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

Average area ratio and normalized total scalar curvature of hyperbolic n-manifolds

Ruojing Jiang the University of Chicago, Department of Mathematics, Chicago, IL 60637 [email protected]
Abstract.

On closed hyperbolic manifolds of dimension n3n\geq 3, we review the definition of the average area ratio of a metric hh with Rhn(n1)R_{h}\geq-n(n-1) relative to the hyperbolic metric h0h_{0}, and we prove that it attains the local minimum of one at h0h_{0}, which solves a local version of Gromov’s conjecture. Additionally, we discuss the relation between the average area ratio and normalized total scalar curvature for hyperbolic nn-manifolds, as well as its relation to the minimal surface entropy if nn is odd.

1. Introduction

On a closed negatively curved manifold MM, a useful fact is that the unit tangent bundle admits a one-dimensional foliation whose leaves are orbits of the geodesic flow. Given a pair of Riemannian metrics hh and h0h_{0} on MM, the comparisons of the geometric objects associated with hh and h0h_{0} have been studied and become well-known. For instance, the geodesic stretch Iμ(h/h0)I_{\mu}(h/h_{0}) measures the stretching of the metric hh relative to the reference metric h0h_{0} and the measure μ\mu. Comparisons related to Iμ(h/h0)I_{\mu}(h/h_{0}) were discussed by Knieper [16]. Another example is the volume entropy EvolE_{vol} that counts the number of closed geodesics, the locally symmetric metric attains the minimum among all metrics on MM with the same volume (see Besson-Courtois-Gallot [2]). The more detailed results will be reviewed later on, for the light they shed on what could be expected and extended to dimensions larger than one.

The first analogue to dimension two was discussed by Gromov in [8]. A two-dimensional foliation of Gr2MGr_{2}M is formed by a family of stable minimal surfaces of MM. In particular, let h0h_{0} denote the hyperbolic metric, there is a canonical foliation of Gr2MGr_{2}M whose leaves are totally geodesic planes. Gromov introduced the average area ratio AreaF(h/h0)\text{Area}_{F}(h/h_{0}) which is similar in spirit to the geodesic stretch, to measure the “stretching” of the area of leaves on Gr2MGr_{2}M of the metric hh relative to the hyperbolic metric h0h_{0} under the map FF. The comparison result and the interpretation of AreaId(h/h0)\text{Area}_{\text{Id}}(h/h_{0}) for 33-manifolds were studied by Lowe-Neves [19] using Ricci flow. In this paper, we focus on the local picture of AreaId(h/h0)\text{Area}_{\text{Id}}(h/h_{0}) for any dimension n3n\geq 3, and investigate a more general case for 33-manifolds.

On the other hand, Calegari-Marques-Neves [4] defined the minimal surface entropy E(h)E(h) based on the construction of surface subgroups by Kahn-Markovic [15], and the equidistribution property of PSL(2,)\text{PSL}(2,\mathbb{R})-action on the space of minimal laminations (initiated by Ratner [20] and Shah [22], and recently formalized by Labourie [17]). E(h)E(h) measures the number of essential minimal surfaces of MM with respect to hh, thus shifting attention from a one-dimensional object (volume entropy) to dimension two. The minimal surface entropy of closed hyperbolic 33-manifolds was computed by Calegari-Marques-Neves [4] (also see Lowe-Neves [19]), and the author extended the result to manifolds of any odd dimension at least 33 in [13], we continue the discussion for higher dimensions in this paper.

1.1. Gromov’s conjecture

Let (M,h0)(M,h_{0}) be a closed hyperbolic 33-manifold, and let NN be a closed 33-manifold with Riemannian metric gg whose scalar curvature satisfies Rg6R_{g}\geq-6. F:NMF:N\rightarrow M is a smooth map with degree dd. The average area ratio of FF is as follows.

AreaF(g/h0)=(y,P)Gr2(M)xF1(y)limδ0areag((dFx)1(Dδ))δdμh0,\text{Area}_{F}(g/h_{0})=\int_{(y,P)\in Gr_{2}(M)}\sum_{x\in F^{-1}(y)}\lim_{\delta\rightarrow 0}\frac{\text{area}_{g}((dF_{x})^{-1}(D_{\delta}))}{\delta}\,d\mu_{h_{0}},

where DδD_{\delta} is a subset of the totally geodesic disc of n\mathbb{H}^{n} which is tangential to PP at xx, DδD_{\delta} has an area equal to δ\delta, and μh0\mu_{h_{0}} stands for the unit volume measure on Gr2(M)Gr_{2}(M) with respect to the metric induced by h0h_{0}. We refer to Section 2 for a more detailed definition. Gromov proved the following inequality using stability (see page 73 of [8]).

AreaF(g/h0)d3.\text{Area}_{F}(g/h_{0})\geq\frac{d}{3}.

The proof indicates that if the equality holds, then (N,g)(N,g) should be hyperbolic, and in the universal covers, the preimage of any totally geodesic disc is totally geodesic as well. However, it means the inequality is not sharp, and therefore, Gromov conjectured that the lower bound could be replaced by dd. Moreover, the higher dimensional case may share the same property.

Conjecture 1.1.

Let (M,h0)(M,h_{0}) be a closed hyperbolic manifold of dimension n3n\geq 3, and let NN be a closed nn-manifold with Riemannian metric gg, and its scalar curvature satisfies Rgn(n1)R_{g}\geq-n(n-1). F:NMF:N\rightarrow M is a smooth map with degree dd. We have

AreaF(g/h0)d.\text{Area}_{F}(g/h_{0})\geq d.

The equality holds if and only if FF is a local isometry.

Recently, Lowe-Neves (Corollary 1.3 of [19]) verified the special case where n=3n=3 and FF is a local diffeomorphism. We will generalize this conclusion in two aspects in Theorem 1.2 and Theorem 1.4.

1.2. Minimal Surface Entropy

Let n\mathbb{H}^{n} denote the hyperbolic nn-space, where n3n\geq 3. In the Poincaré ball model, the asymptotic boundary n\partial_{\infty}\mathbb{H}^{n} can be considered as the (n1)(n-1)-unit sphere Sn1S^{n-1}_{\infty}. A homeomorphism f:Sn1Sn1f:S^{n-1}_{\infty}\rightarrow S^{n-1}_{\infty} is called KK-quasiconformal if K(f)=ess supxSn1Kf(x)KK(f)=\underset{x\in S^{n-1}_{\infty}}{\text{ess sup}}K_{f}(x)\leq K, where the dilatation of ff is

Kf(x)=limsupr0supy,z|f(x)f(y)||f(x)f(z)|,K_{f}(x)=\underset{r\rightarrow 0}{\lim\sup}\,\underset{y,z}{\sup}\frac{|f(x)-f(y)|}{|f(x)-f(z)|},

the supremum is taken over y,zy,z so that r=|xy|=|xz|r=|x-y|=|x-z|. A KK-quasicircle in Sn1S^{n-1}_{\infty} is the image of a round circle under a KK-quasiconformal map.

Let M=n/π1(M)M=\mathbb{H}^{n}/\pi_{1}(M) be a closed orientable nn-manifold (n3)(n\geq 3) that admits a hyperbolic metric h0h_{0}, A closed surface immersed in MM with genus at least 22 is said to be essential if the immersion is π1\pi_{1}-injective, and the image of its fundamental group in π1(M)\pi_{1}(M) is called a surface subgroup. Let S(M,g)S(M,g) denote the set of surface subgroups of genus at most gg up to conjugacy, and let the subset S(M,g,ϵ)S(M,g)S(M,g,\epsilon)\subset S(M,g) consist of the conjugacy classes whose limit sets are (1+ϵ)(1+\epsilon)-quasicircles. Moreover,

Sϵ(M)=g2S(M,g,ϵ).S_{\epsilon}(M)=\underset{g\geq 2}{\cup}S(M,g,\epsilon).

Suppose hh is an arbitrary Riemannian metric on MM. For any ΠS(M,g)\Pi\in S(M,g), we set

areah(Π)=inf{areah(Σ):ΣΠ}.\text{area}_{h}(\Pi)=\inf\{\text{area}_{h}(\Sigma):\Sigma\in\Pi\}.

Then the minimal surface entropy with respect to hh is defined as follows.

(1.1) E(h)=limϵ0liminfLln#{areah(Π)4π(L1):ΠSϵ(M)}LlnL.E(h)=\underset{\epsilon\rightarrow 0}{\lim}\,\underset{L\rightarrow\infty}{\lim\inf}\,\dfrac{\ln\#\{\text{area}_{h}(\Pi)\leq 4\pi(L-1):\Pi\in S_{\epsilon}(M)\}}{L\ln L}.

According to [4] and [13], for odd-dimensional manifolds, and among metrics with sectional curvature less than or equal to 1-1, E(h)E(h) attains the minimum at the hyperbolic metric h0h_{0}, and E(h0)=2E(h_{0})=2. On the other hand, Theorem 1.1 of [19] shows that when n=3n=3, E(h)E(h) is maximized at h0h_{0} among all metrics with scalar curvature greater than or equal to 6-6.

1.3. Main theorems

In the following results, we assume (M,h0)(M,h_{0}) is a closed hyperbolic manifold of dimension n3n\geq 3.

Theorem 1.2.

There exists a small neighborhood 𝒰\mathcal{U} of h0h_{0} in the metric space of MM, such that for any Riemannian metric h𝒰h\in\mathcal{U} on MM with Rhn(n1)R_{h}\geq-n(n-1), we have

AreaId(h/h0)1.\text{Area}_{\text{Id}}(h/h_{0})\geq 1.

The equality holds if and only if h=h0h=h_{0}.

Note that such a neighborhood is “cylindrical”, that is, if h𝒰h\in\mathcal{U}, then any metric hh^{\prime} in the conformal class of hh with Rhn(n1)R_{h^{\prime}}\geq-n(n-1) also belongs to 𝒰\mathcal{U}.

Let (N,g0)(N,g_{0}) be another closed hyperbolic manifold of the same dimension nn, the theorem leads to the following corollary.

Corollary 1.3.

There exists a small neighborhood 𝒰\mathcal{U} of g0g_{0} in the metric space of NN, such that for any metric g𝒰g\in\mathcal{U} with Rgn(n1)R_{g}\geq-n(n-1), and for any local diffeomorphism F:(N,g)(M,h0)F:(N,g)\rightarrow(M,h_{0}) with degree dd, we have

AreaF(g/h0)d.\text{Area}_{F}(g/h_{0})\geq d.

The equality holds if and only if FF is a local isometry between gg and h0h_{0}, i.e., g=F(h0)g=F^{*}(h_{0}).

Proof.

Since FF is a local diffeomorphism,

AreaF(g/h0)=dAreaId(g/F(h0)).\text{Area}_{F}(g/h_{0})=d\,\text{Area}_{\text{Id}}(g/F^{*}(h_{0})).

F(h0)F^{*}(h_{0}) is a hyperbolic metric on NN, so it’s isometric to g0g_{0} due to Mostow rigidity. Thus we obtain from the previous theorem the following inequality.

AreaF(g/h0)=dAreaId(g/g0)d.\text{Area}_{F}(g/h_{0})=d\,\text{Area}_{\text{Id}}(g/g_{0})\geq d.

In particular, when n=3n=3, we prove a more general result for any smooth map between homeomorphic 3-manifolds which is not necessarily a diffeomorphism. Let (M,h0)(M,h_{0}) and (N,g0)(N,g_{0}) be closed hyperbolic 3-manifolds.

Theorem 1.4.

If π1(M)π1(N)\pi_{1}(M)\cong\pi_{1}(N), then there exists a small neighborhood 𝒰\mathcal{U} of g0g_{0} in the C2C^{2}-topology, such that for any metric g𝒰g\in\mathcal{U} with Rg6R_{g}\geq-6, and for any smooth map F:(N,g)(M,h0)F:(N,g)\rightarrow(M,h_{0}), we have

degF=1andAreaF(g/h0)1.\text{deg}\,F=1\quad\text{and}\quad\text{Area}_{F}(g/h_{0})\geq 1.

In a different situation, suppose that π1(N)\pi_{1}(N) is isomorphic to an index dd subgroup G<π1(M)G<\pi_{1}(M). Let M~\tilde{M} denote the covering space of MM with π1(M~)=G\pi_{1}(\tilde{M})=G and let p:M~Mp:\tilde{M}\rightarrow M denote the covering map, the hyperbolic metric on M~\tilde{M} is still represented by h0h_{0}. We have the following corollary.

Corollary 1.5.

There exists a small neighborhood 𝒰\mathcal{U} of g0g_{0} in the C2C^{2}-topology, such that for any metric g𝒰g\in\mathcal{U} with Rg6R_{g}\geq-6, and for any smooth map F:(N,g)(M,h0)F:(N,g)\rightarrow(M,h_{0}) satisfying Fπ1(N)<pπ1(M~)F_{*}\pi_{1}(N)<p_{*}\pi_{1}(\tilde{M}), we have

degF=dandAreaF(g/h0)d.\text{deg}\,F=d\quad\text{and}\quad\text{Area}_{F}(g/h_{0})\geq d.
Proof.

Since Fπ1(N)<pπ1(M~)F_{*}\pi_{1}(N)<p_{*}\pi_{1}(\tilde{M}), FF can be lifted to a smooth map F~:NM~\tilde{F}:N\rightarrow\tilde{M}, so that pF~=Fp\circ\tilde{F}=F. Applying Theorem 1.4 to F~\tilde{F}, we have

degF~=1andAreaF~(g/h0)1.\text{deg}\,\tilde{F}=1\quad\text{and}\quad\text{Area}_{\tilde{F}}(g/h_{0})\geq 1.

Thus,

degF=dandAreaF(g/h0)=dAreaF~(g/h0)d.\text{deg}\,F=d\quad\text{and}\quad\text{Area}_{F}(g/h_{0})=d\,\text{Area}_{\tilde{F}}(g/h_{0})\geq d.

Besides, when the dimension of MM is an odd number, Hamenstädt [9] verified the existence of surface subgroups, and based on the result, the author discussed the corresponding minimal surface entropy and proved that E(h0)=2E(h_{0})=2 in [13]. As a result, we can extend Theorem 1.2 of [19] to any odd dimensions at least 3.

Theorem 1.6.

Let (M,h0)(M,h_{0}) be a closed hyperbolic manifold of an odd dimension n3n\geq 3. For any Riemannian metric hh on MM,

(1.2) AreaId(h/h0)E(h)E(h0)=2,\text{Area}_{\text{Id}}(h/h_{0})E(h)\geq E(h_{0})=2,

the equality holds if and only if h=ch0h=ch_{0} for some constant c>0c>0.

This inequality can be considered as an analogue of the following comparison related to the volume entropy and geodesic stretch (Theorem 1.2 of [16]). More generally, suppose that (M,h0)(M,h_{0}) is a compact space of any dimension with negative curvature, and hh is another metric on MM. It says that

Iμ0(h/h0)Evol(h)Evol(h0),I_{\mu_{0}}(h/h_{0})E_{vol}(h)\geq E_{vol}(h_{0}),

where μ0\mu_{0} is the Bowen-Margulis measure of h0h_{0} which is the unique measure with maximal measure theoretic entropy. The equality holds if and only if the geodesic flows of hh and h0h_{0} are time preserving conjugate after rescaling.

We briefly mention a similar result. Besson-Courtois-Gallot [3] showed that for any compact hyperbolic manifold (it also holds for compact locally symmetric spaces) (M,h0)(M,h_{0}) with dimension n3n\geq 3,

(volh(M)volh0(M))1nEvol(h)Evol(h0),\big{(}\frac{\text{vol}_{h}(M)}{\text{vol}_{h_{0}}(M)}\big{)}^{\frac{1}{n}}E_{vol}(h)\geq E_{vol}(h_{0}),

the inequality is sharp unless hh is isometric to h0h_{0} after rescaling. Calegari-Marques-Neves [4] conjectured the minimal surface analogue for closed hyperbolic 3-manifolds (also see [13] for higher odd-dimensional closed hyperbolic manifolds and locally symmetric spaces).

(volh(M)volh0(M))2nE(h)E(h0)=2,\big{(}\frac{\text{vol}_{h}(M)}{\text{vol}_{h_{0}}(M)}\big{)}^{\frac{2}{n}}E(h)\geq E(h_{0})=2,

the equality holds if and only if hh and h0h_{0} are isometric up to scaling.

The author believes that Theorem 1.2 and Theorem 1.6 also hold for the other compact locally symmetric spaces of rank one, which include complex hyperbolic manifolds, quaternionic hyperbolic manifolds, and the Cayley plane. To check the details, we refer the readers to the construction of the surface subgroups of locally symmetric spaces and the discussion of the equidistribution property in [14] and the corresponding definition of the minimal surface entropy in Section 5 of [13].

The organization of this paper is as follows. Firstly, in Section 2, we introduce notations and definitions that are used throughout this paper. In Section 3, we discuss the average area ratio for dimension n3n\geq 3 and prove Theorem 1.2. In Section 4, we establish the equidistribution property and average area ratio formula which will be used in the last two sections. And in Section 5, we consider the special case for 3-manifolds and give the proof for Theorem 1.4. Finally, Section 6 concentrates on the minimal surface entropy defined on odd-dimensional manifolds, and the proof of Theorem 1.6 will be given.

Acknowledgements

I would like to thank my advisor André Neves for his constant encouragement and all the useful suggestions related to this work.

2. Preliminaries

2.1. Gromov’s average area ratio

Let (M,h0)(M,h_{0}) be a closed hyperbolic manifold of dimension n3n\geq 3, and let (N,g)(N,g) be another closed Riemannian manifold of the same dimension. Suppose that F:(N,g)(M,h0)F:(N,g)\rightarrow(M,h_{0}) is a smooth map. For any (x,L)(x,L) in the Grassmannian bundle Gr2(N)Gr_{2}(N), |Λ2F(x,L)|g|\Lambda^{2}F(x,L)|_{g} denotes the Jacobian of dFxdF_{x} at plane LL. Take an arbitrary regular value yMy\in M of FF, for (y,P)Gr2(M)(y,P)\in Gr_{2}(M), we let

(2.1) |Λ2F|g1(y,P)=xF1(y)1|Λ2F(x,(dFx)1(P)|g.|\Lambda^{2}F|^{-1}_{g}(y,P)=\sum_{x\in F^{-1}(y)}\frac{1}{|\Lambda^{2}F(x,(dF_{x})^{-1}(P)|_{g}}.

This definition ([19]) is equivalent to Gromov’s definition ([8]):

|Λ2F|g1(y,P)=xF1(y)limδ0areag((dFx)1(Dδ))δ,|\Lambda^{2}F|^{-1}_{g}(y,P)=\sum_{x\in F^{-1}(y)}\lim_{\delta\rightarrow 0}\frac{\text{area}_{g}((dF_{x})^{-1}(D_{\delta}))}{\delta},

where DδD_{\delta} is a subset of the totally geodesic disc DnD\subset\mathbb{H}^{n} which is tangential to PP at xx, and DδD_{\delta} has area equal to δ\delta.

The average area ratio of FF is defined in [8] by

(2.2) AreaF(g/h0)=(y,P)Gr2(M)|Λ2F|g1(y,P)𝑑μh0,\text{Area}_{F}(g/h_{0})=\int_{(y,P)\in Gr_{2}(M)}|\Lambda^{2}F|^{-1}_{g}(y,P)\,d\mu_{h_{0}},

where μh0\mu_{h_{0}} stands for the unit volume measure on Gr2(M)Gr_{2}(M) with respect to the metric induced by h0h_{0}.

2.2. Normalized Total Scalar Curvature

Let (M,h0)(M,h_{0}) be a closed hyperbolic manifold of dimension n3n\geq 3, and let \mathcal{M} be the space of Riemannian metrics on MM. The total scalar curvature (or Einstein-Hilbert functional) ~:\tilde{\mathcal{E}}:\mathcal{M}\rightarrow\mathbb{R} is

~(h)=MRh𝑑Vh.\tilde{\mathcal{E}}(h)=\int_{M}R_{h}\,dV_{h}.

It is a Riemannian functional in the sense that it’s invariant under diffeomorphisms, but it’s not scale-invariant. To resolve this issue, we consider

(h)=(volh(M))2n1MRh𝑑Vh.\mathcal{E}(h)=(\text{vol}_{h}(M))^{\frac{2}{n}-1}\int_{M}R_{h}\,dV_{h}.

This is called the normalized total scalar curvature (or normalized Einstein-Hilbert functional) of MM. Under conformal deformations, the first variation of \mathcal{E} is

(h)l=n22nvolh(M)2n1MRhMRh𝑑Vh,lh𝑑Vh.\mathcal{E}^{\prime}(h)\cdot l=\frac{n-2}{2n}\text{vol}_{h}(M)^{\frac{2}{n}-1}\int_{M}\langle R_{h}-\fint_{M}R_{h}\,dV_{h},\,l\rangle_{h}\,dV_{h}.

It equals to zero provided that (M,h)(M,h) has constant scalar curvature. Assuming RhR_{h} is constant, we can simplify the full variation to

(h)l=volh(M)2n1M1nRhRich,lh𝑑Vh.\mathcal{E}^{\prime}(h)\cdot l=\text{vol}_{h}(M)^{\frac{2}{n}-1}\int_{M}\langle\frac{1}{n}R_{h}-Ric_{h},l\rangle_{h}\,dV_{h}.

Thus, a metric hh is critical if and only if (M,h)(M,h) is Einstein. In particular, the hyperbolic metric h0h_{0} is a critical point for \mathcal{E}. Furthermore, since \mathcal{E} is scale-invariant, from now on we may assume that for ht=h0+tlh_{t}=h_{0}+tl,

Mddt|t=0(deth0(ht))dVh0=12Mtrh0l𝑑Vh0=0.\int_{M}\frac{d}{dt}|_{t=0}(\sqrt{\text{det}_{h_{0}}(h_{t})})\,dV_{h_{0}}=\frac{1}{2}\int_{M}\text{tr}_{h_{0}}l\,dV_{h_{0}}=0.

Taking this into account, we obtain the second variation of \mathcal{E} at h0h_{0} as follows.

(2.3) ′′(h0)(l,l)=\displaystyle\mathcal{E}^{\prime\prime}(h_{0})(l,l)= volh0(M)2n1Md2dt2|t=0Rht2(n1)d2dt2|t=0(deth0(ht))\displaystyle\text{vol}_{h_{0}}(M)^{\frac{2}{n}-1}\int_{M}\frac{d^{2}}{dt^{2}}|_{t=0}R_{h_{t}}-2(n-1)\frac{d^{2}}{dt^{2}}|_{t=0}(\sqrt{\text{det}_{h_{0}}(h_{t})})
+2ddt|t=0Rhtddt|t=0(deth0(ht))dVh0,\displaystyle+2\frac{d}{dt}|_{t=0}R_{h_{t}}\,\frac{d}{dt}|_{t=0}(\sqrt{\text{det}_{h_{0}}(h_{t})})\,dV_{h_{0}},

Substituting the formulas

d2dt2deth0(ht)=(14(trhtl)212trht(l2))deth0(ht),\frac{d^{2}}{dt^{2}}\sqrt{\text{det}_{h_{0}}(h_{t})}=\big{(}\frac{1}{4}(\text{tr}_{h_{t}}l)^{2}-\frac{1}{2}\text{tr}_{h_{t}}(l^{2})\big{)}\sqrt{\text{det}_{h_{0}}(h_{t})},
ddtRht=Δht(trhtl)+δht2lRicht,lht,\frac{d}{dt}R_{h_{t}}=-\Delta_{h_{t}}(\text{tr}_{h_{t}}l)+\delta_{h_{t}}^{2}l-\langle Ric_{h_{t}},l\rangle_{h_{t}},
ddtRicht=12Δhtl+12Richt(l)+12l(Richt)Rmhtlδht(δhtl)12ht2(trhtl),\frac{d}{dt}Ric_{h_{t}}=-\frac{1}{2}\Delta_{h_{t}}l+\frac{1}{2}Ric_{h_{t}}(l)+\frac{1}{2}l(Ric_{h_{t}})-Rm_{h_{t}}*l-\delta^{*}_{h_{t}}(\delta_{h_{t}}l)-\frac{1}{2}\nabla_{h_{t}}^{2}(\text{tr}_{h_{t}}l),

where (Rmhtl)ij=Rikjmlkm(Rm_{h_{t}}*l)_{ij}=R_{ikjm}l^{km}, we obtain that

′′(h0)(l,l)=\displaystyle\mathcal{E}^{\prime\prime}(h_{0})(l,l)= volh0(M)2n1M12Δh0l+Rmh0l+δh0(δh0l)\displaystyle\text{vol}_{h_{0}}(M)^{\frac{2}{n}-1}\int_{M}\langle\frac{1}{2}\Delta_{h_{0}}l+Rm_{h_{0}}*l+\delta^{*}_{h_{0}}(\delta_{h_{0}}l)
+n12(trh0l)h012(Δh0(trh0l))h0+(δh02l)h0,lh0dVh0.\displaystyle+\frac{n-1}{2}(\text{tr}_{h_{0}}l)h_{0}-\frac{1}{2}(\Delta_{h_{0}}(\text{tr}_{h_{0}}l))h_{0}+(\delta_{h_{0}}^{2}l)h_{0}\,,l\rangle_{h_{0}}\,dV_{h_{0}}.

According to Ebin’s Slice Theorem (see [6]), for any hh\in\mathcal{M} lying in a small neighborhood of h0h_{0}, there exist ϕDiff(M)\phi\in\text{Diff}(M), fC(M)f\in C^{\infty}(M), and a transverse-traceless tensor lTTl_{TT}, i.e., δh0(lTT)=0\delta_{h_{0}}(l_{TT})=0 and trh0(lTT)=0\text{tr}_{h_{0}}(l_{TT})=0, such that ϕh=h0+fh0+lTT\phi^{*}h=h_{0}+fh_{0}+l_{TT}. Then we can simplify the second variation.

(2.4) ′′(h0)(fh0,fh0)=(n1)(n2)2volh0(M)2n1MΔh0fnf,fh0𝑑Vh0.\mathcal{E}^{\prime\prime}(h_{0})(fh_{0},fh_{0})=-\frac{(n-1)(n-2)}{2}\text{vol}_{h_{0}}(M)^{\frac{2}{n}-1}\int_{M}\langle\Delta_{h_{0}}f-nf,\,f\rangle_{h_{0}}\,dV_{h_{0}}.

And we also get

(2.5) ′′(h0)(lTT,lTT)=volh0(M)2n1M12Δh0lTT+Rmh0lTT,lTTh0𝑑Vh0.\mathcal{E}^{\prime\prime}(h_{0})(l_{TT},l_{TT})=\text{vol}_{h_{0}}(M)^{\frac{2}{n}-1}\int_{M}\langle\frac{1}{2}\Delta_{h_{0}}l_{TT}+Rm_{h_{0}}*l_{TT},l_{TT}\rangle_{h_{0}}\,dV_{h_{0}}.

Thus,

(h)\displaystyle\mathcal{E}(h) =(ϕh)=(h0+fh0+lTT)\displaystyle=\mathcal{E}(\phi^{*}h)=\mathcal{E}(h_{0}+fh_{0}+l_{TT})
=(h0)+′′(h0)(fh0,fh0)+′′(h0)(lTT,lTT)+higher order variations.\displaystyle=\mathcal{E}(h_{0})+\mathcal{E}^{\prime\prime}(h_{0})(fh_{0},fh_{0})+\mathcal{E}^{\prime\prime}(h_{0})(l_{TT},l_{TT})+\text{higher order variations}.

Taking the estimates of (2.4) and (2.5) into consideration, we can use this expansion to discuss the local behavior of \mathcal{E}. In particular, as discussed in [2], \mathcal{E} reaches a local maximum at h0h_{0}, and there exists C>0C>0, so that for any metric hh in a small neighborhood of h0h_{0} in \mathcal{M},

(h0)(h)Cd(h,h0)2,where d(h,h0)=infϕDiff(M)|ϕhh0|H1(M,h0).\mathcal{E}(h_{0})-\mathcal{E}(h)\geq C\,d(h,h_{0})^{2},\quad\text{where }d(h,h_{0})=\underset{\phi\in\text{Diff}(M)}{\inf}|\phi^{*}h-h_{0}|_{H^{1}(M,h_{0})}.

The inequality is sharp unless hh and h0h_{0} are isometric via some ϕDiff(M)\phi\in\text{Diff}(M).

2.3. Equidistribution

Suppose that the hyperbolic manifold MM has an odd dimension n3n\geq 3. According to the construction by Hamenstädt [9], for any small number ϵ>0\epsilon>0, there is an essential surface Σϵ\Sigma_{\epsilon} in MM which is sufficiently well-distributed and (1+ϵ)(1+\epsilon)-quasigeodesic (i.e. the geodesics on the surface with respect to intrinsic distance are (1+ϵ,ϵ)(1+\epsilon,\epsilon)-quasigeodesics in MM). And as discussed in [13], Σϵ\Sigma_{\epsilon} determines an (1+O(ϵ))(1+O(\epsilon))-quasiconformal map on Sn1S_{\infty}^{n-1}, and thus associated with an element in Sϵ(M)S_{\epsilon}(M).

Furthermore, let G(M,g,ϵ)G(M,g,\epsilon) denote the subset of S(M,g,ϵ)S(M,g,\epsilon) consisting of homotopy classes of finite covers of Σϵ\Sigma_{\epsilon} that have genus at most gg. It has cardinality comparable to g2gg^{2g}. Moreover, let SiS_{i} denote the minimal representative of an element in G(M,gi,1i)G(M,g_{i},\frac{1}{i}), then it is homotopic to an (1+1i)(1+\frac{1}{i})-quasigeodesic surface Σi\Sigma_{i}. From Lemma 4.3 of [4], for any continuous function ff on MM, the unit Radon measure induced by integration over SiS_{i} satisfies

(2.6) limi1areah0(Si)Sif𝑑Ah0=ν(f),\underset{i\rightarrow\infty}{\lim}\,\frac{1}{\text{area}_{h_{0}}(S_{i})}\int_{S_{i}}f\,dA_{h_{0}}=\nu(f),

where the limiting measure ν\nu is positive on any non-empty open set of MM.

Notice that the measure ν\nu is not necessarily identified with the unit Radon measure on MM induced by integration over MM, the latter measure is denoted by μ\mu. However, in this paper, we need to find a sequence of minimal surfaces whose Radon measures defined above converge to μ\mu. To solve this problem, we introduce Labourie’s construction [17] in Section 4. And we stress that both methods of Hamenstädt and Labourie require that MM has an odd dimension.

3. Gromov’s Conjecture in Higher Dimensions

Throughout this section, we can choose the dimension of MM to be any integer n3n\geq 3. The proof of Theorem 1.2 separates into two parts. If hh is an arbitrary metric conformal to a metric h¯\bar{h} with constant scalar curvature, we compare their average area ratios in Theorem 3.1. And if hh is a metric with constant scalar curvature different from h0h_{0}, we make use of the evaluations of normalized Einstein-Hilbert functional in [2].

3.1. Conformal deformations

Firstly, given a metric hh on MM, we look at the conformal class of hh. Since MM admits a hyperbolic metric, every conformal class [h][h] must be scalar negative, i.e., it has a metric with negative scalar curvature. And according to the Yamabe problem, after rescaling, there exists a unique metric h¯[h]\bar{h}\in[h] with constant scalar curvature c<0c<0. In Theorem 1.2, we assume that cn(n1)c\geq-n(n-1). Set h¯=cn(n1)h¯\bar{h}^{\prime}=\frac{c}{-n(n-1)}\bar{h}, so Rh¯n(n1)R_{\bar{h}^{\prime}}\equiv-n(n-1), and for any surface or 2-dimensional subset SS in MM, we have

areah¯(S)=cn(n1)areah¯(S)areah¯(S).\text{area}_{\bar{h}^{\prime}}(S)=\frac{c}{-n(n-1)}\text{area}_{\bar{h}}(S)\leq\text{area}_{\bar{h}}(S).

Therefore, to prove the theorem, we may assume that Rh¯n(n1)R_{\bar{h}}\equiv-n(n-1).

Theorem 3.1.

Suppose the scalar curvature Rhn(n1)R_{h}\geq-n(n-1), then we have

(3.1) AreaId(h/h0)AreaId(h¯/h0).\text{Area}_{\text{Id}}(h/h_{0})\geq\text{Area}_{\text{Id}}(\bar{h}/h_{0}).

Furthermore, for any surface subgroup ΠSϵ(M)\Pi\in S_{\epsilon}(M),

(3.2) areah(Π)areah¯(Π),\text{area}_{h}(\Pi)\geq\text{area}_{\bar{h}}(\Pi),

and as an immediate result,

(3.3) E(h)E(h¯).E(h)\leq E(\bar{h}).

Each of the above equalities holds if and only if h=h¯h=\bar{h}.

Proof.

Set h=e2ϕh¯h=e^{2\phi}\bar{h}. The conformal factor ϕ\phi satisfies that

e2ϕRh\displaystyle e^{2\phi}R_{h} =Rh¯2(n1)Δh¯ϕ(n2)(n1)|dϕ|h¯2\displaystyle=R_{\bar{h}}-2(n-1)\Delta_{\bar{h}}\phi-(n-2)(n-1)|d\phi|_{\bar{h}}^{2}
=n(n1)2(n1)Δh¯ϕ(n2)(n1)|dϕ|h¯2.\displaystyle=-n(n-1)-2(n-1)\Delta_{\bar{h}}\phi-(n-2)(n-1)|d\phi|_{\bar{h}}^{2}.

Let ϕmin:=minxMϕ(x)\phi_{min}:=\underset{x\in M}{\min}\,\phi(x), we obtain Δh¯ϕmin0\Delta_{\bar{h}}\phi_{min}\geq 0, and it yields that

(3.4) n(n1)e2ϕmine2ϕminRhn(n1).-n(n-1)e^{2\phi_{min}}\leq e^{2\phi_{min}}R_{h}\leq-n(n-1).

Thus, for any subset DδD_{\delta} of a totally geodesic disc in n\mathbb{H}^{n} with hyperbolic area equal to δ\delta, we have

e2ϕmin1areah(Dδ)=Dδe2ϕ𝑑Ah¯areah¯(Dδ).e^{2\phi_{min}}\geq 1\quad\Longrightarrow\quad\text{area}_{h}(D_{\delta})=\int_{D_{\delta}}e^{2\phi}\,dA_{\bar{h}}\geq\text{area}_{\bar{h}}(D_{\delta}).

The inequality (3.1) follows from the definition of the average area ratio (2.2).

In addition, it also shows that for any surface SMS\in M,

areah(S)=Se2ϕ𝑑Ah¯areah¯(S).\text{area}_{h}(S)=\int_{S}e^{2\phi}\,dA_{\bar{h}}\geq\text{area}_{\bar{h}}(S).

Thus we conclude (3.2), and the comparison of entropy (3.3) is a direct corollary of the definition (1.1).

Moreover, if any equality in the theorem holds, the inequality 3.4 implies that Rhn(n1)R_{h}\equiv-n(n-1), due to the uniqueness of the solution to Yamabe problem, hh must be identical to h¯\bar{h}. ∎

3.2. Definition of functional 𝒜\mathcal{A}

Let \mathcal{M} be the space of all Riemannian metrics on MM, and let R\mathcal{M}_{R} be the subset consisting of metrics with constant scalar curvature n(n1)-n(n-1). From the previous section, it remains to consider metrics in R\mathcal{M}_{R}.

Define a functional 𝒜\mathcal{A} from the space of Riemannian metrics on MM to \mathbb{R} as follows.

𝒜(h)=\displaystyle\mathcal{A}(h)= (x,P)Gr2MRh(x)|Λ2Id|h1(x,P)𝑑μh0\displaystyle\int_{(x,P)\in Gr_{2}M}R_{h}(x)\,|\Lambda^{2}\text{Id}|^{-1}_{h}(x,P)\,d\mu_{h_{0}}
=\displaystyle= limδ0(x,P)Gr2MRh(x)areah(Dδ(P))δ𝑑μh0\displaystyle\lim_{\delta\rightarrow 0}\int_{(x,P)\in Gr_{2}M}R_{h}(x)\,\frac{\text{area}_{h}(D_{\delta}(P))}{\delta}\,d\mu_{h_{0}}
=\displaystyle= limδ0(x,P)Gr2MRh(x)Dδ(P)deth0(h|Dδ(P))𝑑Ah0𝑑μh0,\displaystyle\lim_{\delta\rightarrow 0}\int_{(x,P)\in Gr_{2}M}R_{h}(x)\,\fint_{D_{\delta}(P)}\sqrt{\text{det}_{h_{0}}(h|_{D_{\delta}(P)})}\,dA_{h_{0}}\,d\mu_{h_{0}},

where Dδ(P)D_{\delta}(P) is a subset of the totally geodesic disc in n\mathbb{H}^{n} that is tangential to PP at xx, and Dδ(P)D_{\delta}(P) has hyperbolic area equal to δ\delta, μh0\mu_{h_{0}} is the unit volume measure on Gr2MGr_{2}M with respect to the metric induced by h0h_{0}. Notice that the metric hRh\in\mathcal{M}_{R} satisfies

𝒜(h)\displaystyle\mathcal{A}(h) =n(n1)(x,P)Gr2M|Λ2Id|h1(x,P)𝑑μh0\displaystyle=-n(n-1)\,\int_{(x,P)\in Gr_{2}M}|\Lambda^{2}\text{Id}|^{-1}_{h}(x,P)\,d\mu_{h_{0}}
=𝒜(h0)(x,P)Gr2M|Λ2Id|h1(x,P)𝑑μh0.\displaystyle=\mathcal{A}(h_{0})\,\int_{(x,P)\in Gr_{2}M}|\Lambda^{2}\text{Id}|^{-1}_{h}(x,P)\,d\mu_{h_{0}}.

Therefore, from the definition of the average area ratio (2.2), to deduce that if hh is in a small neighborhood of h0h_{0} in R\mathcal{M}_{R}, then

AreaId(h/h0)=(x,P)Gr2M|Λ2Id|h1(x,P)𝑑μh01,\text{Area}_{\text{Id}}(h/h_{0})=\int_{(x,P)\in Gr_{2}M}|\Lambda^{2}\text{Id}|^{-1}_{h}(x,P)\,d\mu_{h_{0}}\geq 1,

we only need to show that 𝒜\mathcal{A} attains a local maximum at h0h_{0}.

3.3. Proof of Theorem 1.2

To see this, we’ll discuss the first and second variations of 𝒜\mathcal{A} in detail. Before start, we notice that the functional 𝒜\mathcal{A} is scale-invariant, so it suffices to assume volh0(M)=1\text{vol}_{h_{0}}(M)=1, and the symmetric tensor l=hh0l=h-h_{0} satisfies Mtrh0l𝑑Vh0=0\int_{M}\text{tr}_{h_{0}}l\,dV_{h_{0}}=0. Let ht=h0+tlh_{t}=h_{0}+tl, we rewrite 𝒜(ht)\mathcal{A}(h_{t}) as

𝒜(ht)=xMRht(x)aht(x)𝑑Vh0,\mathcal{A}(h_{t})=\int_{x\in M}R_{h_{t}}(x)\,a_{h_{t}}(x)\,dV_{h_{0}},

where

aht(x)=limδ0PGr2MxDδ(P)deth0(ht|Dδ(P))𝑑Vh0𝑑νh0.a_{h_{t}}(x)=\lim_{\delta\rightarrow 0}\fint_{P\in Gr_{2}M_{x}}\fint_{D_{\delta}(P)}\sqrt{\text{det}_{h_{0}}(h_{t}|_{D_{\delta}(P)})}\,dV_{h_{0}}d\nu_{h_{0}}.

Then we have

(3.5) ddt𝒜(ht)=MddtRhtaht+RhtddtahtdVh0.\frac{d}{dt}\mathcal{A}(h_{t})=\int_{M}\frac{d}{dt}R_{h_{t}}\,a_{h_{t}}+R_{h_{t}}\frac{d}{dt}a_{h_{t}}\,dV_{h_{0}}.

In addition,

(3.6) ddtRht=Δht(trhtl)+δht2lRicht,lht,\frac{d}{dt}R_{h_{t}}=-\Delta_{h_{t}}(\text{tr}_{h_{t}}l)+\delta_{h_{t}}^{2}l-\langle Ric_{h_{t}},l\rangle_{h_{t}},

where Δht\Delta_{h_{t}} is the rough Laplacian with negative eigenvalues, and δht2\delta_{h_{t}}^{2} is the double divergence operator. In particular, when t=0t=0, applying the Stokes’ theorem, we have

Mddt|t=0Rhtah0dVh0=MRich0,lh0dVh0=(n1)Mtrh0l𝑑Vh0.\int_{M}\frac{d}{dt}|_{t=0}\,R_{h_{t}}\,a_{h_{0}}\,dV_{h_{0}}=\int_{M}-\langle Ric_{h_{0}},l\rangle_{h_{0}}\,dV_{h_{0}}=(n-1)\int_{M}\text{tr}_{h_{0}}l\,dV_{h_{0}}.

On the other hand, since

ddtdeth0(ht|Dδ(P))=12trhtl|Dδ(P)deth0(ht|Dδ(P)),\frac{d}{dt}\sqrt{\text{det}_{h_{0}}(h_{t}|_{D_{\delta}(P)})}=\frac{1}{2}\text{tr}_{h_{t}}l|_{D_{\delta}(P)}\,\sqrt{\text{det}_{h_{0}}(h_{t}|_{D_{\delta}(P)})},

let λ1,,λn\lambda_{1},\cdots,\lambda_{n} be the eigenvalues of the matrix ll at point xMx\in M with respect to h0h_{0}, we obtain by computation that

ddt|t=0aht\displaystyle\frac{d}{dt}|_{t=0}\,a_{h_{t}} =limδ012PGr2MxDδ(P)trh0l|Dδ(P)dAh0dνh0\displaystyle=\lim_{\delta\rightarrow 0}\frac{1}{2}\fint_{P\in Gr_{2}M_{x}}\fint_{D_{\delta}(P)}\text{tr}_{h_{0}}l|_{D_{\delta}(P)}\,dA_{h_{0}}d\nu_{h_{0}}
=12i<jλi+λj(n2)=12(n1)i=1nλi(n2)\displaystyle=\frac{1}{2}\,\frac{\sum_{i<j}\lambda_{i}+\lambda_{j}}{\binom{n}{2}}=\frac{1}{2}\,\frac{(n-1)\sum_{i=1}^{n}\lambda_{i}}{\binom{n}{2}}
=1ntrh0l.\displaystyle=\frac{1}{n}\text{tr}_{h_{0}}l.

It follows that for any symmetric 2-tensor ll,

𝒜(h0)l=(n1)Mtrh0l𝑑Vh0n(n1)1nMtrh0l𝑑Vh0=0,\mathcal{A}^{\prime}(h_{0})\cdot l=(n-1)\int_{M}\text{tr}_{h_{0}}l\,dV_{h_{0}}-n(n-1)\frac{1}{n}\int_{M}\text{tr}_{h_{0}}l\,dV_{h_{0}}=0,

thus h0h_{0} is a critical point of 𝒜\mathcal{A}.

Now we proceed to compute the second variation at t=0t=0. Note that 𝒜\mathcal{A} is an analogue of the normalized total scalar curvature \mathcal{E}, it’s easier to compare their second variations using the computation in Section 2.2. Based on (3.5) and (2.3),

(3.7) 𝒜′′(h0)(l,l)′′(h0)(l,l)\displaystyle\mathcal{A}^{\prime\prime}(h_{0})(l,l)-\mathcal{E}^{\prime\prime}(h_{0})(l,l)
=\displaystyle= Mn(n1)d2dt2|t=0aht+2(n1)d2dt2|t=0(deth0(ht))dVh0\displaystyle\int_{M}-n(n-1)\frac{d^{2}}{dt^{2}}|_{t=0}\,a_{h_{t}}+2(n-1)\frac{d^{2}}{dt^{2}}|_{t=0}(\sqrt{\text{det}_{h_{0}}(h_{t})})\,dV_{h_{0}}
+M2ddt|t=0Rht(ddt|t=0(ahtdeth0(ht))dVh0\displaystyle+\int_{M}2\frac{d}{dt}|_{t=0}R_{h_{t}}(\frac{d}{dt}|_{t=0}(a_{h_{t}}-\sqrt{\text{det}_{h_{0}}(h_{t})})\,dV_{h_{0}}
=\displaystyle= Mn(n1)d2dt2|t=0aht+2(n1)(14(trh0l)212trh0(l2))dVh0\displaystyle\int_{M}-n(n-1)\frac{d^{2}}{dt^{2}}|_{t=0}\,a_{h_{t}}+2(n-1)\big{(}\frac{1}{4}(\text{tr}_{h_{0}}l)^{2}-\frac{1}{2}\text{tr}_{h_{0}}(l^{2})\big{)}\,dV_{h_{0}}
+2M(Δh0(trh0l)+δh02l+(n1)trh0l)(1n12)trh0l𝑑Vh0.\displaystyle+2\int_{M}\big{(}-\Delta_{h_{0}}(\text{tr}_{h_{0}}l)+\delta_{h_{0}}^{2}l+(n-1)\text{tr}_{h_{0}}l\big{)}(\frac{1}{n}-\frac{1}{2})\text{tr}_{h_{0}}l\,dV_{h_{0}}.

Next, using

d2dt2deth0(ht|Dδ(P))=(14(trhtl|Dδ(P))212trht(l|Dδ(P)2))deth0(ht|Dδ(P)),\frac{d^{2}}{dt^{2}}\sqrt{\text{det}_{h_{0}}(h_{t}|_{D_{\delta}(P)})}=\big{(}\frac{1}{4}(\text{tr}_{h_{t}}l|_{D_{\delta}(P)})^{2}-\frac{1}{2}\text{tr}_{h_{t}}(l|_{D_{\delta}(P)}^{2})\big{)}\sqrt{\text{det}_{h_{0}}(h_{t}|_{D_{\delta}(P)})},

we estimate the first term on the right-hand side of (3.7).

(3.8) n(n1)Md2dt2|t=0ahtdVh0\displaystyle-n(n-1)\int_{M}\frac{d^{2}}{dt^{2}}|_{t=0}\,a_{h_{t}}\,dV_{h_{0}}
=\displaystyle= n(n1)limδ0xMPGr2MxDδ(P)14(trh0l|Dδ(P))212trh0(l|Dδ(P)2)dAh0dνh0dVh0\displaystyle-n(n-1)\lim_{\delta\rightarrow 0}\,\int_{x\in M}\fint_{P\in Gr_{2}M_{x}}\fint_{D_{\delta}(P)}\frac{1}{4}(\text{tr}_{h_{0}}l|_{D_{\delta}(P)})^{2}-\frac{1}{2}\text{tr}_{h_{0}}(l|_{D_{\delta}(P)}^{2})\,dA_{h_{0}}d\nu_{h_{0}}dV_{h_{0}}
=\displaystyle= n(n1)Mi<j14(λi+λj)212(λi2+λj2)(n2)𝑑Vh0\displaystyle-n(n-1)\int_{M}\frac{\sum_{i<j}\frac{1}{4}(\lambda_{i}+\lambda_{j})^{2}-\frac{1}{2}(\lambda_{i}^{2}+\lambda_{j}^{2})}{\binom{n}{2}}\,dV_{h_{0}}
=\displaystyle= n(n1)M14(i=1nλi)2n4i=1nλi2(n2)𝑑Vh0\displaystyle-n(n-1)\int_{M}\frac{\frac{1}{4}(\sum_{i=1}^{n}\lambda_{i})^{2}-\frac{n}{4}\sum_{i=1}^{n}\lambda_{i}^{2}}{\binom{n}{2}}\,dV_{h_{0}}
=\displaystyle= M12(trh0l)2+n2trh0(l2)dVh0.\displaystyle\int_{M}-\frac{1}{2}(\text{tr}_{h_{0}}l)^{2}+\frac{n}{2}\text{tr}_{h_{0}}(l^{2})\,dV_{h_{0}}.

Combining (3.7) and (3.8), we have

(3.9) 𝒜′′(h0)(l,l)=\displaystyle\mathcal{A}^{\prime\prime}(h_{0})(l,l)= ′′(h0)(l,l)Mn22trh0(l2)𝑑Vh0\displaystyle\mathcal{E}^{\prime\prime}(h_{0})(l,l)-\int_{M}\frac{n-2}{2}\text{tr}_{h_{0}}(l^{2})\,dV_{h_{0}}
M(n2)22n(trh0l)2+n2n|h0(trh0l)|2+n2nδh02ltrh0ldVh0.\displaystyle-\int_{M}\frac{(n-2)^{2}}{2n}(\text{tr}_{h_{0}}l)^{2}+\frac{n-2}{n}|\nabla_{h_{0}}(\text{tr}_{h_{0}}l)|^{2}+\frac{n-2}{n}\delta_{h_{0}}^{2}l\,\text{tr}_{h_{0}}l\,dV_{h_{0}}.

To simplify this quadratic form, we decompose ll into three parts. Applying the decomposition of space of symmetric tensors for a compact Einstein manifold other than the standard sphere (Theorem 4.60 of [1]), we have

Th0=C(M)h0Th0(Diff(M)(h0))TTh0,T_{h_{0}}\mathcal{M}=C^{\infty}(M)\cdot h_{0}\oplus T_{h_{0}}(\text{Diff}(M)(h_{0}))\oplus TT_{h_{0}},

where C(M)h0C^{\infty}(M)\cdot h_{0} represents the conformal deformations of h0h_{0}, Diff(M)(h0)\text{Diff}(M)(h_{0}) is the action of the diffeomorphism group on h0h_{0}, and TTh0=kertrh0kerδh0TT_{h_{0}}=\ker\text{tr}_{h_{0}}\cap\ker\delta_{h_{0}} stands for the set of transverse-traceless tensors. Let hRh\in\mathcal{M}_{R}, and let l=hh0l=h-h_{0}, it decomposes into l=fh0+lD+lTTl=fh_{0}+l_{D}+l_{TT}, where fh0C(M)h0fh_{0}\in C^{\infty}(M)\cdot h_{0}, lDTh0(Diff(M)(h0))l_{D}\in T_{h_{0}}(\text{Diff}(M)(h_{0})), and lTTTTh0l_{TT}\in TT_{h_{0}}. The second variation of 𝒜\mathcal{A} has the form

𝒜′′(h0)(l,l)\displaystyle\mathcal{A}^{\prime\prime}(h_{0})(l,l)
=\displaystyle= 𝒜′′(h0)(lD+lTT,lD+lTT)+𝒜′′(h0)(fh0,lD)+𝒜′′(h0)(lD,fh0)\displaystyle\mathcal{A}^{\prime\prime}(h_{0})(l_{D}+l_{TT},l_{D}+l_{TT})+\mathcal{A}^{\prime\prime}(h_{0})(fh_{0},l_{D})+\mathcal{A}^{\prime\prime}(h_{0})(l_{D},fh_{0})
+𝒜′′(h0)(fh0,lTT)+𝒜′′(h0)(lTT,fh0)+𝒜′′(h0)(fh0,fh0)\displaystyle+\mathcal{A}^{\prime\prime}(h_{0})(fh_{0},l_{TT})+\mathcal{A}^{\prime\prime}(h_{0})(l_{TT},fh_{0})+\mathcal{A}^{\prime\prime}(h_{0})(fh_{0},fh_{0})
\displaystyle\leq 𝒜′′(h0)(lD+lTT,lD+lTT)+C0|fh0|H1(M,h0)|l|H1(M,h0).\displaystyle\mathcal{A}^{\prime\prime}(h_{0})(l_{D}+l_{TT},l_{D}+l_{TT})+C_{0}|fh_{0}|_{H^{1}(M,h_{0})}|l|_{H^{1}(M,h_{0})}.

By Theorem 3.1, to check that 𝒜\mathcal{A} reaches a local maximum at h0h_{0} on R\mathcal{M}_{R}, it remains to analyze the sign of 𝒜′′(h0)(lD+lTT,lD+lTT)\mathcal{A}^{\prime\prime}(h_{0})(l_{D}+l_{TT},l_{D}+l_{TT}) and estimate |fh0|H1(M,h0)|fh_{0}|_{H^{1}(M,h_{0})}. To see these, we prove the following lemmas.

Lemma 3.2.

There exists a constant C>0C>0, such that in the decomposition of ll, lDTh0(Diff(M)(h0))l_{D}\in T_{h_{0}}(\text{Diff}(M)(h_{0})) and lTTTTh0l_{TT}\in TT_{h_{0}} satisfy

𝒜′′(h0)((lD+lTT,lD+lTT)C(|lTT|H1(M,h0)2+|X|L2(M,h0)2),\mathcal{A}^{\prime\prime}(h_{0})((l_{D}+l_{TT},l_{D}+l_{TT})\leq-C(|l_{TT}|^{2}_{H^{1}(M,h_{0})}+|X|^{2}_{L^{2}(M,h_{0})}),

where Xh0=lD\mathcal{L}_{X}h_{0}=l_{D}.

Lemma 3.3.

For any ϵ>0\epsilon>0, we can find a C2C^{2}-neighborhood 𝒰ϵ,R\mathcal{U}_{\epsilon,R} of h0h_{0} on R\mathcal{M}_{R} and c>0c>0, such that for any h𝒰ϵ,Rh\in\mathcal{U}_{\epsilon,R}, fh0C(M)h0fh_{0}\in C^{\infty}(M)\cdot h_{0} in the decomposition of l=hh0l=h-h_{0} satisfies

|fh0|H1(M,h0)cϵ|l|H1(M,h0).|fh_{0}|_{H^{1}(M,h_{0})}\leq c\,\epsilon|l|_{H^{1}(M,h_{0})}.
Proof of Lemma 3.2.
(3.10) 𝒜′′(h0)((lD+lTT,lD+lTT)\displaystyle\mathcal{A}^{\prime\prime}(h_{0})((l_{D}+l_{TT},l_{D}+l_{TT})
=\displaystyle= 𝒜′′(h0)(lTT,lTT)+𝒜′′(h0)((lD,lD)+𝒜′′(h0)(lD,lTT)+𝒜′′(h0)(lTT,lD).\displaystyle\mathcal{A}^{\prime\prime}(h_{0})(l_{TT},l_{TT})+\mathcal{A}^{\prime\prime}(h_{0})((l_{D},l_{D})+\mathcal{A}^{\prime\prime}(h_{0})(l_{D},l_{TT})+\mathcal{A}^{\prime\prime}(h_{0})(l_{TT},l_{D}).

To find an upper bound of the first term, we use the estimate in Lemma 2.9 of [2]. There exists a constant C1>0C_{1}>0, such that

′′(h0)(lTT,lTT)=M12Δh0lTT+Rmh0lTT,lTTh0𝑑Vh0C1|lTT|H1(M,h0)2.\mathcal{E}^{\prime\prime}(h_{0})(l_{TT},l_{TT})=\int_{M}\langle\frac{1}{2}\Delta_{h_{0}}l_{TT}+Rm_{h_{0}}*l_{TT},l_{TT}\rangle_{h_{0}}\,dV_{h_{0}}\leq-C_{1}|l_{TT}|^{2}_{H^{1}(M,h_{0})}.

And thus, substituting the above inequality and trh0lTT=δh0lTT=0\text{tr}_{h_{0}}l_{TT}=\delta_{h_{0}}l_{TT}=0 into (3.9), we obtain

(3.11) 𝒜′′(h0)(lTT,lTT)\displaystyle\mathcal{A}^{\prime\prime}(h_{0})(l_{TT},l_{TT}) =′′(h0)(lTT,lTT)n22Mtrh0(lTT2)𝑑Vh0\displaystyle=\mathcal{E}^{\prime\prime}(h_{0})(l_{TT},l_{TT})-\frac{n-2}{2}\int_{M}\text{tr}_{h_{0}}(l_{TT}^{2})\,dV_{h_{0}}
C1|lTT|H1(M,h0)2.\displaystyle\leq-C_{1}|l_{TT}|^{2}_{H^{1}(M,h_{0})}.

To deal with the remaining terms of (3.10), we apply the diffeomorphism invariance property of \mathcal{E}, which says

(3.12) ′′(h0)(lD,)=′′(h0)(,lD)=0.\mathcal{E}^{\prime\prime}(h_{0})(l_{D},\cdot)=\mathcal{E}^{\prime\prime}(h_{0})(\cdot,l_{D})=0.

Moreover, for any ϕDiff(M)\phi\in\text{Diff}(M),

(3.13) Rϕh0Rh0n(n1)Rh0lD=0.R_{\phi^{*}h_{0}}\equiv R_{h_{0}}\equiv-n(n-1)\quad\Longrightarrow\quad R_{h_{0}}^{\prime}\cdot l_{D}=0.

Therefore, the second variations comparison (3.7), in company with (3.8), says that

(3.14) 𝒜′′(h0)((lD,lD)=n22M(trh0lD)2trh0(lD2)dVh0.\mathcal{A}^{\prime\prime}(h_{0})((l_{D},l_{D})=\frac{n-2}{2}\int_{M}(\text{tr}_{h_{0}}l_{D})^{2}-\text{tr}_{h_{0}}(l_{D}^{2})\,dV_{h_{0}}.

Since lDTh0(Diff(M)(h0))l_{D}\in T_{h_{0}}(\text{Diff}(M)(h_{0})), it can be expressed by the Lie derivative of the metric h0h_{0} in the direction XX. And using Helmholtz-Hodge decomposition, XX decomposes further into X=h0r+YX=\nabla_{h_{0}}r+Y, where rr is a scalar function, and YY is a vector field with divh0Y=0\text{div}_{h_{0}}Y=0. By computation,

(3.15) trh0lD=2Δh0r+2divh0Y=2Δh0r.\text{tr}_{h_{0}}l_{D}=2\Delta_{h_{0}}r+2\text{div}_{h_{0}}Y=2\Delta_{h_{0}}r.

In addition,

(3.16) |lD|L2(M,h0)2=\displaystyle|l_{D}|^{2}_{L^{2}(M,h_{0})}= |2h02r+i,j(iYj+jYi)|L2(M,h0)2\displaystyle|2\nabla^{2}_{h_{0}}r+\sum_{i,j}(\nabla_{i}Y_{j}+\nabla_{j}Y_{i})|^{2}_{L^{2}(M,h_{0})}
=\displaystyle= 4|h02r|L2(M,h0)2+|i,j(iYj+jYi)|L2(M,h0)2\displaystyle 4|\nabla^{2}_{h_{0}}r|^{2}_{L^{2}(M,h_{0})}+|\sum_{i,j}(\nabla_{i}Y_{j}+\nabla_{j}Y_{i})|^{2}_{L^{2}(M,h_{0})}
+4h02r,i,j(iYj+jYi)L2(M,h0).\displaystyle+4\langle\nabla^{2}_{h_{0}}r,\sum_{i,j}(\nabla_{i}Y_{j}+\nabla_{j}Y_{i})\rangle_{L^{2}(M,h_{0})}.

where the second term is in the form

(3.17) |i,j(iYj+jYi)|L2(M,h0)2\displaystyle|\sum_{i,j}(\nabla_{i}Y_{j}+\nabla_{j}Y_{i})|^{2}_{L^{2}(M,h_{0})}
=\displaystyle= M4i(iYi)2+ij(iYj)2+(jYi)2+2iYjjYidVh0\displaystyle\int_{M}4\sum_{i}(\nabla_{i}Y_{i})^{2}+\sum_{i\neq j}(\nabla_{i}Y_{j})^{2}+(\nabla_{j}Y_{i})^{2}+2\nabla_{i}Y_{j}\nabla_{j}Y_{i}\,dV_{h_{0}}
=\displaystyle= M4(divh0Y)2+ij2iYijYj+(iYj)2+(jYi)2\displaystyle\int_{M}4(\text{div}_{h_{0}}Y)^{2}+\sum_{i\neq j}-2\nabla_{i}Y_{i}\nabla_{j}Y_{j}+(\nabla_{i}Y_{j})^{2}+(\nabla_{j}Y_{i})^{2}
2iYijYj+2iYjjYidVh0\displaystyle-2\nabla_{i}Y_{i}\nabla_{j}Y_{j}+2\nabla_{i}Y_{j}\nabla_{j}Y_{i}\,dV_{h_{0}}
=\displaystyle= MijYjjiYi+YiijYj+(iYj)2+(jYi)2\displaystyle\int_{M}\sum_{i\neq j}Y_{j}\nabla_{j}\nabla_{i}Y_{i}+Y_{i}\nabla_{i}\nabla_{j}Y_{j}+(\nabla_{i}Y_{j})^{2}+(\nabla_{j}Y_{i})^{2}
+YiijYjYijiYj+YjjiYiYjijYidVh0\displaystyle+Y_{i}\nabla_{i}\nabla_{j}Y_{j}-Y_{i}\nabla_{j}\nabla_{i}Y_{j}+Y_{j}\nabla_{j}\nabla_{i}Y_{i}-Y_{j}\nabla_{i}\nabla_{j}Y_{i}\,dV_{h_{0}}
=\displaystyle= MijYj(ijYiRijjiYj)+Yi(jiYjRjiijYi)+(iYj)2+(jYi)2\displaystyle\int_{M}\sum_{i\neq j}Y_{j}(\nabla_{i}\nabla_{j}Y_{i}-R_{ijj}^{i}Y_{j})+Y_{i}(\nabla_{j}\nabla_{i}Y_{j}-R_{jii}^{j}Y_{i})+(\nabla_{i}Y_{j})^{2}+(\nabla_{j}Y_{i})^{2}
+YiijYjYi(ijYj+RjiijYi)+YjjiYiYj(jiYi+RijjiYj)dVh0\displaystyle+Y_{i}\nabla_{i}\nabla_{j}Y_{j}-Y_{i}(\nabla_{i}\nabla_{j}Y_{j}+R_{jii}^{j}Y_{i})+Y_{j}\nabla_{j}\nabla_{i}Y_{i}-Y_{j}(\nabla_{j}\nabla_{i}Y_{i}+R_{ijj}^{i}Y_{j})\,dV_{h_{0}}
=\displaystyle= MijYi2+Yj22jYiiYj+(iYj)2+(jYi)2+Yi2+Yj2dVh0\displaystyle\int_{M}\sum_{i\neq j}Y_{i}^{2}+Y_{j}^{2}-2\nabla_{j}Y_{i}\nabla_{i}Y_{j}+(\nabla_{i}Y_{j})^{2}+(\nabla_{j}Y_{i})^{2}+Y_{i}^{2}+Y_{j}^{2}\,dV_{h_{0}}
=\displaystyle= 4(n1)|Y|L2(M,h0)2+|i,j(iYjjYi)|L2(M,h0)2,\displaystyle 4(n-1)|Y|^{2}_{L^{2}(M,h_{0})}+|\sum_{i,j}(\nabla_{i}Y_{j}-\nabla_{j}Y_{i})|^{2}_{L^{2}(M,h_{0})},

and the last term of (3.16) vanishes, since

(3.18) 4h02r,i,j(iYj+jYi)L2(M,h0)\displaystyle 4\langle\nabla^{2}_{h_{0}}r,\sum_{i,j}(\nabla_{i}Y_{j}+\nabla_{j}Y_{i})\rangle_{L^{2}(M,h_{0})}
=\displaystyle= 4M2iiiriYi+ijijr(iYj+jYi)dVh0\displaystyle 4\int_{M}2\sum_{i}\nabla_{i}\nabla_{i}r\nabla_{i}Y_{i}+\sum_{i\neq j}\nabla_{i}\nabla_{j}r(\nabla_{i}Y_{j}+\nabla_{j}Y_{i})\,dV_{h_{0}}
=\displaystyle= 4M2iiiriYiij(irjiYj+jrijYi)dVh0\displaystyle 4\int_{M}2\sum_{i}\nabla_{i}\nabla_{i}r\nabla_{i}Y_{i}-\sum_{i\neq j}(\nabla_{i}r\nabla_{j}\nabla_{i}Y_{j}+\nabla_{j}r\nabla_{i}\nabla_{j}Y_{i})\,dV_{h_{0}}
=\displaystyle= 4M2iiiriYiij(ir(ijYj+RjiijYi)+jr(jiYi+RijjiYj))dVh0\displaystyle 4\int_{M}2\sum_{i}\nabla_{i}\nabla_{i}r\nabla_{i}Y_{i}-\sum_{i\neq j}\big{(}\nabla_{i}r(\nabla_{i}\nabla_{j}Y_{j}+R_{jii}^{j}Y_{i})+\nabla_{j}r(\nabla_{j}\nabla_{i}Y_{i}+R_{ijj}^{i}Y_{j})\big{)}\,dV_{h_{0}}
=\displaystyle= 4M2Δh0rdivh0Yij(iirjYj+jjriYi)\displaystyle 4\int_{M}2\Delta_{h_{0}}r\,\text{div}_{h_{0}}Y-\sum_{i\neq j}(\nabla_{i}\nabla_{i}r\nabla_{j}Y_{j}+\nabla_{j}\nabla_{j}r\nabla_{i}Y_{i})
+ij(iirjYj+jjriYi)+ijirYi+jrYjdVh0\displaystyle+\sum_{i\neq j}(\nabla_{i}\nabla_{i}r\nabla_{j}Y_{j}+\nabla_{j}\nabla_{j}r\nabla_{i}Y_{i})+\sum_{i\neq j}\nabla_{i}rY_{i}+\nabla_{j}rY_{j}\,dV_{h_{0}}
=\displaystyle= 8(n1)h0r,YL2(M,h0)=0.\displaystyle 8(n-1)\langle\nabla_{h_{0}}r,Y\rangle_{L^{2}(M,h_{0})}=0.

Substituting (3.15)-(3.18) into (3.14), then applying Bochner’s formula, we obtain

(3.19) 𝒜′′(h0)(lD,lD)\displaystyle\mathcal{A}^{\prime\prime}(h_{0})(l_{D},l_{D})
=\displaystyle= n22(M4(Δh0r)24|h02r|2dVh0\displaystyle\frac{n-2}{2}\big{(}\int_{M}4(\Delta_{h_{0}}r)^{2}-4|\nabla_{h_{0}}^{2}r|^{2}\,dV_{h_{0}}
4(n1)|Y|L2(M,h0)2|i,j(iYjjYi)|L2(M,h0)2)\displaystyle-4(n-1)|Y|^{2}_{L^{2}(M,h_{0})}-|\sum_{i,j}(\nabla_{i}Y_{j}-\nabla_{j}Y_{i})|^{2}_{L^{2}(M,h_{0})}\big{)}
=\displaystyle= n22(M4Ric(h0r,h0r)𝑑Vh04(n1)|Y|L2(M,h0)2|i,j(iYjjYi)|L2(M,h0)2)\displaystyle\frac{n-2}{2}\big{(}\int_{M}4Ric(\nabla_{h_{0}}r,\nabla_{h_{0}}r)\,dV_{h_{0}}-4(n-1)|Y|^{2}_{L^{2}(M,h_{0})}-|\sum_{i,j}(\nabla_{i}Y_{j}-\nabla_{j}Y_{i})|^{2}_{L^{2}(M,h_{0})}\big{)}
\displaystyle\leq 2(n1)(n2)|X|L2(M,h0)2.\displaystyle-2(n-1)(n-2)|X|^{2}_{L^{2}(M,h_{0})}.

Furthermore, after eliminating some terms using (3.12), (3.13) and trh0lTT=0tr_{h_{0}}l_{TT}=0, we get

(3.20) 𝒜′′(h0)(lD,lTT)=n22MlD,lTTh0𝑑Vh0=0,\mathcal{A}^{\prime\prime}(h_{0})(l_{D},l_{TT})=-\frac{n-2}{2}\int_{M}\langle l_{D},l_{TT}\rangle_{h_{0}}\,dV_{h_{0}}=0,

it vanishes because of the L2L^{2}-orthogonality between Th0(Diff(M)(h0))T_{h_{0}}(\text{Diff}(M)(h_{0})) and TTh0TT_{h_{0}}. Similarly, the traceless-transverse property of lTTl_{TT} simplifies (3.9 to

(3.21) 𝒜′′(h0)(lTT,lD)=n22MlTT,lDh0𝑑Vh0=0.\mathcal{A}^{\prime\prime}(h_{0})(l_{TT},l_{D})=-\frac{n-2}{2}\int_{M}\langle l_{TT},l_{D}\rangle_{h_{0}}\,dV_{h_{0}}=0.

Substituting (3.11), (3.19), (3.20) and (3.21) into (3.10), we complete the proof.

Proof of Lemma 3.3.

Since hRh\in\mathcal{M}_{R}, we have

0=RhRh0=01Rh0+tll𝑑t,0=R_{h}-R_{h_{0}}=\int_{0}^{1}R_{h_{0}+tl}^{\prime}\cdot l\,dt,

it leads to

(3.22) Rh0l,trh0(fh0)L2(M,h0)=01(Rh0+tlRh0)l,trh0(fh0)L2(M,h0)𝑑t.\langle R_{h_{0}}^{\prime}\cdot l,\text{tr}_{h_{0}}(fh_{0})\rangle_{L^{2}(M,h_{0})}=-\int_{0}^{1}\langle(R^{\prime}_{h_{0}+tl}-R^{\prime}_{h_{0}})\cdot l,\text{tr}_{h_{0}}(fh_{0})\rangle_{L^{2}(M,h_{0})}\,dt.

On the left-hand side, we have Rh0l=Rh0(fh0)R_{h_{0}}^{\prime}\cdot l=R_{h_{0}}^{\prime}\cdot(fh_{0}) as analyzed earlier, and it follows from (3.6) that

(3.23) Rh0l,trh0(fh0)h0\displaystyle\langle R_{h_{0}}^{\prime}\cdot l,\text{tr}_{h_{0}}(fh_{0})\rangle_{h_{0}}
=\displaystyle= MΔh0(trh0(fh0))+δh02(fh0)+(n1)trh0(fh0),trh0(fh0)h0𝑑Vh0\displaystyle\int_{M}\langle-\Delta_{h_{0}}(\text{tr}_{h_{0}}(fh_{0}))+\delta^{2}_{h_{0}}(fh_{0})+(n-1)\text{tr}_{h_{0}}(fh_{0}),\text{tr}_{h_{0}}(fh_{0})\rangle_{h_{0}}\,dV_{h_{0}}
=\displaystyle= Mn(n1)|h0f|2+n2(n1)|f|2dVh0\displaystyle\int_{M}n(n-1)|\nabla_{h_{0}}f|^{2}+n^{2}(n-1)|f|^{2}\,dV_{h_{0}}
\displaystyle\geq (n1)|fh0|H1(M,h0)2.\displaystyle(n-1)|fh_{0}|^{2}_{H^{1}(M,h_{0})}.

On the other hand, the right-hand side of (3.22) can be estimated using the continuity of htrhlh\rightarrow\text{tr}_{h}l, h,L2(M,h)h\rightarrow\langle\cdot,\cdot\rangle_{L^{2}(M,h)}, hhlh\rightarrow\nabla_{h}l, and hδhlh\rightarrow\delta_{h}l. For any ϵ>0\epsilon>0, after shrinking the neighborhood of h0h_{0} in R\mathcal{M}_{R}, we have

(3.24) 01(Rh0+tlRh0)l,trh0(fh0)L2(M,h0)𝑑tϵ|hh0|C2|l|H1(M,h0)|fh0|H1(M,h0).-\int_{0}^{1}\langle(R^{\prime}_{h_{0}+tl}-R^{\prime}_{h_{0}})\cdot l,\text{tr}_{h_{0}}(fh_{0})\rangle_{L^{2}(M,h_{0})}\,dt\leq\epsilon|h-h_{0}|_{C^{2}}|l|_{H^{1}(M,h_{0})}|fh_{0}|_{H^{1}(M,h_{0})}.

Therefore, taken (3.23) and (3.24) into account, (3.22) leads to the result.

To end this section, we discuss the equality condition of Theorem 1.2. There exists t(0,1)t\in(0,1), such that

𝒜(h)=𝒜(h0)+𝒜(h0)′′(l,l)2+𝒜′′′(h0+tl)(l,l,l)6.\mathcal{A}(h)=\mathcal{A}(h_{0})+\frac{\mathcal{A}(h_{0})^{\prime\prime}(l,l)}{2}+\frac{\mathcal{A}^{\prime\prime\prime}(h_{0}+tl)(l,l,l)}{6}.

Following the same procedure to compute 𝒜′′′(h0)\mathcal{A}^{\prime\prime\prime}(h_{0}) and discuss the continuity near h0h_{0}, we can see that 𝒜′′′(h0+tl)(l,l,l)=O(|l|H1(M,h0)3)\mathcal{A}^{\prime\prime\prime}(h_{0}+tl)(l,l,l)=O(|l|^{3}_{H^{1}(M,h_{0})}). Thus, from the above lemmas, there exists C>0C^{\prime}>0,

𝒜(h)𝒜(h0)C(|fh0+lTT|H1(M,h0)2+|X|L2(M,h0)2)+O(|hh0|H1(M,h0)3).\mathcal{A}(h)\leq\mathcal{A}(h_{0})-C^{\prime}(|fh_{0}+l_{TT}|^{2}_{H^{1}(M,h_{0})}+|X|^{2}_{L^{2}(M,h_{0})})+O(|h-h_{0}|^{3}_{H^{1}(M,h_{0})}).

From the computation of (3.19) and its derivative, we can see that as hh0h\rightarrow h_{0}, |X|L2(M,h0)|X|_{L^{2}(M,h_{0})} and |lD|H1(M,h0)|l_{D}|_{H^{1}(M,h_{0})} decay at the same rate, so the norm |fh0+lTT|H1(M,h0)2+|X|L2(M,h0)2C′′|hh0|H1(M,h0)2|fh_{0}+l_{TT}|^{2}_{H^{1}(M,h_{0})}+|X|^{2}_{L^{2}(M,h_{0})}\sim C^{\prime\prime}|h-h_{0}|^{2}_{H^{1}(M,h_{0})}. Consequently, the following expansion holds for metric hh in a small neighborhood of h0h_{0} with Rhn(n1)R_{h}\geq-n(n-1).

AreaId(h/h0)1+C′′′(|fh0+lTT|H1(M,h0)2+|X|L2(M,h0)2)+O(|hh0|H1(M,h0)3).\text{Area}_{\text{Id}}(h/h_{0})\geq 1+C^{\prime\prime\prime}(|fh_{0}+l_{TT}|^{2}_{H^{1}(M,h_{0})}+|X|^{2}_{L^{2}(M,h_{0})})+O(|h-h_{0}|^{3}_{H^{1}(M,h_{0})}).

where C′′′>0C^{\prime\prime\prime}>0. Note that AreaId(h/h0)=1\text{Area}_{\text{Id}}(h/h_{0})=1 requires that fh0=lTT=X=0fh_{0}=l_{TT}=X=0, and thus h=h0h=h_{0}.

4. Equidistribution property and average area ratio formula

In this section, we extend the notations of laminations and associated properties for hyperbolic 3-manifolds (see Labourie [17] and Lowe-Neves [19]) to the higher, odd-dimensional case. The purpose of this section is to deduce the average area ratio formula in Lemma 4.5, as an important tool in the proofs of Theorem 1.4 and Theorem 1.6.

4.1. Laminations and laminar measures

Let M=n/π1(M)M=\mathbb{H}^{n}/\pi_{1}(M) be a closed hyperbolic manifold of dimension n3n\geq 3. And let (n,ϵ)\mathcal{F}(\mathbb{H}^{n},\epsilon) be the space of conformal minimal immersions Φ:2n\Phi:\mathbb{H}^{2}\rightarrow\mathbb{H}^{n}, such that Φ(2)\Phi(\partial_{\infty}\mathbb{H}^{2}) is an (1+ϵ)(1+\epsilon)-quasicircle. As discussed in Section 2 of [13], when ϵ\epsilon is sufficiently small, Φ(2)\Phi(\mathbb{H}^{2}) is a stable embedded disc in n\mathbb{H}^{n}. The space (n,ϵ)\mathcal{F}(\mathbb{H}^{n},\epsilon) equips with the topology of uniform convergence on compact sets, and we take

(M,ϵ):=(n,ϵ)/π1(M)\mathcal{F}(M,\epsilon):=\mathcal{F}(\mathbb{H}^{n},\epsilon)/\pi_{1}(M)

with the quotient topology. The space (M,ϵ)\mathcal{F}(M,\epsilon) together with the action of PSL(2,)\text{PSL}(2,\mathbb{R}) by pre-composition

(4.1) γ:(M,ϵ)(M,ϵ),γ(ϕ)=ϕγ1,γPSL(2,)\mathcal{R}_{\gamma}:\mathcal{F}(M,\epsilon)\rightarrow\mathcal{F}(M,\epsilon),\quad\mathcal{R}_{\gamma}(\phi)=\phi\circ\gamma^{-1},\quad\forall\gamma\in\text{PSL}(2,\mathbb{R})

is called the conformal minimal lamination of MM. A laminar measure on (M,ϵ)\mathcal{F}(M,\epsilon) stands for a probability measure which is invariant under the PSL(2,)\text{PSL}(2,\mathbb{R})-action defined as above. The space (M,ϵ)\mathcal{F}(M,\epsilon) is sequentially compact, but the space of laminar measures is not necessarily weakly compact. In light of that, we consider a continuous map from (M,ϵ)\mathcal{F}(M,\epsilon) to the frame bundle F(M)F(M) of MM, the latter space is compact, so the space of probability measures on F(M)F(M) is compact in weak-* topology.

Firstly we define a map from (M,ϵ)\mathcal{F}(M,\epsilon) to the 2-vector bundle F2(M)F_{2}(M) on MM consisting of (x,v1,v2)M×SxM×SxM(x,v_{1},v_{2})\in M\times S_{x}M\times S_{x}M, where SxMS_{x}M denotes the unit sphere in the tangent space to MM at xx. Let {e1,e2}\{e_{1},e_{2}\} be an orthonormal basis of 2\mathbb{H}^{2}, and for any ϕ(M,ϵ)\phi\in\mathcal{F}(M,\epsilon), let ϕ(h0)=Cϕ2h2\phi^{*}(h_{0})=C_{\phi}^{2}h_{\mathbb{H}^{2}}, where Cϕ2>0C_{\phi}^{2}>0 denotes the conformal factor between the hyperbolic metric on 2\mathbb{H}^{2} and the pull-back metric of h0h_{0} by ϕ\phi. Let e1(ϕ)=dϕ(e1)Cϕe_{1}(\phi)=\dfrac{d\phi(e_{1})}{C_{\phi}} and e2(ϕ)=dϕ(e2)Cϕe_{2}(\phi)=\dfrac{d\phi(e_{2})}{C_{\phi}}. We define the following continuous map.

Ω:(M,ϵ)F2(M),Ω(ϕ)=(ϕ(i),e1(ϕ),e2(ϕ)).\Omega:\mathcal{F}(M,\epsilon)\rightarrow F_{2}(M),\quad\Omega(\phi)=(\phi(i),e_{1}(\phi),e_{2}(\phi)).

Furthermore, it induces a map from (M,ϵ)\mathcal{F}(M,\epsilon) to the frame bundle F(M)F(M) by parallel transport:

Ω¯:(M,ϵ)F(M),Ω¯(ϕ)=(ϕ(i),e1(ϕ),e2(ϕ),,en(ϕ)).\overline{\Omega}:\mathcal{F}(M,\epsilon)\rightarrow F(M),\quad\overline{\Omega}(\phi)=(\phi(i),e_{1}(\phi),e_{2}(\phi),\cdots,e_{n}(\phi)).

And define the projection

P:F(M)F2(M),P(x,e1,e2,,en)=(x,e1,e2).P:F(M)\rightarrow F_{2}(M),\quad P(x,e_{1},e_{2},\cdots,e_{n})=(x,e_{1},e_{2}).

We consider the subspace (n,0)(n,ϵ)\mathcal{F}(\mathbb{H}^{n},0)\subset\mathcal{F}(\mathbb{H}^{n},\epsilon), it contains isometric immersions ϕ0:2n\phi_{0}:\mathbb{H}^{2}\rightarrow\mathbb{H}^{n} whose images are totally geodesic discs in n\mathbb{H}^{n}. Conversely, each totally geodesic disc is uniquely determined by ϕ(i)\phi(i), and tangent vectors e1(ϕ)e_{1}(\phi), e2(ϕ)e_{2}(\phi). Let Ω0:(M,0)F2(M)\Omega_{0}:\mathcal{F}(M,0)\rightarrow F_{2}(M) be the restriction of Ω\Omega to (M,0)\mathcal{F}(M,0), it’s therefore a bijection. Using (4.1), We can define the PSL(2,)\text{PSL}(2,\mathbb{R})-action on F(M)F(M) as follows.

Rγ:F(M)F(M),Rγ(x)=Ω¯γΩ01P(x),γPSL(2,).R_{\gamma}:F(M)\rightarrow F(M),\quad R_{\gamma}(x)=\overline{\Omega}\circ\mathcal{R}_{\gamma}\circ\Omega_{0}^{-1}\circ P(x),\quad\forall\gamma\in\text{PSL}(2,\mathbb{R}).

This definition coincides with the homogeneous action of PSL(2,)\text{PSL}(2,\mathbb{R}) on F(M)F(M). Following the discussion of Lemma 3.2 of [19], we conclude the following result.

Proposition 4.1.

Given any sequence of laminar measures μi\mu_{i} on (M,1i)\mathcal{F}(M,\frac{1}{i}), the sequence of induced measures Ω¯μi\overline{\Omega}_{*}\mu_{i} on F(M)F(M) converges weakly to a probability measure ν\nu, then ν\nu is invariant under the homogeneous action of PSL(2,)\text{PSL}(2,\mathbb{R}).

Let G<PSL(2,)G<\text{PSL}(2,\mathbb{R}) be a Fuchsian subgroup, then 2/G\mathbb{H}^{2}/G is a closed hyperbolic surface with genus 2\geq 2 whose fundamental domain is represented by UU. And let ϕ(M,ϵ)\phi\in\mathcal{F}(M,\epsilon) equivariant with respect to a representation Π\Pi of GG in π1(M)<SO(n,1)\pi_{1}(M)<SO(n,1). The image of ϕ(2)\phi(\mathbb{H}^{2}) in MM is closed minimal surface in MM whose fundamental group is Π\Pi. We define a laminar measure associated with ϕ\phi as follows.

(4.2) δϕ(f)=1vol(U)Uf(ϕγ)𝑑ν0(γ),fC0((M,ϵ)),\delta_{\phi}(f)=\frac{1}{vol(U)}\int_{U}f(\phi\circ\gamma)d\nu_{0}(\gamma),\quad\forall f\in C^{0}(\mathcal{F}(M,\epsilon)),

where ν0\nu_{0} denotes the bi-invariant measure on PSL(2,)\text{PSL}(2,\mathbb{R}).

4.2. Equidistribution

In this section, we assume the dimension of MM is odd. Adapting the methods of Proposition 6.1 of [19] and Theorem 5.7 in [17], we prove the following result.

Proposition 4.2.

For any ii\in\mathbb{N}, there is a lamination ϕi\phi_{i} in (M,1i)\mathcal{F}(M,\frac{1}{i}) equivariant with respect to a representation of a Fuchsian group Gi<PSL(2,)G_{i}<\text{PSL}(2,\mathbb{R}) in π1(M)\pi_{1}(M), such that Ω¯δϕi\overline{\Omega}_{*}\delta_{\phi_{i}} converges to the Lebesgue measure μLeb\mu_{Leb} on F(M)F(M) as ii\rightarrow\infty.

Sketch of the proof.

Let 𝒯~\tilde{\mathcal{T}} be the space of tripods X~=(x1,x2,x3)\tilde{X}=(x_{1},x_{2},x_{3}), where x1,x2,x3nx_{1},x_{2},x_{3}\in\partial_{\infty}\mathbb{H}^{n}. Each element X~\tilde{X} determines an ideal triangle Δ(X~)\Delta(\tilde{X}) in n\mathbb{H}^{n}. Let b(X~)b(\tilde{X}) be the barycenter of Δ(X~)\Delta(\tilde{X}). Denote by (e1(X~),e2(X~))(e_{1}(\tilde{X}),e_{2}(\tilde{X})) the orthonormal basis of Δ(X~)\Delta(\tilde{X}), and denote by (e3(X~,,en(X~))(e_{3}(\tilde{X},\cdots,e_{n}(\tilde{X})) the orthonormal basis of the normal bundle of Δ(X~)\Delta(\tilde{X}) in n\mathbb{H}^{n}. Thus, each tripod X~\tilde{X} determines a point (b(X~),e1(X~),,en(X~))(b(\tilde{X}),e_{1}(\tilde{X}),\cdots,e_{n}(\tilde{X})) in F(n)F(\mathbb{H}^{n}), which represents the frame bundle of n\mathbb{H}^{n}.

Consider the closed manifold M=n/π1(M)M=\mathbb{H}^{n}/\pi_{1}(M). Let XX be the corresponding point of X~\tilde{X} in 𝒯:=𝒯~/π1(M)\mathcal{T}:=\tilde{\mathcal{T}}/\pi_{1}(M), and let F(X)F(X) be the point in the frame bundle F(M)F(M) corresponding to (b(X~),e1(X~),,en(X~))(b(\tilde{X}),e_{1}(\tilde{X}),\cdots,e_{n}(\tilde{X})) in F(n)F(\mathbb{H}^{n}).

A quintuple (X,Y,l1,l2,l3)(X,Y,l_{1},l_{2},l_{3}) is called a triconnected pair of tripods (Definition 10.1.1 of [14]) if X,Y𝒯X,Y\in\mathcal{T} and l1,l2,l3l_{1},l_{2},l_{3} are three distinct homotopy classes of paths connecting XX to YY. The space of triconnected pair of tripods is denoted by 𝒯𝒯\mathcal{TT}. Let π1\pi^{1} and π2\pi^{2} be the forgetting maps from 𝒯𝒯\mathcal{TT} to 𝒯\mathcal{T}

π1:(X,Y,l1,l2,l3)X,π2:(X,Y,l1,l2,l3)Y.\pi^{1}:(X,Y,l_{1},l_{2},l_{3})\mapsto X,\quad\pi^{2}:(X,Y,l_{1},l_{2},l_{3})\mapsto Y.

Moreover, let π1¯\overline{\pi^{1}} and π2¯\overline{\pi^{2}} be the corresponding maps from 𝒯𝒯\mathcal{TT} to F(M)F(M)

π1¯:(X,Y,l1,l2,l3)F(X),π2¯:(X,Y,l1,l2,l3)F(Y).\overline{\pi^{1}}:(X,Y,l_{1},l_{2},l_{3})\mapsto F(X),\quad\overline{\pi^{2}}:(X,Y,l_{1},l_{2},l_{3})\mapsto F(Y).

In addition, there is a weighted measure μϵ,R\mu_{\epsilon,R} on 𝒯𝒯\mathcal{TT} as defined in Definition 11.2.3 of [14]. If (X,Y,l1,l2,l3)(X,Y,l_{1},l_{2},l_{3}) is in the support of μϵ,R\mu_{\epsilon,R}, then the ideal triangles determined by XX and YY can be glued to an (ϵ,R)(\epsilon,R)-almost closing pair of pants (see Definition 9.1.1 of [14]). Moreover, it follows from the mixing property that for fixed ϵ\epsilon, as RR\rightarrow\infty, π1¯μϵ,R\overline{\pi^{1}}_{*}\mu_{\epsilon,R} and π2¯μϵ,R\overline{\pi^{2}}_{*}\mu_{\epsilon,R} both converge to the Lebesgue measure μLeb\mu_{Leb} on F(M)F(M).

Arguing like Theorem 5.7 of [17], we can choose a sequence RjR_{j}\rightarrow\infty as jj\rightarrow\infty, and a sequence of measures μ1j,Rj\mu_{\frac{1}{j},R_{j}}. Then we approximate each μ1j,Rj\mu_{\frac{1}{j},R_{j}} by another weighted measure νj\nu_{j} supported in finitely many pleated pair of pants Pj1,,PjNjP^{1}_{j},\cdots,P^{N_{j}}_{j}, which can be glued together to get essential surfaces Σj1,,ΣjMj\Sigma_{j}^{1},\cdots,\Sigma_{j}^{M_{j}} in MM. When jj is sufficiently large and for each 1kMj1\leq k\leq M_{j}, Σjk\Sigma_{j}^{k} is (1+1j)(1+\frac{1}{j})-quasigeodesic, and the projection from Σjk\Sigma_{j}^{k} to the unique minimal surface SjkS_{j}^{k} homotopic to Σjk\Sigma_{j}^{k} is (1+1j)(1+\frac{1}{j})-bi-Lipschitz and it has distance uniformly bounded by O(1j)O(\frac{1}{j}). For this reason, we can further approximate νj\nu_{j} by a weighted measure supported in Sj1SjMjS_{j}^{1}\cup\cdots\cup S_{j}^{M_{j}}. SjkS_{j}^{k} is obtained by a lamination ϕjk(M,1j)\phi_{j}^{k}\in\mathcal{F}(M,\frac{1}{j}), in fact, it’s the image of ϕjk(2)\phi_{j}^{k}(\mathbb{H}^{2}) in MM, and thus associated with the laminar measure δϕjk\delta_{\phi_{j}^{k}}, we have the following lemma.

Lemma 4.3.

For any jj\in\mathbb{N}, there exist a finite sequence of laminations ϕj1,,ϕjMj\phi_{j}^{1},\cdots,\phi_{j}^{M_{j}} in (M,1j)\mathcal{F}(M,\frac{1}{j}), and θj1,,θjMj(0,1)\theta_{j}^{1},\cdots,\theta_{j}^{M_{j}}\in(0,1) with θj1++θjMj=1\theta_{j}^{1}+\cdots+\theta_{j}^{M_{j}}=1, such that each ϕjk\phi_{j}^{k} is equivariant with respect to a representation of a Fuchsian group in π1(M)\pi_{1}(M), and the laminar measure

μj=k=1Mjθjkδϕjk\mu_{j}=\sum_{k=1}^{M_{j}}\theta_{j}^{k}\delta_{\phi_{j}^{k}}

satisfies that Ω¯μj\overline{\Omega}_{*}\mu_{j} converges to the Lebesgue measure μLeb\mu_{Leb} on F(M)F(M) as jj\rightarrow\infty.

Next, for 2ln12\leq l\leq n-1, we define

𝒫l:={F(P)F(M),\displaystyle\mathcal{P}_{l}:=\{F(P)\subset F(M), where P is a l-dimensional closed totally geodesic\displaystyle\text{ where }P\text{ is a $l$-dimensional closed totally geodesic}
submanifold of M}.\displaystyle\text{ submanifold of }M\}.

Then 𝒫:l=2n1𝒫l\mathcal{P}:\underset{l=2}{\overset{n-1}{\cup}}\mathcal{P}_{l} contains at most countably many candidates. Therefore, we can find a decreasing sequence of tubular neighborhoods {Bk}F(M)\{B_{k}\}\subset F(M), so that for any kk\in\mathbb{N}, BkB_{k} covers 𝒫\mathcal{P} and it satisfies μLeb(Bk)<22k1\mu_{Leb}(B^{k})<2^{-2k-1} and μLeb(Bk)=0\mu_{Leb}(\partial B_{k})=0. In consequence of previous lemma, after passing to a subsequence, we have Ω¯μj(Bk)<22k\overline{\Omega}_{*}\mu_{j}(B_{k})<2^{-2k}. Additionally, as argued in Lemma 6.2 of [19], we can find a subsequence {ji}\{j_{i}\}, and ϕi{ϕji1,,ϕjiMji}\phi_{i}\in\{\phi_{j_{i}}^{1},\cdots,\phi_{j_{i}}^{M_{j_{i}}}\}, such that Ω¯(δϕi)(Bk)<2k\overline{\Omega}_{*}(\delta_{\phi_{i}})(B_{k})<2^{-k}.

As a result of Proposition 4.1, as ii\rightarrow\infty, Ω¯δϕi\overline{\Omega}_{*}\delta_{\phi_{i}} converges weakly to a probability measure ν\nu on F(M)F(M). ν\nu is invariant under the homogeneous action of PSL(2,)\text{PSL}(2,\mathbb{R}), and it satisfies that

(4.3) ν(Bk)<2k.\nu(B_{k})<2^{-k}.

To finish the proof, we need the following lemma.

Lemma 4.4.

ν=μLeb\nu=\mu_{Leb}.

Proof.

According to the ergodic decomposition theorem ([11]), ν\nu can be expressed by a linear combination of the ergodic measures for PSL(2,)\text{PSL}(2,\mathbb{R})-action on F(M)F(M). Moreover, Ratner’s measure classification theorem (see [20] or [22]) says that any ergodic PSL(2,)\text{PSL}(2,\mathbb{R})-invariant measure on F(M)F(M) is either an invariant probability measure supported on a finite union of {Pk}𝒫\{P_{k}\}\subset\mathcal{P}, or it is identical to μLeb\mu_{Leb}. Thus, we can write ν\nu as

ν=a1μLeb+a2μ𝒫2++an1μ𝒫n1,\nu=a_{1}\mu_{Leb}+a_{2}\mu_{\mathcal{P}_{2}}+\cdots+a_{n-1}\mu_{\mathcal{P}_{n-1}},

where a1+a2++an1=1a_{1}+a_{2}+\cdots+a_{n-1}=1 and μ𝒫l\mu_{\mathcal{P}_{l}} represents an ergodic measure supported on 𝒫l\mathcal{P}_{l}, 2ln12\leq l\leq n-1. By (4.3), for all kk\in\mathbb{N},

a2++an1=a2μ𝒫2(Bk)++an1μ𝒫n1(Bk)ν(Bk)<2k.a_{2}+\cdots+a_{n-1}=a_{2}\mu_{\mathcal{P}_{2}}(B_{k})+\cdots+a_{n-1}\mu_{\mathcal{P}_{n-1}}(B_{k})\leq\nu(B_{k})<2^{-k}.

So

a1=1a2an1>12k,k.a_{1}=1-a_{2}-\cdots-a_{n-1}>1-2^{-k},\quad\forall k\in\mathbb{N}.

We must have a1=1a_{1}=1, and therefore ν=μLeb\nu=\mu_{Leb}. ∎

Proposition 4.2 follows immediately from the lemma. ∎

4.3. Average area ratio formula

Lemma 4.5 (average area ratio formula).

Let (N,g)(N,g) be a closed Riemannian manifold that also has odd dimension nn, and let FF be a smooth map that takes (N,g)(N,g) to (M,h0)(M,h_{0}). For ii\in\mathbb{N}, we pick a lamination ϕi(M,1i)\phi_{i}\in\mathcal{F}(M,\frac{1}{i}) equivariant with respect to a representation of Gi<PSL(2,)G_{i}<\text{PSL}(2,\mathbb{R}) in π1(M)\pi_{1}(M), and it satisfies Proposition 4.2. Let SiS_{i} be the image of ϕi(2)\phi_{i}(\mathbb{H}^{2}) in MM. Then we have

AreaF(g/h0)=limiareag(F1(Si))4π(gi1).\text{Area}_{F}(g/h_{0})=\lim_{i\rightarrow\infty}\frac{\text{area}_{g}(F^{-1}(S_{i}))}{4\pi(g_{i}-1)}.
Proof.

Recall that |Λ2F|g1|\Lambda^{2}F|^{-1}_{g} is a function defined almost everywhere on Gr2(M)Gr_{2}(M). Since |Λ2F|g|\Lambda^{2}F|_{g} can be regarded as a smooth function on F(M)F(M) by

|Λ2F|g:F(M),(x,e1,e2,en)|Λ2F|g(x,span(e1,e2)),|\Lambda^{2}F|_{g}:F(M)\rightarrow\mathbb{R},\quad(x,e_{1},e_{2}\cdots,e_{n})\mapsto|\Lambda^{2}F|_{g}(x,\text{span}(e_{1},e_{2})),

based on the definition (2.1), |Λ2F|g1|\Lambda^{2}F|^{-1}_{g} is also seen as a function defined almost everywhere on F(M)F(M). Thus, Proposition 4.2 implies that

AreaF(g/h0)=μLeb(|Λ2F|g1)=limiΩ¯δϕi(|Λ2F|g1).\text{Area}_{F}(g/h_{0})=\mu_{Leb}(|\Lambda^{2}F|^{-1}_{g})=\lim_{i\rightarrow\infty}\overline{\Omega}_{*}\delta_{\phi_{i}}(|\Lambda^{2}F|^{-1}_{g}).

In light of the definition of laminar measure δϕi\delta_{\phi_{i}} in (4.2), we have

Ω¯δϕi(|Λ2F|g1)=1vol(Ui)Ui|Λ2F|g1Ω¯(ϕiγ)𝑑ν0(γ),\overline{\Omega}_{*}\delta_{\phi_{i}}(|\Lambda^{2}F|^{-1}_{g})=\frac{1}{vol(U_{i})}\int_{U_{i}}|\Lambda^{2}F|^{-1}_{g}\circ\overline{\Omega}(\phi_{i}\circ\gamma)d\nu_{0}(\gamma),

where UiU_{i} is the fundamental domain of PSL(2,)/Gi\text{PSL}(2,\mathbb{R})/G_{i}. Set x=γ(i)x=\gamma(i). Since the hyperbolic surface 2/Gi\mathbb{H}^{2}/G_{i} has area equal to 4π(gi1)4\pi(g_{i}-1), where gi2g_{i}\geq 2 denotes the genus. The above expression also can be written as

Ω¯δϕi(|Λ2F|g1)\displaystyle\overline{\Omega}_{*}\delta_{\phi_{i}}(|\Lambda^{2}F|^{-1}_{g}) =14π(gi1)2/Gi|Λ2F|g1(ϕi(x),(dϕi)xTx2)𝑑Ah2/Gi(x)\displaystyle=\frac{1}{4\pi(g_{i}-1)}\int_{\mathbb{H}^{2}/G_{i}}|\Lambda^{2}F|^{-1}_{g}(\phi_{i}(x),(d\phi_{i})_{x}T_{x}\mathbb{H}^{2})dA_{h_{\mathbb{H}^{2}/G_{i}}}(x)
=14π(gi1)Si|Λ2F|g1(y,TySi)Ci2(ϕi1(y))𝑑Ah0(y),\displaystyle=\frac{1}{4\pi(g_{i}-1)}\int_{S_{i}}\frac{|\Lambda^{2}F|^{-1}_{g}(y,T_{y}S_{i})}{C_{i}^{2}(\phi_{i}^{-1}(y))}dA_{h_{0}}(y),

where Ci2>0C_{i}^{2}>0 denotes the conformal factor between the hyperbolic metric on 2/Gi\mathbb{H}^{2}/G_{i} and the pull-back metric of h0h_{0} by ϕi\phi_{i}, namely ϕi(h0)=Ci2h2/Gi\phi_{i}^{*}(h_{0})=C_{i}^{2}h_{\mathbb{H}^{2}/G_{i}}. Since the Gaussian curvature on SiS_{i} has the form 112|A|2(x)-1-\frac{1}{2}|A|^{2}(x), we have

11Ci21+12|A|L(Si)2.1\leq\frac{1}{C_{i}^{2}}\leq 1+\frac{1}{2}|A|^{2}_{L^{\infty}(S_{i})}.

On the other hand, the co-area formula yields that

Si|Λ2F|g1(y,TySi)𝑑Ah0(y)=areag(F1(Si)).\int_{S_{i}}|\Lambda^{2}F|^{-1}_{g}(y,T_{y}S_{i})dA_{h_{0}}(y)=\text{area}_{g}(F^{-1}(S_{i})).

Combining these formulas, we have

areag(F1(Si))4π(gi1)Ω¯δϕi(|Λ2F|g1)(1+12|A|L(Si)2)areag(F1(Si))4π(gi1).\frac{\text{area}_{g}(F^{-1}(S_{i}))}{4\pi(g_{i}-1)}\leq\overline{\Omega}_{*}\delta_{\phi_{i}}(|\Lambda^{2}F|^{-1}_{g})\leq(1+\frac{1}{2}|A|^{2}_{L^{\infty}(S_{i})})\frac{\text{area}_{g}(F^{-1}(S_{i}))}{4\pi(g_{i}-1)}.

Since |A|L(Si)20|A|^{2}_{L^{\infty}(S_{i})}\rightarrow 0 as ii\rightarrow\infty (see [13]), the lemma follows immediately from the squeeze theorem. ∎

5. Gromov’s conjecture in dimension three

In this section, we discuss the proof of Theorem 1.4. First of all, the fact that degF=1\text{deg}\,F=1 follows from Corollary 0.3 of [25] and the geometrization theorem for 3-manifolds. Next, it’s easy to see, the induced map F:π1(N)π1(M)F_{*}:\pi_{1}(N)\rightarrow\pi_{1}(M) is surjective, since otherwise, it factors through a dd-fold covering space of MM with d>1d>1, and thus degFd>1\text{deg}\,F\geq d>1, violating the degree one observation. In addition, π1(M)\pi_{1}(M) is a Hopfian group (for example, see 15.13 of [12]), so the surjectivity of FF_{*} can be upgraded to be an isomorphism, which makes FF a homotopy equivalence between NN and MM due to Whitehead’s theorem. Furthermore, the Mostow rigidity theorem indicates that FF is homotopic to an isometry. For this reason, we can simplify the conditions of Theorem 1.4 as follows.

5.1. A simpler version of Theorem 1.4

Theorem 1.4’.

Let (M,h0)(M,h_{0}) be a closed hyperbolic 3-manifold. There exists a small neighborhood 𝒰\mathcal{U} of h0h_{0} in the C2C^{2}-topology, such that for all Riemannian metric h𝒰h\in\mathcal{U} with Rh6R_{h}\geq-6, and for any smooth map F:(M,h)(M,h0)F:(M,h)\rightarrow(M,h_{0}), it has degF=1\text{deg}\,F=1 and it is homotopic to the identity, we have

AreaF(h/h0)1.\quad\text{Area}_{F}(h/h_{0})\geq 1.

Moreover, the equality holds if and only if FF is an isometry between hh and h0h_{0}.

Let SiS_{i} be the minimal surface in MM with respect to h0h_{0} defined in Lemma 4.5. The inverse F1(Si)F^{-1}(S_{i}) is also a closed surface in MM, but note that F1(Si)F^{-1}(S_{i}) is not necessarily homotopic to SiS_{i}. In fact, we can only find the following relation of their genus. The Gromov norms of SiS_{i} and F1(Si)F^{-1}(S_{i}) satisfies that

|degF|SiF1(Si).|\deg F|||S_{i}||\leq||F^{-1}(S_{i})||.

Here degF=1\deg F=1. And for any closed surface SS with genus g(S)g(S), S=4π(g(S)1)v2||S||=\frac{4\pi(g(S)-1)}{v_{2}}, where v2v_{2} is a fixed number representing the supreme area of geodesic 2-simplices in 2\mathbb{H}^{2}. As an immediate result, we have g(F1(Si))g(Si)g(F^{-1}(S_{i}))\geq g(S_{i}).

To compare the areas of surfaces with respect to the induced metric of hh in different homotopic classes, we hope to find a global area-minimizing surface. In general, the existence and the topology of such a surface are complicated. But if there is a minimal surface with suitable curvature conditions, then adapting Uhlenbeck’s method in [24], we can check the uniqueness of a closed minimal surface of any type, which is the key point of the proof.

5.2. Proof of Theorem 1.4’.

For each ii\in\mathbb{N}, let M~i\tilde{M}_{i} be the covering space of MM such that π1(M~i)π1(Si)\pi_{1}(\tilde{M}_{i})\cong\pi_{1}(S_{i}). Let F~i\tilde{F}_{i} be the corresponding lift of FF that maps F~i1(M~i)M~i\tilde{F}_{i}^{-1}(\tilde{M}_{i})\simeq\tilde{M}_{i} to M~i\tilde{M}_{i}. The lift of SiS_{i} in M~i\tilde{M}_{i} still has fundamental group π1(Si)\pi_{1}(S_{i}), so we denote it by SiS_{i} as well. By assumption, FF is homotopic to identity, thus there is a continuous map H:M×[0,1]MH:M\times[0,1]\rightarrow M with H(x,0)=xH(x,0)=x and H(x,1)=F(x)H(x,1)=F(x) for any xMx\in M. Since MM is compact, the length of the path of HH between xx and F(x)F(x) is uniformly bounded by a constant C>0C>0. Now let H~i\tilde{H}_{i} be the lift of HH that connects F~i\tilde{F}_{i} to the identity map on M~i\tilde{M}_{i}. For all yM~iy\in\tilde{M}_{i}, the length of the path between yy and F~i(y)\tilde{F}_{i}(y) is therefore uniformly bounded by the same constant CC. So F~i\tilde{F}_{i} is proper, meaning F~i1(Si)\tilde{F}^{-1}_{i}(S_{i}) is a closed set, and therefore it is a kk-fold cover of F1(Si)F^{-1}(S_{i}) for some finite number kk. If k>1k>1, the image of F~i1(Si)\tilde{F}^{-1}_{i}(S_{i}) under F~i\tilde{F}_{i} is either a closed surface with Euler characteristic equal to kχ(Si)k\chi(S_{i}), and therefore having genus kg(Si)k+1>g(Si)kg(S_{i})-k+1>g(S_{i}), or it is a union of at least two surfaces with genus g(Si)\geq g(S_{i}). However, both cases are impossible because the image cannot be identified with SiM~iS_{i}\subset\tilde{M}_{i}. We must have k=1k=1. Consequently, the covering map from F~i1(Si)\tilde{F}_{i}^{-1}(S_{i}) to F1(Si)F^{-1}(S_{i}) is one-to-one.

On the other hand, the classical result [21] verifies the existence of area-minimizing surface Σi(M,h)\Sigma_{i}\subset(M,h) in the homotopy class of SiS_{i}. And based on Theorem 4.3 of [18], there exist a C2C^{2}-neighborhood 𝒰0\mathcal{U}_{0} of h0h_{0} and N0N_{0}\in\mathbb{N}, so that when h𝒰0h\in\mathcal{U}_{0} and iN0i\geq N_{0}, Σi\Sigma_{i} is the unique minimal surface in (M,h)(M,h) homotopic to SiS_{i}. Furthermore, let DiD_{i} (Ωi\Omega_{i}) be the lifts of SiS_{i} (Σi\Sigma_{i}, respectively) in B3B^{3}. These discs DiD_{i} and Ωi\Omega_{i} are asymptotic and at a uniformly bounded Hausdorff distance to each other, as hh0h\rightarrow h_{0}, Ωi\Omega_{i} converges uniformly on compact sets to DiD_{i} in C2,αC^{2,\alpha}. Therefore, replacing 𝒰0\mathcal{U}_{0} by a smaller subset or replacing N0N_{0} by a larger integer if necessary, we can assume that if h𝒰0h\in\mathcal{U}_{0} and iN0i\geq N_{0}, then there exists a smooth map fif_{i} on DiD_{i} with |fi|C2,α<1|f_{i}|_{C^{2,\alpha}}<1, such that Ωi\Omega_{i} can be represented by a graph of fif_{i} over DiD_{i}. More precisely, Let nin_{i} be the unit normal vector field of DiD_{i}, then we have the following diffeomorphism.

Fi:DiΩi,Fi(x)=cosh(fi(x))x+sinh(fi(x))ni(x).F_{i}:D_{i}\rightarrow\Omega_{i},\quad F_{i}(x)=\cosh(f_{i}(x))x+\sinh(f_{i}(x))n_{i}(x).

Notice that the minimal disc Ωi\Omega_{i} has mean curvature equal to zero with respect to hh, so the mean curvature Hh0(Ωi)H_{h_{0}}(\Omega_{i}) with respect to h0h_{0} has a uniform bound determined by the perturbation of hh and h\nabla h. Since hh is C2C^{2}-close to h0h_{0}, we have

(5.1) |Hh0(Ωi)|C0,α=O(|hh0|C2),iN0.|H_{h_{0}}(\Omega_{i})|_{C^{0,\alpha}}=O(|h-h_{0}|_{C^{2}}),\quad\forall i\geq N_{0}.

According to the Schauder estimates for elliptic PDE, there exists a constant c0>0c_{0}>0, such that for any iN0i\geq N_{0},

(5.2) |fi|C2,αc0(|fi|L+|Hh0(Ωi)|C0,α).|f_{i}|_{C^{2,\alpha}}\leq c_{0}\big{(}|f_{i}|_{L^{\infty}}+|H_{h_{0}}(\Omega_{i})|_{C^{0,\alpha}}\big{)}.

Besides, for iN0i\geq N_{0}, suppose the principle curvature of DiD_{i} with respect to h0h_{0} satisfies that supiN0|λ(Di)|L<1\underset{i\geq N_{0}}{\sup}|\lambda(D_{i})|_{L^{\infty}}<1. Uhlenbeck ([24]) shows that 3\mathbb{H}^{3} is foliated by a sequence of equidistant discs relative to DiD_{i}. We denote by DirD_{i}^{r} the disc with a fix distance rr to DiD_{i}, it has mean curvature

Hh0(Dir)=2(1λ(Di)2)tanhr1λ(Di)2tanh2r.H_{h_{0}}(D_{i}^{r})=\frac{2(1-\lambda(D_{i})^{2})\tanh r}{1-\lambda(D_{i})^{2}\tanh^{2}r}.

Let Ri+R_{i}^{+} and RiR_{i}^{-} be the supremum and infimum of rr such that Ωi\Omega_{i} meets DirD_{i}^{r}, and the intersections points are xi+x_{i}^{+}, xix_{i}^{-}, respectively. Since Ri+,Ri0R_{i}^{+},R_{i}^{-}\rightarrow 0 as hh0h\rightarrow h_{0}, we may assume

min{ddr(tanhr)|r=Ri+,ddr(tanhr)|r=Ri}=min{1cosh2Ri+,1cosh2Ri}12,\min\{\frac{d}{dr}(\tanh r)|_{r=R_{i}^{+}},\,\frac{d}{dr}(\tanh r)|_{r=R_{i}^{-}}\}=\min\{\frac{1}{\cosh^{2}R_{i}^{+}},\,\frac{1}{\cosh^{2}R_{i}^{-}}\}\geq\frac{1}{2},

then we have

|Hh0(Ωi)|L\displaystyle|H_{h_{0}}(\Omega_{i})|_{L^{\infty}} max{|Hh0(DiRi+(xi+)|,|Hh0(DiRi(xi)|}\displaystyle\geq\max\{|H_{h_{0}}(D_{i}^{R_{i}^{+}}(x_{i}^{+})|,\,|H_{h_{0}}(D_{i}^{R_{i}^{-}}(x_{i}^{-})|\}
2(1|λ(Di)|L2)max{|tanhRi+|,|tanhRi|}\displaystyle\geq 2(1-|\lambda(D_{i})|^{2}_{L^{\infty}})\max\{|\tanh R_{i}^{+}|,\,|\tanh R_{i}^{-}|\}
(1|λ(Di)|L2)max{|Ri+|,|Ri|}.\displaystyle\geq(1-|\lambda(D_{i})|^{2}_{L^{\infty}})\max\{|R_{i}^{+}|,\,|R_{i}^{-}|\}.

Since Ωi\Omega_{i}, described by the graph fif_{i}, is bounded between DiRi+D_{i}^{R_{i}^{+}} and DiRiD_{i}^{R_{i}^{-}}, the above result indicates the existence of a uniform constant c1>0c_{1}>0, such that

(5.3) |fi|Lc1|Hh0(Ωi)|L,iN0.|f_{i}|_{L^{\infty}}\leq c_{1}|H_{h_{0}}(\Omega_{i})|_{L^{\infty}},\quad\forall i\geq N_{0}.

Combining (5.1)-(5.3), we obtain

|fi|C2,α=O(|hh0|C2),iN0.|f_{i}|_{C^{2,\alpha}}=O(|h-h_{0}|_{C^{2}}),\quad\forall i\geq N_{0}.

And therefore, the principal curvatures of Σi\Sigma_{i} with respect to h0h_{0} and hh satisfy that

|λh0(Σi)|L2=O(|hh0|C22),iN0,\displaystyle|\lambda_{h_{0}}(\Sigma_{i})|^{2}_{L^{\infty}}=O(|h-h_{0}|^{2}_{C^{2}}),\quad\forall i\geq N_{0},
\displaystyle\Longrightarrow |λh(Σi)|L2=|λh0(Σi)|L2+O(|hh0|C22)=O(|hh0|C22),iN0.\displaystyle\,|\lambda_{h}(\Sigma_{i})|^{2}_{L^{\infty}}=|\lambda_{h_{0}}(\Sigma_{i})|^{2}_{L^{\infty}}+O(|h-h_{0}|^{2}_{C^{2}})=O(|h-h_{0}|^{2}_{C^{2}}),\quad\forall i\geq N_{0}.

Clearly, the sectional curvature of (M,h)(M,h) has the property

|Kh|L=1+O(|hh0|C2),iN0.|K_{h}|_{L^{\infty}}=-1+O(|h-h_{0}|_{C^{2}}),\quad\forall i\geq N_{0}.

Thus, we can find 𝒰𝒰0\mathcal{U}\subset\mathcal{U}_{0} and NN0N\geq N_{0}, such that if h𝒰h\in\mathcal{U} and iNi\geq N, then the principal curvatures of Σi\Sigma_{i} with respect to hh and the sectional curvature of (M,h)(M,h) satisfy that

|λh(Σi)|L2<supKh.|\lambda_{h}(\Sigma_{i})|_{L^{\infty}}^{2}<-\sup K_{h}.

In the lemma below, we apply Uhlenbeck’s method [24], as well as the comparison result associated with Riccati equations, to prove that Σi\Sigma_{i} is the unique closed minimal surface in (M~i,h)(\tilde{M}_{i},h), thus minimizing the area among all closed surfaces.

Lemma 5.1.

Let Σ\Sigma be a minimal surface in (M,h)(M,h) whose fundamental group injectively includes in π1(M)\pi_{1}(M), the principal curvatures of Σ\Sigma with respect to hh, denoted by ±λ\pm\lambda, and the sectional curvature KK of (M,h)(M,h) satisfy that

(5.4) |λ(Σ)|L2<supK.|\lambda(\Sigma)|^{2}_{L^{\infty}}<-\sup K.

Let M~\tilde{M} be the cover of MM with π1(M~)π1(Σ)\pi_{1}(\tilde{M})\cong\pi_{1}(\Sigma). Then Σ\Sigma is the unique closed minimal surface in M~\tilde{M}.

Proof.

Denote the supreme of the sectional curvature on (M,h)(M,h) by k2-k^{2}, where k>0k>0. Let f(x)f(x) be the distance function from a fixed point in M~Σ\tilde{M}\setminus\Sigma to xΣx\in\Sigma. By (5.4), for any XTxΣX\in T_{x}\Sigma,

(5.5) D2f(X,X)=Hessf(X,X)A(X,X)(f)ktanh(kf(x))k>0.D^{2}f(X,X)=\text{Hess}\,f(X,X)-A(X,X)(f)\geq\frac{k}{\tanh(kf(x))}-k>0.

It turns out that ff is a convex function, thus, there’s only one critical point that attains the minimum. As a result, exp|NΣ\exp|_{N\Sigma} maps injectively from the normal bundle NΣN\Sigma to M~\tilde{M}.

Furthermore, we show that exp|NΣ\exp|_{N\Sigma} is a diffeomorphism, and thus M~\tilde{M} is foliated by a family of surfaces {Σr}r\{\Sigma_{r}\}_{r\in\mathbb{R}}, where Σr\Sigma_{r} is the surface at the fixed distance rr to Σ\Sigma. To see this, we introduce some notations beforehand. For xΣx\in\Sigma, choose an oriented, orthonormal basis {e1,e2}\{e_{1},e_{2}\} for TxΣT_{x}\Sigma, and a unit vector e3e_{3} for NxΣN_{x}\Sigma. Then we obtain an orthonormal frame by applying parallel transport along exp|NΣ\exp|_{N\Sigma}. Since Σ\Sigma is a minimal surface, the principal curvatures satisfy that λ1=λ2:=λ\lambda_{1}=-\lambda_{2}:=\lambda, we assume λ0\lambda\geq 0 in the following computation. Let Vi(r)=vi(r)ei(r)V_{i}(r)=v^{i}(r)e_{i}(r) be the Jacobi field along exp(re3)\exp(re_{3}), where i=1,2i=1,2, it satisfies that vi(0)=1v^{i}(0)=1, (vi)(0)=λi=±λ(v^{i})^{\prime}(0)=\lambda_{i}=\pm\lambda.

On the other hand, let Σ¯\overline{\Sigma} be a minimal surface in M~\tilde{M} with respect to an ambient metric h¯\overline{h} of constant sectional curvature k2-k^{2}, and its principal curvature satisfies that λ¯=λ\overline{\lambda}=\lambda. We do not require the existence of Σ¯\overline{\Sigma}, it’s only used for comparison in the computation. Similar to the notations defined above, let e¯1,,e¯3\overline{e}_{1},\cdots,\overline{e}_{3} be the corresponding frame on M~\tilde{M} with respect to h¯\overline{h}, and let V¯i(r)=v¯i(r)e¯i(r)\overline{V}_{i}(r)=\overline{v}^{i}(r)\overline{e}_{i}(r) be the Jacobi field along exp(re¯3)\exp(r\overline{e}_{3}) which shares the same initial data with Vi(r)V_{i}(r). Since λ<k\lambda<k, we have

v¯i(r)=coshkr±λsinhkrk>0,i=1,2.\overline{v}^{i}(r)=\cosh{kr}\pm\dfrac{\lambda\sinh{kr}}{k}>0,\quad i=1,2.

From

(vi)′′=R(e3,ei,ei,e3)vik2v¯i=(v¯i)′′(v^{i})^{\prime\prime}=-R(e_{3},e_{i},e_{i},e_{3})v^{i}\geq k^{2}\,\overline{v}^{i}=(\overline{v}^{i})^{\prime\prime}

and the initial data, the graph of viv^{i} lies above that of v¯i\overline{v}^{i}, thus above the horizontal axis. The non-vanishing Jacobi fields ensure that the induced metric on Σr\Sigma_{r} in (M~,h)(\tilde{M},h) is nonsingular for all rr\in\mathbb{R}. In addition, we’ve seen that exp|NΣ:NΣM~\exp|_{N\Sigma}:N\Sigma\rightarrow\tilde{M} is injective, and therefore also bijective, so it is a diffeomorphism and M~\tilde{M} admits a foliation structure.

Next, let λi(r)(i=1,2)\lambda_{i}(r)\,(i=1,2) be the principal curvatures on Σr\Sigma_{r}, and denote by λ¯i(r)(i=1,2)\overline{\lambda}_{i}(r)\,(i=1,2) the principal curvatures of the rr-equidistant surface to Σ¯\overline{\Sigma} with respect to h¯\overline{h}. Notice that each λi(r)\lambda_{i}(r) satisfies the Riccati equation

λi(r)=λi2(r)+R(e3,ei,ei,e3)(r).\lambda_{i}^{\prime}(r)=\lambda^{2}_{i}(r)+R(e_{3},e_{i},e_{i},e_{3})(r).

Then it follows from the comparison theorem associated with Riccati equations (for instance, see Theorem 3.1 of [26]) that

λ1(r)\displaystyle\lambda_{1}(r) λ¯1(r)=kktanh(kr)+λk+λtanh(kr)>0,\displaystyle\geq\overline{\lambda}_{1}(r)=k\,\frac{k\tanh(kr)+\lambda}{k+\lambda\tanh(kr)}>0,
λ2(r)\displaystyle\lambda_{2}(r) λ¯2(r)=kktanh(kr)λkλtanh(kr).\displaystyle\geq\overline{\lambda}_{2}(r)=k\,\frac{k\tanh(kr)-\lambda}{k-\lambda\tanh(kr)}.

It follows from λ2k2\lambda^{2}\leq k^{2} that

λ1(r)+λ2(r)2k(k2λ2)tanh(kr)k2λ2tanh2(kr)>0.\lambda_{1}(r)+\lambda_{2}(r)\geq 2k\,\frac{(k^{2}-\lambda^{2})\tanh(kr)}{k^{2}-\lambda^{2}\tanh^{2}(kr)}>0.

Therefore, for any rr\in\mathbb{R}, Σr\Sigma_{r} is strictly mean convex with respect to the metric induced by hh.

Finally, we prove the uniqueness. Assume that Σ\Sigma^{\prime} is another closed minimal surface in (M~,h)(\tilde{M},h), and let R+R_{+} and RR_{-} be the supremum and infimum of rr such that Σ\Sigma^{\prime} intersects Σr\Sigma_{r}, respectively, then R+R_{+} and RR_{-} are both finite. However, due to the maximum principle, Σ\Sigma^{\prime} cannot be tangential to any strictly mean convex slice Σr\Sigma_{r} with r0r\neq 0. Therefore, we must have ΣΣ0=Σ\Sigma^{\prime}\subset\Sigma_{0}=\Sigma. ∎

Now we finish the proof of Theorem 1.4. From the previous lemma, when ii\in\mathbb{N} is sufficiently large, Σi\Sigma_{i} is the area-minimizer among all closed surfaces in M~i\tilde{M}_{i} with respect to the induced metric of hh, it yields that

areah(F1(Si))=areah(F~i1(Si))areah(Σi).\text{area}_{h}(F^{-1}(S_{i}))=\text{area}_{h}(\tilde{F}_{i}^{-1}(S_{i}))\geq\text{area}_{h}(\Sigma_{i}).

Combining it with the area comparison in Theorem 5.1 of [19], we have

AreaF(h/h0)=limiareah(F1(Si))4π(gi1)limiareah(Σi)4π(gi1)1.\text{Area}_{F}(h/h_{0})=\lim_{i\rightarrow\infty}\frac{\text{area}_{h}(F^{-1}(S_{i}))}{4\pi(g_{i}-1)}\geq\lim_{i\rightarrow\infty}\frac{\text{area}_{h}(\Sigma_{i})}{4\pi(g_{i}-1)}\geq 1.

Moreover, when the equality holds, it follows from the equality of Theorem 5.1 of [19] that h=F(h0)h=F^{*}(h_{0}), FF is an isometry between hh and h0h_{0}.

6. Proof of Theorem 1.6

In the end, we prove Theorem 1.6 in this last section.

6.1. Proof of inequality

The proof follows directly from [19], but for readers’ convenience, it is stated as follows.

We let

α:=limiareah(Πi)4π(gi1).\alpha:=\lim_{i\rightarrow\infty}\frac{\text{area}_{h}(\Pi_{i})}{4\pi(g_{i}-1)}.

For any δ>0\delta>0, we can take ii sufficiently large, such that

limiareah(Πi)4π(gi1)<α+δ.\lim_{i\rightarrow\infty}\frac{\text{area}_{h}(\Pi_{i})}{4\pi(g_{i}-1)}<\alpha+\delta.

Let Πik\Pi_{i}^{k} be the kk-cover of Πi\Pi_{i}. Since Πi\Pi_{i} has genus gi2g_{i}\geq 2,

4π(gik1)k 4π(gi1),4\pi(g_{i}^{k}-1)\geq k\,4\pi(g_{i}-1),

then the least area surface in the homotopy class of Πik\Pi_{i}^{k} with respect to hh satisfies that

(6.1) limiareah(Πik)4π(gik1)limikareah(Πi)k 4π(gi1)=limiareah(Πi)4π(gi1)<α+δ.\lim_{i\rightarrow\infty}\frac{\text{area}_{h}(\Pi_{i}^{k})}{4\pi(g_{i}^{k}-1)}\leq\lim_{i\rightarrow\infty}\frac{k\,\text{area}_{h}(\Pi_{i})}{k\,4\pi(g_{i}-1)}=\lim_{i\rightarrow\infty}\frac{\text{area}_{h}(\Pi_{i})}{4\pi(g_{i}-1)}<\alpha+\delta.

According to Müller-Puchta’s formula (see [15]), there exists a constant cic_{i} that depends only on MM and ii, such that the following is true when gig_{i} is large.

(6.2) s(M,gik,1i)(cigik)2gik.s(M,g_{i}^{k},\frac{1}{i})\geq(c_{i}g_{i}^{k})^{2g_{i}^{k}}.

Define LikL_{i}^{k} in the following way

(6.3) 4π(Lik1)=(α+δ)4π(gik1)limkgikLik=1α+δ.4\pi(L_{i}^{k}-1)=(\alpha+\delta)4\pi(g_{i}^{k}-1)\Longrightarrow\lim_{k\rightarrow\infty}\frac{g_{i}^{k}}{L_{i}^{k}}=\frac{1}{\alpha+\delta}.

Combining (6.2) with (6.1), we have that

#{areah(Π)4π(Lik1):ΠS1i(M)}(cigik)2gik,\#\{\text{area}_{h}(\Pi)\leq 4\pi(L_{i}^{k}-1):\Pi\in S_{\frac{1}{i}}(M)\}\geq(c_{i}g_{i}^{k})^{2g_{i}^{k}},

Therefore, by (6.3),

E(h)\displaystyle E(h) =limiliminfkln#{areah(Π)4π(Lik1):ΠS1i(M)}LiklnLik\displaystyle=\underset{i\rightarrow\infty}{\lim}\,\underset{k\rightarrow\infty}{\lim\inf}\,\dfrac{\ln\#\{\text{area}_{h}(\Pi)\leq 4\pi(L_{i}^{k}-1):\Pi\in S_{\frac{1}{i}}(M)\}}{L_{i}^{k}\ln L_{i}^{k}}
limiliminfk(cigik)2gikLiklnLik\displaystyle\geq\underset{i\rightarrow\infty}{\lim}\,\underset{k\rightarrow\infty}{\lim\inf}\,\dfrac{(c_{i}g_{i}^{k})^{2g_{i}^{k}}}{L_{i}^{k}\ln L_{i}^{k}}
=2α+δ.\displaystyle=\frac{2}{\alpha+\delta}.

Since δ\delta is an arbitrarily small positive number, we conclude that

(6.4) AreaId(h/h0)E(h)=limiareah(Si)4π(gi1)E(h)limiareah(Πi)4π(gi1)E(h)2.\text{Area}_{\text{Id}}(h/h_{0})E(h)=\lim_{i\rightarrow\infty}\frac{\text{area}_{h}(S_{i})}{4\pi(g_{i}-1)}\,E(h)\geq\lim_{i\rightarrow\infty}\frac{\text{area}_{h}(\Pi_{i})}{4\pi(g_{i}-1)}\,E(h)\geq 2.

6.2. Proof of rigidity

If AreaId(h/h0)E(h)=2\text{Area}_{\text{Id}}(h/h_{0})E(h)=2, then (6.4) yields that

(6.5) limiareah(Πi)areah(Si)=1.\lim_{i\rightarrow\infty}\frac{\text{area}_{h}(\Pi_{i})}{\text{area}_{h}(S_{i})}=1.

To make use of this equality, we run the mean curvature flow in (Bn,h)(B^{n},h) with initial condition DiD_{i}, which is the lift of SiS_{i} in n\mathbb{H}^{n}, then we estimate the decay rate of the area. Firstly of all, we need to review and establish some tools for complete, noncompact surfaces moving by mean curvature. The classical short time existence theorem for compact manifolds moving by mean curvature is well-known [10]. However, the general theory for complete, noncompact manifolds has not been established in the literature. There are only several essential contributions in some special cases: Ecker-Huisken [7] proved the codimension one case in which only a local Lipschitz condition on the initial hypersurface was required. For higher codimensions, Chau-Chen-He [5] discussed the case of nonparametric mean curvature flow for flat metrics. The result related to our case is listed as follows.

Lemma 6.1.

There exist T>0T>0 and C>0C>0 depending only on MM, so that for sufficiently large ii\in\mathbb{N}, we can find a solution Di(t)D_{i}(t) to the mean curvature flow in (M,h)(M,h) with initial condition Di(0)=DiD_{i}(0)=D_{i}, where 0tT0\leq t\leq T. Additionally, the mean curvature of Di(t)D_{i}(t) and its derivative are both bounded uniformly by CC.

Proof.

Notice that after passing to a subsequence, DiD_{i} converges smoothly on compact sets to a disc DD, and each of them is a cover of a compact surface in MM. Take xDx\in D, the standard theory indicates that there is a number T0>0T_{0}>0, such that for any kk\in\mathbb{N}, we can find a solution Dik(t)D_{i}^{k}(t) to the mean curvature flow with initial condition B(x,k)¯Di\overline{B(x,k)}\cap D_{i}, where 0tT00\leq t\leq T_{0}. Since T0T_{0} depends only on the second fundamental form of DD, in particular, it’s independent of ii and kk.

Next, in order to apply the Arzela-Ascoli theorem and estimate the mean curvature of Di(t)D_{i}(t) and its derivative for any small time tt, we need the following preparation. Claim that for any δ>0\delta>0, and any spacetime Xik=(xik,t)X_{i}^{k}=(x_{i}^{k},t) of Dik(t)D_{i}^{k}(t), there exists an open neighborhood UikU_{i}^{k} of XikX_{i}^{k}, so that the Guassian density ratio

Θ(Dik(t),Xik,r):=yDik(tr2)14πr2exp(|yxik|24r2)𝑑2(y)\Theta(D_{i}^{k}(t),X_{i}^{k},r):=\int_{y\in D_{i}^{k}(t-r^{2})}\frac{1}{4\pi r^{2}}\exp\big{(}-\frac{|y-x_{i}^{k}|^{2}}{4r^{2}}\big{)}\,d\mathcal{H}^{2}(y)

satisfies that

(6.6) Θ(Dik(t),Xik,r)1+δ, 0<r<d(Xik,Uik).\Theta(D_{i}^{k}(t),X_{i}^{k},r)\leq 1+\delta,\quad\forall\,0<r<d(X_{i}^{k},U_{i}^{k}).

If this wasn’t true for some integers ii and kk, then we could pick a sequence λj\lambda_{j}\rightarrow\infty as jj\rightarrow\infty, and

Θ(𝒟λj(Dik(t)Xik),0,λjr)>1+δ,\Theta(\mathcal{D}_{\lambda_{j}}(D_{i}^{k}(t)-X_{i}^{k}),0,\lambda_{j}r)>1+\delta,

where 𝒟λ\mathcal{D}_{\lambda} denotes the parabolic dilation 𝒟λ(y,t)=(λy,λ2t)\mathcal{D}_{\lambda}(y,t)=(\lambda y,\lambda^{2}t). Since the second fundamental form satisfies |A|L(𝒟λj(Dik(t))Xik)20|A|^{2}_{L^{\infty}(\mathcal{D}_{\lambda_{j}}(D_{i}^{k}(t))-X_{i}^{k})}\rightarrow 0 as jj\rightarrow\infty, 𝒟λj(Dik(t)Xik)\mathcal{D}_{\lambda_{j}}(D_{i}^{k}(t)-X_{i}^{k}) converges smoothly to a disc D¯ik\bar{D}_{i}^{k} whose second fundamental form vanishes. However, the inequality above implies that

limjΘ(D¯ik,0,λjr)>1,\lim_{j\rightarrow\infty}\Theta(\bar{D}_{i}^{k},0,\lambda_{j}r)>1,

which contradicts the topology of D¯ik\bar{D}_{i}^{k}.

We’ve seen that (6.6) holds, so due to the local regularity theorem in [27], there is a uniform constant C0C_{0} that is independent of ii and kk, so that at any spacetime Xik=(xik,t)X_{i}^{k}=(x_{i}^{k},t),

(6.7) |A|2(Xik)d(Xik,Uik)C0.|A|^{2}(X_{i}^{k})d(X_{i}^{k},U_{i}^{k})\leq C_{0}.

Therefore, Arzela-Ascoli theorem (see page 1494 of [27]) implies the short-time existence of the mean curvature flow with noncompact initial condition Di(0)=DiD_{i}(0)=D_{i} on time interval [0,T0][0,T_{0}]. Moreover, the interior estimate (6.7) validates the condition of the maximum principle ([7], Theorem 4.3). Arguing like Theorem 4.4 of [7], we can find T>0T>0 and C>0C>0 that make the lemma hold.

Next, following the method of Lemma 6.5 in [19], we prove a similar result for the case of the higher codimensions.

Lemma 6.2.
limi1areah(Di)Di|Hh|2𝑑Ah=0,\underset{i\rightarrow\infty}{\lim}\,\dfrac{1}{\text{area}_{h}(D_{i})}\int_{D_{i}}|H_{h}|^{2}dA_{h}=0,

where HhH_{h} denotes the mean curvature of each disc DiD_{i} in (M,h)(M,h).

Proof.

Suppose by contradiction that there exists ϵ>0\epsilon>0, such that after passing to a subsequence, and for ii\in\mathbb{N} large enough,

(6.8) 1areah(Di)Di|Hh|2𝑑Ah>2ϵ.\dfrac{1}{\text{area}_{h}(D_{i})}\int_{D_{i}}|H_{h}|^{2}dA_{h}>2\epsilon.

Under the mean curvature flow, the mean curvature satisfies the following evolution equation on the time interval t[0,T]t\in[0,T] (see [23]).

ddt|Hh(t)|2=\displaystyle\nabla_{\frac{d}{dt}}|H_{h}(t)|^{2}= Δ|Hh(t)|22|Hh(t)|2+4Ahjk(t),Hh(t)ht(Ah)jk(t),Hh(t)ht\displaystyle\Delta|H_{h}(t)|^{2}-2|\nabla H_{h}(t)|^{2}+4\langle A_{h}^{jk}(t),H_{h}(t)\rangle_{h_{t}}\langle(A_{h})_{jk}(t),H_{h}(t)\rangle_{h_{t}}
+2(Rh)jklm(t)Hhj(t)(Fi)pk(t)Hhl(t)Fimp(t),\displaystyle+2(R_{h})_{jklm}(t)H_{h}^{j}(t)(F_{i})_{p}^{k}(t)H_{h}^{l}(t)F_{i}^{mp}(t),

where Hh(t)H_{h}(t), Ah(t)A_{h}(t) and Fi(t)F_{i}(t) represent the mean curvature, second fundamental form and the immersion Fi(t):Di(t)MF_{i}(t):D_{i}(t)\rightarrow M, respectively.

Using the result of Lemma 6.1, we can pick a uniform constant C1>0C_{1}>0, such that for all sufficiently large ii\in\mathbb{N}, and for any t[0,T]t\in[0,T],

ddtDi(t)|Hh(t)|2𝑑AhC1areah(Di(t))C1areah(Di),{\frac{d}{dt}}\int_{D_{i}(t)}|H_{h}(t)|^{2}dA_{h}\geq-C_{1}\text{area}_{h}(D_{i}(t))\geq-C_{1}\text{area}_{h}(D_{i}),

the latter inequality follows from the fact

ddtareah(Di(t))=12Di(t)trddthjk,hjk𝑑Ah=Di(t)|Hh(t)|2𝑑Ah0.\frac{d}{dt}\text{area}_{h}(D_{i}(t))=\frac{1}{2}\int_{D_{i}(t)}tr\langle\frac{d}{dt}h_{jk},h^{jk}\rangle\,dA_{h}=-\int_{D_{i}(t)}|H_{h}(t)|^{2}dA_{h}\leq 0.

We can choose T1<min{ϵC1,T}T_{1}<\min\{\frac{\epsilon}{C_{1}},T\}, by assumption (6.8), for any ii\in\mathbb{N} and t[0,T1]t\in[0,T_{1}],

Di(t)|Hh(t)|2𝑑Ahϵareah(Di)ϵareah(Di(t)).\int_{D_{i}(t)}|H_{h}(t)|^{2}dA_{h}\geq\epsilon\,\text{area}_{h}(D_{i})\geq\epsilon\,\text{area}_{h}(D_{i}(t)).

Then we obtain

ddtareah(Di(t))=Di(t)|Hh(t)|2𝑑Ahϵareah(Di(t)).\frac{d}{dt}\text{area}_{h}(D_{i}(t))=-\int_{D_{i}(t)}|H_{h}(t)|^{2}dA_{h}\leq-\epsilon\,\text{area}_{h}(D_{i}(t)).

Thus, for any sufficiently large ii\in\mathbb{N},

areah(Πi)areah(Di)areah(Di(t))areah(Di)eϵT1areah(Di)areah(Di)=eϵT1<1,\frac{\text{area}_{h}(\Pi_{i})}{\text{area}_{h}(D_{i})}\leq\frac{\text{area}_{h}(D_{i}(t))}{\text{area}_{h}(D_{i})}\leq\frac{e^{-\epsilon T_{1}}\text{area}_{h}(D_{i})}{\text{area}_{h}(D_{i})}=e^{-\epsilon T_{1}}<1,

which violates (6.5). ∎

Furthermore, arguing like Lemma 4.2 of [13], we deduce the following result from Lemma 6.2. For any round circle cnc\subset\partial_{\infty}\mathbb{H}^{n}, it has a dense π1(M)\pi_{1}(M)-orbit in n\partial_{\infty}\mathbb{H}^{n}. In addition, cc can be represented by limiΛ(ϕiΠiϕi1)\underset{i\rightarrow\infty}{\lim}\Lambda(\phi_{i}\Pi_{i}\phi_{i}^{-1}), where ϕiπ1(M)\phi_{i}\in\pi_{1}(M), and Λ(ϕiΠiϕi1)\Lambda(\phi_{i}\Pi_{i}\phi_{i}^{-1}) represents the limit set of ϕiΠiϕi1\phi_{i}\Pi_{i}\phi_{i}^{-1}. Redefine DiD_{i} by the lifts of SiS_{i} to n\mathbb{H}^{n} preserved by ϕiΠiϕi1\phi_{i}\Pi_{i}\phi_{i}^{-1}. It has the property that

limiDiBRi(0)|Hh|2𝑑Ah=0,Ri.\lim_{i\rightarrow\infty}\int_{D_{i}\cap B_{R_{i}}(0)}|H_{h}|^{2}dA_{h}=0,\quad R_{i}\rightarrow\infty.

Note that after passing to a subsequence, DiD_{i} converges to the totally geodesic disc D(c)nD(c)\subset\mathbb{H}^{n} that is asymptotic to cc. Therefore, the mean curvature HhH_{h} vanishes on D(c)D(c), namely, D(c)D(c) is a minimal disc of BnB^{n} with respect to the metric hh. And since cc is chosen arbitrarily, every totally geodesic disc of n\mathbb{H}^{n} must be minimal for hh.

We apply the result below for surfaces in 3-manifolds, the proof can be found in [19].

Lemma 6.3.

Every totally geodesic disc in 3\mathbb{H}^{3} is minimal with respect to another metric hh if and only if for any geodesic γ3\gamma\subset\mathbb{H}^{3}, the following function is a constant

t|h|h012h(γ(t),γ(t)).t\mapsto|h|_{h_{0}}^{-\frac{1}{2}}h\,(\gamma^{\prime}(t),\gamma^{\prime}(t)).

Because of the ergodicity of the geodesic flow in (M,h0)(M,h_{0}), we can choose a geodesic γ\gamma of MM whose orbit is dense in the unit tangent bundle. Let γ~\tilde{\gamma} be the lift of γ\gamma to n\mathbb{H}^{n}. γ~\tilde{\gamma} must be contained in a hyperbolic 3-ball B3B\approx\mathbb{H}^{3}. Applying the previous lemma to the geodesic γ~\tilde{\gamma} and ambient manifold BB, we conclude that

t|h|B|h012h|B(γ~(t),γ~(t))t\mapsto|h|_{B}|_{h_{0}}^{-\frac{1}{2}}h|_{B}\,(\tilde{\gamma}^{\prime}(t),\tilde{\gamma}^{\prime}(t))

is constant. So the projection γ\gamma in MM also satisfies that

t|h|h012h(γ(t),γ(t))t\mapsto|h|_{h_{0}}^{-\frac{1}{2}}h\,(\gamma^{\prime}(t),\gamma^{\prime}(t))

is constant. Thus due to the density, there is a constant c>0c>0, such that for any vector field XX of the unit tangent bundle of MM,

|h|h012h(X,X)=ch0(X,X)|h|h012h=ch0.|h|_{h_{0}}^{-\frac{1}{2}}h(X,X)=c\,h_{0}(X,X)\quad\Longrightarrow\quad|h|_{h_{0}}^{-\frac{1}{2}}h=c\,h_{0}.

As a result, hh coincides with a multiple of h0h_{0}.

References

  • [1] A. L. Besse. Einstein Manifolds, volume 10 of Ergebnisse der Mathematik und ihrer Grenzgebiete. Springer Berlin, Heidelberg, 1987.
  • [2] G. Besson, G. Courtois, and S. Gallot. Volume et entropie minimale des espaces localement symétriques. Inventiones mathematicae, 103(2):417–446, 1991.
  • [3] G. Besson, G. Courtois, and S. Gallot. Entropies et rigidités des espaces localement symétriques de courbure strictement négative. Geometric and Functional Analysis, 5:731–799, 1995.
  • [4] D. Calegari, F. Marques, and A. Neves. Counting minimal surfaces in negatively curved 3-manifolds. Duke Mathematical Journal, 2022.
  • [5] A. Chau, J. Chen, and W. He. Lagrangian mean curvature flow for entire lipschitz graphs. Calculus of Variations and Partial Differential Equations, 44:199–220, 2012.
  • [6] D. G. Ebin. On the space of Riemannian metrics. Bulletin of the American Mathematical Society, 74(5):1001–1003, 1968.
  • [7] K. Ecker and G. Huisken. Interior estimates for hypersurfaces moving by mean curvature. Inventiones mathematicae, 105(3):547–570, 1991.
  • [8] M. Gromov. Foliated plateau problem, part I: Minimal varieties. Geometric and Functional Analysis, 1:14–79, 1991.
  • [9] U. Hamenstädt. Incompressible surfaces in rank one locally symmetric spaces. Geometric and Functional Analysis, 25, 02 2014.
  • [10] R. S. Hamilton. Heat equations in geometry. Lecture notes, Honolulu, Hawaii, 1989.
  • [11] B. Hasselblatt and A. Katok. Chapter 1 Principal structures, volume 1 of Handbook of Dynamical Systems, pages 1–203. Elsevier Science, 2002.
  • [12] J. Hempel. 3-manifolds. Annals of Mathematics Studies 86, Princeton University Press, 1976.
  • [13] R. Jiang. Counting essential minimal surfaces in closed negatively curved n-manifolds. arXiv:2108.01796, 2021.
  • [14] J. Kahn, F. Labourie, and S. Mozes. Surface groups in uniform lattices of some semi-simple groups. arXiv: Differential Geometry, 2018.
  • [15] J. Kahn and V. Markovic. Counting essential surfaces in a closed hyperbolic 3-manifold. Geom. Topol., 16:601–624, 2012.
  • [16] G. Knieper. Volume growth, entropy and the geodesic stretch. Mathematical Research Letters, 2:39–58, 1995.
  • [17] F. Labourie. Asymptotic counting of minimal surfaces in hyperbolic manifolds [according to Calegari, Marques and Neves]. arXiv:2203.09366, 2022.
  • [18] B. Lowe. Deformations of totally geodesic foliations and minimal surfaces in negatively curved 3-manifolds. Geom. Funct. Anal., 31:895–929, 2021.
  • [19] B. Lowe and A. Neves. Minimal surface entropy and average area ratio. arXiv:2110.09451, 2021.
  • [20] M. Ratner. Raghunathan’s topological conjecture and distributions of unipotent flows. Duke Mathematical Journal, 63(1):235–280, 1991.
  • [21] R. Schoen and S. T. Yau. Existence of incompressible minimal surfaces and the topology of three dimensional manifolds with non-negative scalar curvature. Annals of Mathematics, 110(1):127–142, 1979.
  • [22] N. A. Shah. Closures of totally geodesic immersions in manifolds of constant negative curvature. Singapore: World Scientific, 1991.
  • [23] K. Smoczyk. Mean curvature flow in higher codimension - introduction and survey. In Global Differential Geometry, pages 231–274. Springer Berlin Heidelberg, 2012.
  • [24] K. Uhlenbeck. Closed minimal surfaces in hyperbolic 3-manifolds, pages 147–168. Seminar On Minimal Submanifolds. (AM-103). Princeton University Press, 1983.
  • [25] S. Wang. The π1\pi_{1}-injectivity of self-maps of nonzero degree on 3-manifolds. Mathematische Annalen, 297(1):171–190, 1993.
  • [26] F. W. Warner. Extension of the Rauch comparison theorem to submanifolds. Transactions of the American Mathematical Society, 122(2):341–356, 1966.
  • [27] B. White. A local regularity theorem for mean curvature flow. Annals of mathematics, 161(3):1487–1519, 05 2005.