^{\theta}}\delta^{\prime}_{ij} as

(47) Tijθrijθδij=p<q(ki=p,kj=qrijθ)dist([ep,fp],[eq,fq]).\displaystyle\sum_{T_{ij}^{\theta}}\sqrt{r_{ij}^{\theta}}\delta^{\prime}_{ij}=\sum_{p<q}\left(\sum_{\ k_{i}=p,\ k_{j}=q}\sqrt{r_{ij}^{\theta}}\right)\mathop{\rm dist}([e_{p},f_{p}],[e_{q},f_{q}]).

Then it follows by Lemma 16 that there exists an integer K0[1,N1]K_{0}\in[1,N-1] such that

(48) p<q(ki=p,kj=qrijθ)dist([ep,fp],[eq,fq])\displaystyle\sum_{p<q}\left(\sum_{\ k_{i}=p,\ k_{j}=q}\sqrt{r_{ij}^{\theta}}\right)\mathop{\rm dist}([e_{p},f_{p}],[e_{q},f_{q}])
(fNe1k=1N|fkek|)1kiK0<kjNrijθ\displaystyle\geq(f_{N}-e_{1}-\sum_{k=1}^{N}|f_{k}-e_{k}|)\sum_{1\leq k_{i}\leq K_{0}<k_{j}\leq N}\sqrt{r_{ij}^{\theta}}
(1kiK0<kjNrijθ)(maxi,j|biaj|i=1m|biai|).\displaystyle\geq\left(\sum_{1\leq k_{i}\leq K_{0}<k_{j}\leq N}\sqrt{r_{ij}^{\theta}}\right)\left(\max_{i,j}|b_{i}-a_{j}|-\sum_{i=1}^{m}|b_{i}-a_{i}|\right).

Note that the unions of all the center closed geodesics of collars TijθT_{ij}^{\theta}, which satisfy 1kiK0<kjN1\leq k_{i}\leq K_{0}<k_{j}\leq N, divide XgX_{g} into at least two components: one contains a MiM_{i} with kiK0k_{i}\leq K_{0}; another one contains a MjM_{j} with kj>K0k_{j}>K_{0}. So by the definition of 1(Xg)\mathcal{L}_{1}(X_{g}) we have

(49) 1kiK0<kjNrijθ1(Xg)\sum_{1\leq k_{i}\leq K_{0}<k_{j}\leq N}r_{ij}^{\theta}\geq\mathcal{L}_{1}(X_{g})

implying that

(50) 1kiK0<kjNrijθ1kiK0<kjNrijθ1(Xg).\displaystyle\sum_{1\leq k_{i}\leq K_{0}<k_{j}\leq N}\sqrt{r_{ij}^{\theta}}\geq\sqrt{\sum_{1\leq k_{i}\leq K_{0}<k_{j}\leq N}r_{ij}^{\theta}}\geq\sqrt{\mathcal{L}_{1}(X_{g})}.

Then it follows by (46), (47), (48) and (50) that

(51) Tijθrijθδij1(Xg)(maxi,j|biaj|i=1m|biai|)\displaystyle\sum_{T_{ij}^{\theta}}\sqrt{r_{ij}^{\theta}}\delta_{ij}\geq\sqrt{\mathcal{L}_{1}(X_{g})}\left(\max_{i,j}|b_{i}-a_{j}|-\sum_{i=1}^{m}|b_{i}-a_{i}|\right)

which completes the proof. ∎

Now we are ready to prove Theorem 1.

Proof of Theorem 1.

Recall that (25) says that

λ1(Xg)min{14,1(Xg)24Vol(Xg)2}.\lambda_{1}(X_{g})\geq\min\{\frac{1}{4},\frac{\mathcal{L}_{1}(X_{g})^{2}}{4\mathop{\rm Vol}(X_{g})^{2}}\}.

Let ε>0\varepsilon>0 be the uniform constant in Lemma 4. First we assume that

(52) c(ε)1(Xg)1000ε<164and1(Xg)ε,c(\varepsilon)\sqrt{\frac{\mathcal{L}_{1}(X_{g})}{1000\varepsilon}}<\frac{1}{64}\quad\text{and}\quad\mathcal{L}_{1}(X_{g})\leq\varepsilon,

otherwise we have

(53) 1(Xg)2Vol(Xg)21000ε642c(ε)21(Xg)Vol(Xg)2or1(Xg)2Vol(Xg)2ε1(Xg)Vol(Xg)2\displaystyle\frac{\mathcal{L}_{1}(X_{g})^{2}}{\mathop{\rm Vol}(X_{g})^{2}}\geq\frac{1000\varepsilon}{64^{2}c(\varepsilon)^{2}}\frac{\mathcal{L}_{1}(X_{g})}{\mathop{\rm Vol}(X_{g})^{2}}\ \text{or}\ \frac{\mathcal{L}_{1}(X_{g})^{2}}{\mathop{\rm Vol}(X_{g})^{2}}\geq\varepsilon\frac{\mathcal{L}_{1}(X_{g})}{\mathop{\rm Vol}(X_{g})^{2}}

which together with (25) complete the proof. Now we also assume that

(54) λ1(Xg)11000ε1(Xg)Vol(Xg)2,\displaystyle\lambda_{1}(X_{g})\leq\frac{1}{1000\varepsilon}\frac{\mathcal{L}_{1}(X_{g})}{\mathop{\rm Vol}(X_{g})^{2}},

otherwise we are done. By our assumption (52) and (54), we apply Proposition 15 to get

(55) |maxx𝒜φ(x)minx𝒜φ(x)|i=1mOsc(i)164Vol(Xg).|\max_{x\in\mathcal{A}}\varphi(x)-\min_{x\in\mathcal{A}}\varphi(x)|-\sum_{i=1}^{m}\mathop{\rm Osc}(i)\geq\frac{1}{64\sqrt{\mathop{\rm Vol}(X_{g})}}.

Thus, by combining (39) and Proposition 17 we have

(56) λ1(Xg)\displaystyle\lambda_{1}(X_{g}) 112(g1)(Tijθrijθδij)2\displaystyle\geq\frac{1}{12(g-1)}\left(\sum_{T_{ij}^{\theta}}\sqrt{r_{ij}^{\theta}}\delta_{ij}\right)^{2}\
1(Xg)12(g1)(|maxxAφ(x)minxAφ(x)|i=1mOsc(i))2.\displaystyle\geq\frac{\mathcal{L}_{1}(X_{g})}{12(g-1)}\left(|\max_{x\in A}\varphi(x)-\min_{x\in A}\varphi(x)|-\sum_{i=1}^{m}\mathop{\rm Osc}(i)\right)^{2}.
1(Xg)491521(g1)Vol(Xg).\displaystyle\geq\frac{\mathcal{L}_{1}(X_{g})}{49152}\frac{1}{(g-1)\cdot\mathop{\rm Vol}(X_{g})}.

Recall that Vol(Xg)=4π(g1)\mathop{\rm Vol}(X_{g})=4\pi(g-1). Therefore, the discussions above imply that

(59) λ1(Xg)\displaystyle\lambda_{1}(X_{g}) min{14,ε1(Xg)4Vol(Xg)2,1000ε642c(ε)21(Xg)4Vol(Xg)2,11000ε1(Xg)Vol(Xg)2,1(Xg)491521(g1)Vol(Xg)}\displaystyle\geq\min\left\{\begin{array}[]{c}\frac{1}{4},\varepsilon\cdot\frac{\mathcal{L}_{1}(X_{g})}{4\mathop{\rm Vol}(X_{g})^{2}},\frac{1000\varepsilon}{64^{2}c(\varepsilon)^{2}}\cdot\frac{\mathcal{L}_{1}(X_{g})}{4\mathop{\rm Vol}(X_{g})^{2}},\\ \frac{1}{1000\varepsilon}\frac{\mathcal{L}_{1}(X_{g})}{\mathop{\rm Vol}(X_{g})^{2}},\frac{\mathcal{L}_{1}(X_{g})}{49152}\frac{1}{(g-1)\cdot\mathop{\rm Vol}(X_{g})}\\ \end{array}\right\}
min{14,c(ε)1(Xg)g2}\displaystyle\geq\min\{\frac{1}{4},c(\varepsilon)\cdot\frac{\mathcal{L}_{1}(X_{g})}{g^{2}}\}

where c(ε)>0c(\varepsilon)>0 is a uniform constant only depending on ε\varepsilon. By Lemma 6 we know that

1(Xg)cg\mathcal{L}_{1}(X_{g})\leq c^{\prime}\cdot g

for some uniform constant c>0c^{\prime}>0. So for large enough gg, c(ε)1(Xg)g2<14c(\varepsilon)\cdot\frac{\mathcal{L}_{1}(X_{g})}{g^{2}}<\frac{1}{4}. Thus, we get

(60) λ1(Xg)K11(Xg)g2\lambda_{1}(X_{g})\geq K_{1}\cdot\frac{\mathcal{L}_{1}(X_{g})}{g^{2}}

for some uniform constant K1>0K_{1}>0 independent of gg. The proof is complete. ∎

5. An optimal example

In this section, we use the results in [17] to explain that the lower bound of λ1(Xg)\lambda_{1}(X_{g}) in Theorem 1 is optimal as gg\to\infty. That is, for all genus g2g\geq 2, there exists some Riemann surfaces 𝒳g\mathcal{X}_{g} of genus gg such that

λ1(𝒳g)K21(𝒳g)g2\lambda_{1}(\mathcal{X}_{g})\leq K_{2}\frac{\mathcal{L}_{1}(\mathcal{X}_{g})}{g^{2}}

for some uniform constant K2>0K_{2}>0, especially when 1(𝒳g)\mathcal{L}_{1}(\mathcal{X}_{g}) is arbitrarily small.

The following construction of 𝒳g\mathcal{X}_{g} is given in [17].

Recall that a pair of pants is a compact Riemann surface of 0 genus with 33 boundary closed geodesics. The complex structure is uniquely determined by the lengths of the three boundary closed geodesics.

For a fixed

<2arcsinh1,\ell<2\mathop{\rm arcsinh}1,

consider the pair of pants 𝒫\mathcal{P}_{\ell} whose boundary curves all have length equal to \ell. We construct 𝒳g\mathcal{X}_{g} by gluing 3g33g-3 pants 𝒫\mathcal{P}_{\ell} along boundary loops (with arbitrary twist parameters) as shown in figure 3.

Figure 3. Surface 𝒳g\mathcal{X}_{g}

For any closed geodesic γ𝒳g\gamma\subset\mathcal{X}_{g}, the curve γ\gamma is either one of the boundary closed geodesic of certain 𝒫\mathcal{P}_{\ell} or must intersect at least one of the boundary closed geodesic of certain 𝒫\mathcal{P}_{\ell}. Recall that the Collar Lemma 2 implies that for any two intersected closed geodesics, then at least one of them have length larger than 2arcsinh12\mathop{\rm arcsinh}1. So if γ\gamma is a systolic curve of 𝒳g\mathcal{X}_{g}, then γ\gamma must be one of the boundary curve of 𝒫\mathcal{P}_{\ell}. Thus, the systole of XgX_{g} and 1(𝒳g)\mathcal{L}_{1}(\mathcal{X}_{g}) are both equal to \ell. That is,

1(𝒳g)=sys(𝒳g)=.\mathcal{L}_{1}(\mathcal{X}_{g})=\mathop{\rm sys}(\mathcal{X}_{g})=\ell.

It was proved [17, Proposition 15 or Remark 16] that

(61) λ1(𝒳g)β()g2.\lambda_{1}(\mathcal{X}_{g})\leq\frac{\beta(\ell)}{g^{2}}.

Where for the case that <2arcsinh1\ell<2\mathop{\rm arcsinh}1 and 𝒳g\mathcal{X}_{g} constructed as above,

β()K2\beta(\ell)\leq K_{2}\cdot\ell

for some universal constant K2>0K_{2}>0. So we have

(62) λ1(𝒳g)K21(𝒳g)g2\lambda_{1}(\mathcal{X}_{g})\leq K_{2}\frac{\mathcal{L}_{1}(\mathcal{X}_{g})}{g^{2}}

which tells that the lower bound of λ1(Xg)\lambda_{1}(X_{g}) in Theorem 1 is optimal either as gg\to\infty or as 1(Xg)0\mathcal{L}_{1}(X_{g})\to 0.

References

  • [1] Werner Ballmann, Henrik Matthiesen, and Sugata Mondal. Small eigenvalues of closed surfaces. J. Differential Geom., 103(1):1–13, 2016.
  • [2] Werner Ballmann, Henrik Matthiesen, and Sugata Mondal. On the analytic systole of Riemannian surfaces of finite type. Geom. Funct. Anal., 27(5):1070–1105, 2017.
  • [3] Lipman Bers. An inequality for Riemann surfaces. In Differential geometry and complex analysis, pages 87–93. Springer, Berlin, 1985.
  • [4] Peter Buser. Riemannsche Flächen mit Eigenwerten in (0,(0, 1/4)1/4). Comment. Math. Helv., 52(1):25–34, 1977.
  • [5] Peter Buser. Geometry and spectra of compact Riemann surfaces, volume 106 of Progress in Mathematics. Birkhäuser Boston, Inc., Boston, MA, 1992.
  • [6] Isaac Chavel. Eigenvalues in Riemannian geometry, volume 115 of Pure and Applied Mathematics. Academic Press, Inc., Orlando, FL, 1984. Including a chapter by Burton Randol, With an appendix by Jozef Dodziuk.
  • [7] Jeff Cheeger. A lower bound for the smallest eigenvalue of the Laplacian. In Problems in analysis (Papers dedicated to Salomon Bochner, 1969), pages 195–199. 1970.
  • [8] Shiu Yuen Cheng. Eigenfunctions and nodal sets. Comment. Math. Helv., 51(1):43–55, 1976.
  • [9] Jozef Dodziuk, Thea Pignataro, Burton Randol, and Dennis Sullivan. Estimating small eigenvalues of Riemann surfaces. In The legacy of Sonya Kovalevskaya (Cambridge, Mass., and Amherst, Mass., 1985), volume 64 of Contemp. Math., pages 93–121.
  • [10] Jozef Dodziuk and Burton Randol. Lower bounds for λ1\lambda_{1} on a finite-volume hyperbolic manifold. J. Differential Geom., 24(1):133–139, 1986.
  • [11] Michael Lipnowski and Mark Stern. Geometry of the smallest 1-form Laplacian eigenvalue on hyperbolic manifolds. Geom. Funct. Anal., 28(6):1717–1755, 2018.
  • [12] Sugata Mondal. Systole and λ2g2\lambda_{2g-2} of closed hyperbolic surfaces of genus gg. Enseign. Math., 60(1-2):3–24, 2014.
  • [13] Xin Nie, Yunhui Wu, and Yuhao Xue. Large genus asymptotics for lengths of separating closed geodesics on random surfaces. arXiv e-prints, page arXiv:2009.07538, September 2020.
  • [14] Jean-Pierre Otal and Eulalio Rosas. Pour toute surface hyperbolique de genre g,λ2g2>1/4g,\ \lambda_{2g-2}>1/4. Duke Math. J., 150(1):101–115, 2009.
  • [15] R. Schoen, S. Wolpert, and S. T. Yau. Geometric bounds on the low eigenvalues of a compact surface. In Geometry of the Laplace operator (Proc. Sympos. Pure Math., Univ. Hawaii, Honolulu, Hawaii, 1979), Proc. Sympos. Pure Math., XXXVI, pages 279–285. Amer. Math. Soc., Providence, R.I., 1980.
  • [16] Michael E. Taylor. Partial differential equations I. Basic theory, volume 115 of Applied Mathematical Sciences. Springer, New York, second edition, 2011.
  • [17] Yunhui Wu and Yuhao Xue. Small eigenvalues of closed Riemann surfaces for large genus. arXiv e-prints, page arXiv:1809.07449, Sep 2018.