Average area ratio and normalized total scalar curvature of hyperbolic n-manifolds
Abstract.
On closed hyperbolic manifolds of dimension , we review the definition of the average area ratio of a metric with relative to the hyperbolic metric , and we prove that it attains the local minimum of one at , 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 -manifolds, as well as its relation to the minimal surface entropy if is odd.
1. Introduction
On a closed negatively curved manifold , 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 and on , the comparisons of the geometric objects associated with and have been studied and become well-known. For instance, the geodesic stretch measures the stretching of the metric relative to the reference metric and the measure . Comparisons related to were discussed by Knieper [16]. Another example is the volume entropy that counts the number of closed geodesics, the locally symmetric metric attains the minimum among all metrics on 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 is formed by a family of stable minimal surfaces of . In particular, let denote the hyperbolic metric, there is a canonical foliation of whose leaves are totally geodesic planes. Gromov introduced the average area ratio which is similar in spirit to the geodesic stretch, to measure the “stretching” of the area of leaves on of the metric relative to the hyperbolic metric under the map . The comparison result and the interpretation of for -manifolds were studied by Lowe-Neves [19] using Ricci flow. In this paper, we focus on the local picture of for any dimension , and investigate a more general case for -manifolds.
On the other hand, Calegari-Marques-Neves [4] defined the minimal surface entropy based on the construction of surface subgroups by Kahn-Markovic [15], and the equidistribution property of -action on the space of minimal laminations (initiated by Ratner [20] and Shah [22], and recently formalized by Labourie [17]). measures the number of essential minimal surfaces of with respect to , thus shifting attention from a one-dimensional object (volume entropy) to dimension two. The minimal surface entropy of closed hyperbolic -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 in [13], we continue the discussion for higher dimensions in this paper.
1.1. Gromov’s conjecture
Let be a closed hyperbolic -manifold, and let be a closed -manifold with Riemannian metric whose scalar curvature satisfies . is a smooth map with degree . The average area ratio of is as follows.
where is a subset of the totally geodesic disc of which is tangential to at , has an area equal to , and stands for the unit volume measure on with respect to the metric induced by . We refer to Section 2 for a more detailed definition. Gromov proved the following inequality using stability (see page 73 of [8]).
The proof indicates that if the equality holds, then 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 . Moreover, the higher dimensional case may share the same property.
Conjecture 1.1.
Let be a closed hyperbolic manifold of dimension , and let be a closed -manifold with Riemannian metric , and its scalar curvature satisfies . is a smooth map with degree . We have
The equality holds if and only if is a local isometry.
1.2. Minimal Surface Entropy
Let denote the hyperbolic -space, where . In the Poincaré ball model, the asymptotic boundary can be considered as the -unit sphere . A homeomorphism is called -quasiconformal if , where the dilatation of is
the supremum is taken over so that . A -quasicircle in is the image of a round circle under a -quasiconformal map.
Let be a closed orientable -manifold that admits a hyperbolic metric , A closed surface immersed in with genus at least is said to be essential if the immersion is -injective, and the image of its fundamental group in is called a surface subgroup. Let denote the set of surface subgroups of genus at most up to conjugacy, and let the subset consist of the conjugacy classes whose limit sets are -quasicircles. Moreover,
Suppose is an arbitrary Riemannian metric on . For any , we set
Then the minimal surface entropy with respect to is defined as follows.
(1.1) |
According to [4] and [13], for odd-dimensional manifolds, and among metrics with sectional curvature less than or equal to , attains the minimum at the hyperbolic metric , and . On the other hand, Theorem 1.1 of [19] shows that when , is maximized at among all metrics with scalar curvature greater than or equal to .
1.3. Main theorems
In the following results, we assume is a closed hyperbolic manifold of dimension .
Theorem 1.2.
There exists a small neighborhood of in the metric space of , such that for any Riemannian metric on with , we have
The equality holds if and only if .
Note that such a neighborhood is “cylindrical”, that is, if , then any metric in the conformal class of with also belongs to .
Let be another closed hyperbolic manifold of the same dimension , the theorem leads to the following corollary.
Corollary 1.3.
There exists a small neighborhood of in the metric space of , such that for any metric with , and for any local diffeomorphism with degree , we have
The equality holds if and only if is a local isometry between and , i.e., .
Proof.
Since is a local diffeomorphism,
is a hyperbolic metric on , so it’s isometric to due to Mostow rigidity. Thus we obtain from the previous theorem the following inequality.
∎
In particular, when , we prove a more general result for any smooth map between homeomorphic 3-manifolds which is not necessarily a diffeomorphism. Let and be closed hyperbolic 3-manifolds.
Theorem 1.4.
If , then there exists a small neighborhood of in the -topology, such that for any metric with , and for any smooth map , we have
In a different situation, suppose that is isomorphic to an index subgroup . Let denote the covering space of with and let denote the covering map, the hyperbolic metric on is still represented by . We have the following corollary.
Corollary 1.5.
There exists a small neighborhood of in the -topology, such that for any metric with , and for any smooth map satisfying , we have
Proof.
Besides, when the dimension of 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 in [13]. As a result, we can extend Theorem 1.2 of [19] to any odd dimensions at least 3.
Theorem 1.6.
Let be a closed hyperbolic manifold of an odd dimension . For any Riemannian metric on ,
(1.2) |
the equality holds if and only if for some constant .
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 is a compact space of any dimension with negative curvature, and is another metric on . It says that
where is the Bowen-Margulis measure of which is the unique measure with maximal measure theoretic entropy. The equality holds if and only if the geodesic flows of and 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) with dimension ,
the inequality is sharp unless is isometric to 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).
the equality holds if and only if and 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 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 be a closed hyperbolic manifold of dimension , and let be another closed Riemannian manifold of the same dimension. Suppose that is a smooth map. For any in the Grassmannian bundle , denotes the Jacobian of at plane . Take an arbitrary regular value of , for , we let
(2.1) |
This definition ([19]) is equivalent to Gromov’s definition ([8]):
where is a subset of the totally geodesic disc which is tangential to at , and has area equal to .
The average area ratio of is defined in [8] by
(2.2) |
where stands for the unit volume measure on with respect to the metric induced by .
2.2. Normalized Total Scalar Curvature
Let be a closed hyperbolic manifold of dimension , and let be the space of Riemannian metrics on . The total scalar curvature (or Einstein-Hilbert functional) is
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
This is called the normalized total scalar curvature (or normalized Einstein-Hilbert functional) of . Under conformal deformations, the first variation of is
It equals to zero provided that has constant scalar curvature. Assuming is constant, we can simplify the full variation to
Thus, a metric is critical if and only if is Einstein. In particular, the hyperbolic metric is a critical point for . Furthermore, since is scale-invariant, from now on we may assume that for ,
Taking this into account, we obtain the second variation of at as follows.
(2.3) | ||||
Substituting the formulas
where , we obtain that
According to Ebin’s Slice Theorem (see [6]), for any lying in a small neighborhood of , there exist , , and a transverse-traceless tensor , i.e., and , such that . Then we can simplify the second variation.
(2.4) |
And we also get
(2.5) |
Thus,
Taking the estimates of (2.4) and (2.5) into consideration, we can use this expansion to discuss the local behavior of . In particular, as discussed in [2], reaches a local maximum at , and there exists , so that for any metric in a small neighborhood of in ,
The inequality is sharp unless and are isometric via some .
2.3. Equidistribution
Suppose that the hyperbolic manifold has an odd dimension . According to the construction by Hamenstädt [9], for any small number , there is an essential surface in which is sufficiently well-distributed and -quasigeodesic (i.e. the geodesics on the surface with respect to intrinsic distance are -quasigeodesics in ). And as discussed in [13], determines an -quasiconformal map on , and thus associated with an element in .
Furthermore, let denote the subset of consisting of homotopy classes of finite covers of that have genus at most . It has cardinality comparable to . Moreover, let denote the minimal representative of an element in , then it is homotopic to an -quasigeodesic surface . From Lemma 4.3 of [4], for any continuous function on , the unit Radon measure induced by integration over satisfies
(2.6) |
where the limiting measure is positive on any non-empty open set of .
Notice that the measure is not necessarily identified with the unit Radon measure on induced by integration over , the latter measure is denoted by . However, in this paper, we need to find a sequence of minimal surfaces whose Radon measures defined above converge to . 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 has an odd dimension.
3. Gromov’s Conjecture in Higher Dimensions
Throughout this section, we can choose the dimension of to be any integer . The proof of Theorem 1.2 separates into two parts. If is an arbitrary metric conformal to a metric with constant scalar curvature, we compare their average area ratios in Theorem 3.1. And if is a metric with constant scalar curvature different from , we make use of the evaluations of normalized Einstein-Hilbert functional in [2].
3.1. Conformal deformations
Firstly, given a metric on , we look at the conformal class of . Since admits a hyperbolic metric, every conformal class 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 with constant scalar curvature . In Theorem 1.2, we assume that . Set , so , and for any surface or 2-dimensional subset in , we have
Therefore, to prove the theorem, we may assume that .
Theorem 3.1.
Suppose the scalar curvature , then we have
(3.1) |
Furthermore, for any surface subgroup ,
(3.2) |
and as an immediate result,
(3.3) |
Each of the above equalities holds if and only if .
Proof.
Set . The conformal factor satisfies that
Let , we obtain , and it yields that
(3.4) |
Thus, for any subset of a totally geodesic disc in with hyperbolic area equal to , we have
The inequality (3.1) follows from the definition of the average area ratio (2.2).
In addition, it also shows that for any surface ,
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 , due to the uniqueness of the solution to Yamabe problem, must be identical to . ∎
3.2. Definition of functional
Let be the space of all Riemannian metrics on , and let be the subset consisting of metrics with constant scalar curvature . From the previous section, it remains to consider metrics in .
Define a functional from the space of Riemannian metrics on to as follows.
where is a subset of the totally geodesic disc in that is tangential to at , and has hyperbolic area equal to , is the unit volume measure on with respect to the metric induced by . Notice that the metric satisfies
Therefore, from the definition of the average area ratio (2.2), to deduce that if is in a small neighborhood of in , then
we only need to show that attains a local maximum at .
3.3. Proof of Theorem 1.2
To see this, we’ll discuss the first and second variations of in detail. Before start, we notice that the functional is scale-invariant, so it suffices to assume , and the symmetric tensor satisfies . Let , we rewrite as
where
Then we have
(3.5) |
In addition,
(3.6) |
where is the rough Laplacian with negative eigenvalues, and is the double divergence operator. In particular, when , applying the Stokes’ theorem, we have
On the other hand, since
let be the eigenvalues of the matrix at point with respect to , we obtain by computation that
It follows that for any symmetric 2-tensor ,
thus is a critical point of .
Now we proceed to compute the second variation at . Note that is an analogue of the normalized total scalar curvature , it’s easier to compare their second variations using the computation in Section 2.2. Based on (3.5) and (2.3),
(3.7) | ||||
Next, using
we estimate the first term on the right-hand side of (3.7).
(3.8) | ||||
Combining (3.7) and (3.8), we have
(3.9) | ||||
To simplify this quadratic form, we decompose 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
where represents the conformal deformations of , is the action of the diffeomorphism group on , and stands for the set of transverse-traceless tensors. Let , and let , it decomposes into , where , , and . The second variation of has the form
By Theorem 3.1, to check that reaches a local maximum at on , it remains to analyze the sign of and estimate . To see these, we prove the following lemmas.
Lemma 3.2.
There exists a constant , such that in the decomposition of , and satisfy
where .
Lemma 3.3.
For any , we can find a -neighborhood of on and , such that for any , in the decomposition of satisfies
Proof of Lemma 3.2.
(3.10) | ||||
To find an upper bound of the first term, we use the estimate in Lemma 2.9 of [2]. There exists a constant , such that
And thus, substituting the above inequality and into (3.9), we obtain
(3.11) | ||||
To deal with the remaining terms of (3.10), we apply the diffeomorphism invariance property of , which says
(3.12) |
Moreover, for any ,
(3.13) |
Therefore, the second variations comparison (3.7), in company with (3.8), says that
(3.14) |
Since , it can be expressed by the Lie derivative of the metric in the direction . And using Helmholtz-Hodge decomposition, decomposes further into , where is a scalar function, and is a vector field with . By computation,
(3.15) |
In addition,
(3.16) | ||||
where the second term is in the form
(3.17) | ||||
and the last term of (3.16) vanishes, since
(3.18) | ||||
Substituting (3.15)-(3.18) into (3.14), then applying Bochner’s formula, we obtain
(3.19) | ||||
Furthermore, after eliminating some terms using (3.12), (3.13) and , we get
(3.20) |
it vanishes because of the -orthogonality between and . Similarly, the traceless-transverse property of simplifies (3.9 to
(3.21) |
∎
Proof of Lemma 3.3.
Since , we have
it leads to
(3.22) |
On the left-hand side, we have as analyzed earlier, and it follows from (3.6) that
(3.23) | ||||
On the other hand, the right-hand side of (3.22) can be estimated using the continuity of , , , and . For any , after shrinking the neighborhood of in , we have
(3.24) |
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 , such that
Following the same procedure to compute and discuss the continuity near , we can see that . Thus, from the above lemmas, there exists ,
From the computation of (3.19) and its derivative, we can see that as , and decay at the same rate, so the norm . Consequently, the following expansion holds for metric in a small neighborhood of with .
where . Note that requires that , and thus .
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 be a closed hyperbolic manifold of dimension . And let be the space of conformal minimal immersions , such that is an -quasicircle. As discussed in Section 2 of [13], when is sufficiently small, is a stable embedded disc in . The space equips with the topology of uniform convergence on compact sets, and we take
with the quotient topology. The space together with the action of by pre-composition
(4.1) |
is called the conformal minimal lamination of . A laminar measure on stands for a probability measure which is invariant under the -action defined as above. The space is sequentially compact, but the space of laminar measures is not necessarily weakly compact. In light of that, we consider a continuous map from to the frame bundle of , the latter space is compact, so the space of probability measures on is compact in weak- topology.
Firstly we define a map from to the 2-vector bundle on consisting of , where denotes the unit sphere in the tangent space to at . Let be an orthonormal basis of , and for any , let , where denotes the conformal factor between the hyperbolic metric on and the pull-back metric of by . Let and . We define the following continuous map.
Furthermore, it induces a map from to the frame bundle by parallel transport:
And define the projection
We consider the subspace , it contains isometric immersions whose images are totally geodesic discs in . Conversely, each totally geodesic disc is uniquely determined by , and tangent vectors , . Let be the restriction of to , it’s therefore a bijection. Using (4.1), We can define the -action on as follows.
This definition coincides with the homogeneous action of on . Following the discussion of Lemma 3.2 of [19], we conclude the following result.
Proposition 4.1.
Given any sequence of laminar measures on , the sequence of induced measures on converges weakly to a probability measure , then is invariant under the homogeneous action of .
Let be a Fuchsian subgroup, then is a closed hyperbolic surface with genus whose fundamental domain is represented by . And let equivariant with respect to a representation of in . The image of in is closed minimal surface in whose fundamental group is . We define a laminar measure associated with as follows.
(4.2) |
where denotes the bi-invariant measure on .
4.2. Equidistribution
In this section, we assume the dimension of 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 , there is a lamination in equivariant with respect to a representation of a Fuchsian group in , such that converges to the Lebesgue measure on as .
Sketch of the proof.
Let be the space of tripods , where . Each element determines an ideal triangle in . Let be the barycenter of . Denote by the orthonormal basis of , and denote by the orthonormal basis of the normal bundle of in . Thus, each tripod determines a point in , which represents the frame bundle of .
Consider the closed manifold . Let be the corresponding point of in , and let be the point in the frame bundle corresponding to in .
A quintuple is called a triconnected pair of tripods (Definition 10.1.1 of [14]) if and are three distinct homotopy classes of paths connecting to . The space of triconnected pair of tripods is denoted by . Let and be the forgetting maps from to
Moreover, let and be the corresponding maps from to
In addition, there is a weighted measure on as defined in Definition 11.2.3 of [14]. If is in the support of , then the ideal triangles determined by and can be glued to an -almost closing pair of pants (see Definition 9.1.1 of [14]). Moreover, it follows from the mixing property that for fixed , as , and both converge to the Lebesgue measure on .
Arguing like Theorem 5.7 of [17], we can choose a sequence as , and a sequence of measures . Then we approximate each by another weighted measure supported in finitely many pleated pair of pants , which can be glued together to get essential surfaces in . When is sufficiently large and for each , is -quasigeodesic, and the projection from to the unique minimal surface homotopic to is -bi-Lipschitz and it has distance uniformly bounded by . For this reason, we can further approximate by a weighted measure supported in . is obtained by a lamination , in fact, it’s the image of in , and thus associated with the laminar measure , we have the following lemma.
Lemma 4.3.
For any , there exist a finite sequence of laminations in , and with , such that each is equivariant with respect to a representation of a Fuchsian group in , and the laminar measure
satisfies that converges to the Lebesgue measure on as .
Next, for , we define
Then contains at most countably many candidates. Therefore, we can find a decreasing sequence of tubular neighborhoods , so that for any , covers and it satisfies and . In consequence of previous lemma, after passing to a subsequence, we have . Additionally, as argued in Lemma 6.2 of [19], we can find a subsequence , and , such that .
As a result of Proposition 4.1, as , converges weakly to a probability measure on . is invariant under the homogeneous action of , and it satisfies that
(4.3) |
To finish the proof, we need the following lemma.
Lemma 4.4.
.
Proof.
According to the ergodic decomposition theorem ([11]), can be expressed by a linear combination of the ergodic measures for -action on . Moreover, Ratner’s measure classification theorem (see [20] or [22]) says that any ergodic -invariant measure on is either an invariant probability measure supported on a finite union of , or it is identical to . Thus, we can write as
where and represents an ergodic measure supported on , . By (4.3), for all ,
So
We must have , and therefore . ∎
Proposition 4.2 follows immediately from the lemma. ∎
4.3. Average area ratio formula
Lemma 4.5 (average area ratio formula).
Let be a closed Riemannian manifold that also has odd dimension , and let be a smooth map that takes to . For , we pick a lamination equivariant with respect to a representation of in , and it satisfies Proposition 4.2. Let be the image of in . Then we have
Proof.
Recall that is a function defined almost everywhere on . Since can be regarded as a smooth function on by
based on the definition (2.1), is also seen as a function defined almost everywhere on . Thus, Proposition 4.2 implies that
In light of the definition of laminar measure in (4.2), we have
where is the fundamental domain of . Set . Since the hyperbolic surface has area equal to , where denotes the genus. The above expression also can be written as
where denotes the conformal factor between the hyperbolic metric on and the pull-back metric of by , namely . Since the Gaussian curvature on has the form , we have
On the other hand, the co-area formula yields that
Combining these formulas, we have
Since as (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 follows from Corollary 0.3 of [25] and the geometrization theorem for 3-manifolds. Next, it’s easy to see, the induced map is surjective, since otherwise, it factors through a -fold covering space of with , and thus , violating the degree one observation. In addition, is a Hopfian group (for example, see 15.13 of [12]), so the surjectivity of can be upgraded to be an isomorphism, which makes a homotopy equivalence between and due to Whitehead’s theorem. Furthermore, the Mostow rigidity theorem indicates that 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 be a closed hyperbolic 3-manifold. There exists a small neighborhood of in the -topology, such that for all Riemannian metric with , and for any smooth map , it has and it is homotopic to the identity, we have
Moreover, the equality holds if and only if is an isometry between and .
Let be the minimal surface in with respect to defined in Lemma 4.5. The inverse is also a closed surface in , but note that is not necessarily homotopic to . In fact, we can only find the following relation of their genus. The Gromov norms of and satisfies that
Here . And for any closed surface with genus , , where is a fixed number representing the supreme area of geodesic 2-simplices in . As an immediate result, we have .
To compare the areas of surfaces with respect to the induced metric of 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 , let be the covering space of such that . Let be the corresponding lift of that maps to . The lift of in still has fundamental group , so we denote it by as well. By assumption, is homotopic to identity, thus there is a continuous map with and for any . Since is compact, the length of the path of between and is uniformly bounded by a constant . Now let be the lift of that connects to the identity map on . For all , the length of the path between and is therefore uniformly bounded by the same constant . So is proper, meaning is a closed set, and therefore it is a -fold cover of for some finite number . If , the image of under is either a closed surface with Euler characteristic equal to , and therefore having genus , or it is a union of at least two surfaces with genus . However, both cases are impossible because the image cannot be identified with . We must have . Consequently, the covering map from to is one-to-one.
On the other hand, the classical result [21] verifies the existence of area-minimizing surface in the homotopy class of . And based on Theorem 4.3 of [18], there exist a -neighborhood of and , so that when and , is the unique minimal surface in homotopic to . Furthermore, let () be the lifts of (, respectively) in . These discs and are asymptotic and at a uniformly bounded Hausdorff distance to each other, as , converges uniformly on compact sets to in . Therefore, replacing by a smaller subset or replacing by a larger integer if necessary, we can assume that if and , then there exists a smooth map on with , such that can be represented by a graph of over . More precisely, Let be the unit normal vector field of , then we have the following diffeomorphism.
Notice that the minimal disc has mean curvature equal to zero with respect to , so the mean curvature with respect to has a uniform bound determined by the perturbation of and . Since is -close to , we have
(5.1) |
According to the Schauder estimates for elliptic PDE, there exists a constant , such that for any ,
(5.2) |
Besides, for , suppose the principle curvature of with respect to satisfies that . Uhlenbeck ([24]) shows that is foliated by a sequence of equidistant discs relative to . We denote by the disc with a fix distance to , it has mean curvature
Let and be the supremum and infimum of such that meets , and the intersections points are , , respectively. Since as , we may assume
then we have
Since , described by the graph , is bounded between and , the above result indicates the existence of a uniform constant , such that
(5.3) |
Combining (5.1)-(5.3), we obtain
And therefore, the principal curvatures of with respect to and satisfy that
Clearly, the sectional curvature of has the property
Thus, we can find and , such that if and , then the principal curvatures of with respect to and the sectional curvature of satisfy that
In the lemma below, we apply Uhlenbeck’s method [24], as well as the comparison result associated with Riccati equations, to prove that is the unique closed minimal surface in , thus minimizing the area among all closed surfaces.
Lemma 5.1.
Let be a minimal surface in whose fundamental group injectively includes in , the principal curvatures of with respect to , denoted by , and the sectional curvature of satisfy that
(5.4) |
Let be the cover of with . Then is the unique closed minimal surface in .
Proof.
Denote the supreme of the sectional curvature on by , where . Let be the distance function from a fixed point in to . By (5.4), for any ,
(5.5) |
It turns out that is a convex function, thus, there’s only one critical point that attains the minimum. As a result, maps injectively from the normal bundle to .
Furthermore, we show that is a diffeomorphism, and thus is foliated by a family of surfaces , where is the surface at the fixed distance to . To see this, we introduce some notations beforehand. For , choose an oriented, orthonormal basis for , and a unit vector for . Then we obtain an orthonormal frame by applying parallel transport along . Since is a minimal surface, the principal curvatures satisfy that , we assume in the following computation. Let be the Jacobi field along , where , it satisfies that , .
On the other hand, let be a minimal surface in with respect to an ambient metric of constant sectional curvature , and its principal curvature satisfies that . We do not require the existence of , it’s only used for comparison in the computation. Similar to the notations defined above, let be the corresponding frame on with respect to , and let be the Jacobi field along which shares the same initial data with . Since , we have
From
and the initial data, the graph of lies above that of , thus above the horizontal axis. The non-vanishing Jacobi fields ensure that the induced metric on in is nonsingular for all . In addition, we’ve seen that is injective, and therefore also bijective, so it is a diffeomorphism and admits a foliation structure.
Next, let be the principal curvatures on , and denote by the principal curvatures of the -equidistant surface to with respect to . Notice that each satisfies the Riccati equation
Then it follows from the comparison theorem associated with Riccati equations (for instance, see Theorem 3.1 of [26]) that
It follows from that
Therefore, for any , is strictly mean convex with respect to the metric induced by .
Finally, we prove the uniqueness. Assume that is another closed minimal surface in , and let and be the supremum and infimum of such that intersects , respectively, then and are both finite. However, due to the maximum principle, cannot be tangential to any strictly mean convex slice with . Therefore, we must have . ∎
Now we finish the proof of Theorem 1.4. From the previous lemma, when is sufficiently large, is the area-minimizer among all closed surfaces in with respect to the induced metric of , it yields that
Combining it with the area comparison in Theorem 5.1 of [19], we have
Moreover, when the equality holds, it follows from the equality of Theorem 5.1 of [19] that , is an isometry between and .
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
For any , we can take sufficiently large, such that
Let be the -cover of . Since has genus ,
then the least area surface in the homotopy class of with respect to satisfies that
(6.1) |
According to Müller-Puchta’s formula (see [15]), there exists a constant that depends only on and , such that the following is true when is large.
(6.2) |
Define in the following way
(6.3) |
Combining (6.2) with (6.1), we have that
Therefore, by (6.3),
Since is an arbitrarily small positive number, we conclude that
(6.4) |
6.2. Proof of rigidity
If , then (6.4) yields that
(6.5) |
To make use of this equality, we run the mean curvature flow in with initial condition , which is the lift of in , 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 and depending only on , so that for sufficiently large , we can find a solution to the mean curvature flow in with initial condition , where . Additionally, the mean curvature of and its derivative are both bounded uniformly by .
Proof.
Notice that after passing to a subsequence, converges smoothly on compact sets to a disc , and each of them is a cover of a compact surface in . Take , the standard theory indicates that there is a number , such that for any , we can find a solution to the mean curvature flow with initial condition , where . Since depends only on the second fundamental form of , in particular, it’s independent of and .
Next, in order to apply the Arzela-Ascoli theorem and estimate the mean curvature of and its derivative for any small time , we need the following preparation. Claim that for any , and any spacetime of , there exists an open neighborhood of , so that the Guassian density ratio
satisfies that
(6.6) |
If this wasn’t true for some integers and , then we could pick a sequence as , and
where denotes the parabolic dilation . Since the second fundamental form satisfies as , converges smoothly to a disc whose second fundamental form vanishes. However, the inequality above implies that
which contradicts the topology of .
We’ve seen that (6.6) holds, so due to the local regularity theorem in [27], there is a uniform constant that is independent of and , so that at any spacetime ,
(6.7) |
Therefore, Arzela-Ascoli theorem (see page 1494 of [27]) implies the short-time existence of the mean curvature flow with noncompact initial condition on time interval . 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 and 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.
where denotes the mean curvature of each disc in .
Proof.
Suppose by contradiction that there exists , such that after passing to a subsequence, and for large enough,
(6.8) |
Under the mean curvature flow, the mean curvature satisfies the following evolution equation on the time interval (see [23]).
where , and represent the mean curvature, second fundamental form and the immersion , respectively.
Using the result of Lemma 6.1, we can pick a uniform constant , such that for all sufficiently large , and for any ,
the latter inequality follows from the fact
Furthermore, arguing like Lemma 4.2 of [13], we deduce the following result from Lemma 6.2. For any round circle , it has a dense -orbit in . In addition, can be represented by , where , and represents the limit set of . Redefine by the lifts of to preserved by . It has the property that
Note that after passing to a subsequence, converges to the totally geodesic disc that is asymptotic to . Therefore, the mean curvature vanishes on , namely, is a minimal disc of with respect to the metric . And since is chosen arbitrarily, every totally geodesic disc of must be minimal for .
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 is minimal with respect to another metric if and only if for any geodesic , the following function is a constant
Because of the ergodicity of the geodesic flow in , we can choose a geodesic of whose orbit is dense in the unit tangent bundle. Let be the lift of to . must be contained in a hyperbolic 3-ball . Applying the previous lemma to the geodesic and ambient manifold , we conclude that
is constant. So the projection in also satisfies that
is constant. Thus due to the density, there is a constant , such that for any vector field of the unit tangent bundle of ,
As a result, coincides with a multiple of .
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 -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.