Entropy and stability of hyperbolic manifolds
Abstract.
Let be a closed oriented hyperbolic manifold of dimension at least . By the volume entropy inequality of G. Besson, G. Courtois and S. Gallot, for any Riemannian metric on with same volume as , its volume entropy satisfies with equality only when is isometric to . We show that the hyperbolic metric is stable in the following sense: if is a sequence of Riemaniann metrics on of same volume as and if converges to , then there are smooth subsets such that both and tend to , and converges to in the measured Gromov-Hausdorff topology. The proof relies on showing that any spherical Plateau solution for is intrinsically isomorphic to .
Introduction
Let be a hyperbolic manifold of dimension at least with hyperbolic metric . If is a Riemannian metric on , let denote its volume entropy:
where denotes the geodesic -ball centered at some point in the universal cover of . The fundamental volume entropy inequality, proved by Besson-Courtois-Gallot in [BCG95, BCG96], asserts that for any Riemannian metric on of same volume as , we have
(1) |
Moreover, Besson-Courtois-Gallot showed that this inequality is rigid in the sense that if equality holds, then is isometric to . How stable is the volume entropy inequality? We find that stability holds after removing negligible subsets:
Theorem 0.1.
Let be a closed oriented hyperbolic manifold of dimension at least . Let be a sequence of Riemannian metrics on with . If
then there is a sequence of smooth subsets such that
and converges to in the measured Gromov-Hausdorff topology.
In the statement of Theorem 0.1, is the metric space where the distance between two points is given by the infimum of the -lengths of curves joining to inside . A sequence of manifolds converges in the measured Gromov-Hausdorff topology if it converges both in the Gromov-Hausdorff and Gromov-Prokhorov topologies (for a definition of those topologies, see [V+09, Chapter 27, page 778]). Gromov-Prokhorov convergence implies . On the other hand, the conclusion that is a strong additional property.
It is elementary to see that naive stability for the Gromov-Hausdorff topology does not hold. Indeed, by adding thin and long threads to the hyperbolic metric , we get a new metric whose volume and volume entropy are arbitrarily close to and respectively. In this example, is far from in the Gromov-Hausdorff topology, although it is still close to in the Gromov-Prokhorov topology. The following question remains open: under the assumptions of Theorem 0.1, does converge to in the Gromov-Prokhorov topology? In Remark 3.9, we discuss the optimality of Theorem 0.1 with a notion of “coarse dimension” for Riemannian manifolds.
Historical comments
The question of stability for the volume entropy was raised by Courtois in [Cou98], and variants of this problem have been previously studied by Bessières-Besson-Courtois-Gallot [BBCG12] under a lower bound on the Ricci curvature (see also [LW11]), by Guillarmou-Lefeuvre [GL19] and Guillarmou-Knieper-Lefeuvre [GKL22] for neighborhoods of negatively curved manifolds, and Butt [But22] assuming uniform negative curvature bounds. We note that the differential rigidity result of [BBCG12] should follow from Theorem 0.1 and the theory of Cheeger-Colding [CC97, Theorem A.1.12].
The stability of geometric inequalities for Riemannian manifolds is a theme that has been extensively studied. We emphasize that in Theorem 0.1, no a priori curvature bound is required. The proof of this result has thus a quite different flavor compared to stability results under curvature bounds. Theorem 0.1 provides a stability result after removing “negligible” subset. This is formally similar to a stability result we recently proved with Conghan Dong for the Positive Mass theorem [DS23], which settles a conjecture of Huisken-Ilmanen. For stability results in the context of curvature bounds, see [Col96b, Col96a, Pet99, Aub05, CRX19, CDNZ+21]… for Ricci curvature, see [LS14, HLS17, S+21, LNN20, All21, CL22, DS23]… for scalar curvature. For spectral isoperimetric inequalities on surfaces, see [KNPS21] and references therein.
Main ingredients
The first main input in the proof of Theorem 0.1 is the theory of integral currents in metric spaces from geometric measure theory [AK00a, Lan11, Wen11, SW11] In particular, we make essential use of a compactness theorem due to Wenger [Wen11] which is formulated in terms of the intrinsic flat topology for integral current spaces [SW11]. With some hindsight, revisiting Besson-Courtois-Gallot’s original work using tools from geometric measure theory is especially natural, which is one of the main points of this paper. For instance, this combination leads directly to the “spherical Plateau problem” described in the next subsection, which enjoys rigidity properties at least as strong as for the minimal volume entropy problem.
The second ingredient in the proof of Theorem 0.1 is a sharp comparison result for the volume entropy of manifolds almost metrically dominated by a closed hyperbolic manifold, Theorem 3.5. Its proof relies on the equidistribution of geodesic spheres in closed hyperbolic manifolds. As a side note, together with Demetre Kazaras and Kai Xu, we recently applied this comparison result together with a “drawstring” construction to provide counterexamples to a conjecture of Agol-Storm-Thurston relating scalar curvature and volume entropy [KSX23].
The spherical Plateau problem
The proof of Theorem 0.1 is closely related to a variational problem in infinite dimension, called the spherical Plateau problem. Let be a closed oriented hyperbolic manifold with its hyperbolic metric and let . Consider the unit sphere in the Hilbert space and let act on by the regular representation . Denote by the corresponding quotient manifold, endowed with the standard round Hilbert Riemannian metric . There is a unique homotopy class of smooth immersions from to inducing an isomorphism on the fundamental groups. Besson-Courtois-Gallot define the spherical volume of [BCG91] as follows
A key step in Besson-Courtois-Gallot’s proof [BCG95, BCG96] of the entropy inequality (1) is to establish that
That result led us to consider in [Son23] the corresponding volume minimization problem, in particular the study of “limits of minimizing sequences”. Consider any minimizing sequence of maps , namely a sequence such that
Then by Wenger’s compactness theorem [Wen11], the images subsequentially converge as integral current spaces to an integral current space
in the intrinsic flat topology, in the sense of Sormani-Wenger [SW11]. Here is a metric space, is an integral current in the completion of , see Subsection 1.1. We call any such limit a spherical Plateau solution for .
Our second main theorem concerns the intrinsic uniqueness of spherical Plateau solutions for hyperbolic manifolds. The notion of “intrinsic isomorphism” between two integral current spaces will be defined in Definition 2.7.
Theorem 0.2.
If is a closed oriented hyperbolic manifold of dimension , then any spherical Plateau solution for is intrinsically isomorphic to .
Theorem 0.2 leads to a rigidity result with a representation theoretic flavor for , see [Son23, Corollary 4.3]. Conjecturally, the spherical Plateau solution for a closed oriented hyperbolic manifold is unique [Son23, Question 8]. The spherical Plateau problem is of independent geometric interest: in [Son23], we sketch the proof of the intrinsic uniqueness of spherical Plateau solutions for all oriented closed -manifolds, and the construction of higher dimensional analogues of hyperbolic Dehn fillings. Strictly speaking, the statement of Theorem 0.2 is not necessary to show Theorem 0.1. However, the methods in its proof do play a central role.
Outline of the proofs
For Theorem 0.2: In order to describe the proof, it is helpful to recall how Besson-Courtois-Gallot were able to compute the spherical volume
Their main tool was the barycenter map . In our setting, this is a Lipschitz map which under some technical conditions sends cycles in of the form , where , to the rescaled hyperbolic manifold with topological degree . Roughly speaking, the Jacobian of restriction of the barycenter map satisfies [BCG95, BCG96]
(2) |
which in particular implies that . Then the opposite inequality is checked by finding an explicit sequence of embeddings such that
In order to show that spherical Plateau solutions are unique up to intrinsic isomorphism, we try to argue as follows. Consider a minimizing sequence of maps , and denote by the integral currents of induced by pushing forward the fundamental class of by . The barycenter map enjoys the Jacobian bound (2) which is almost achieved on a region that covers almost all of as . Nontrivially, this implies a local Lipschitz bound for , which holds on a whole neighborhood of , and the differential of at points of can be shown to be close to a linear isometry. We can assume, by Wenger’s compactness theorem, that converges to a spherical Plateau solution
(the fact that such a limit exists is crucial). We then construct a limit map from the support of to :
sending the current structure to the natural current structure supported on . Heuristically, as goes to infinity, the Jacobian bound (2) for should be almost saturated almost everywhere, which means that the differential of should be close to a linear isometry almost everywhere. In other words, are almost Riemannian isometries. Passing to the limit, we should be able to deduce that is an isometry for the intrinsic metrics, which would essentially conclude the proof. This strategy of constructing a limit barycenter map has been exploited in the rigidity theorems of [BCG95, Proposition 7.1] and [BBCG12] where curvature bounds are assumed. There, the authors can argue that since their limit barycenter map is -Lipschitz and preserves the volume, it has to be an isometry, see [BCG95, Proposition C.1] and [BBCG12, Sections 3, 4, 5]. Related or more general “Lipschitz-volume” rigidity results were obtained in [DP23, Theorem 1.1], [BCS23, Theorem 1.1] and [Züs23, Theorem 1.2].
However, all those results depend either on the regularity of the convergence to the limit space outside of a small singular set, or on the -Lipschitz continuity of the limit map. The new challenge in our case is the lack of a priori regularity for spherical Plateau solutions and the fact that the limit map is never -Lipschitz in our situation (even though it will a posteriori follow that it is -Lipschitz for the intrinsic metric on ). To address this issue, we show in Proposition 1.4 that under some natural assumptions, limits of almost Riemannian isometries are Riemannian isometries. The proof uses a “curve lifting” argument, which in turn is based on an averaging argument involving the coarea formula.
For Theorem 0.1:
Consider a Riemannian metric on with same volume as and with entropy close to . Then, there is a uniformly Lipschitz map
which is almost a Riemannian isometry to its image, as observed by Besson-Courtois-Gallot [BCG91]. We apply again Proposition 1.4 as in the proof of Theorem 0.2 to instead of . We deduce that, for smooth subset ,
-
•
and are both small,
-
•
is close in the intrinsic flat topology to a space ,
-
•
there is a bi-Lipschitz, -Lipschitz map
-
•
is Gromov-Hausdorff close to via a topologically natural map.
The properties of are not as good as those of , so unlike Theorem 0.2, we cannot readily conclude that is an isometry for the intrinsic metrics. We need to remove a small subset from to get the Gromov-Hausdorff closeness property above.
In order to prove that the map above is, in fact, an isometry, we rely on a volume entropy comparison result, Theorem 3.5. The latter roughly says that if is naturally Gromov-Hausdorff close to a metric space and if there is a -Lipschitz map from to , then either is an isometry or the volume entropy of is strictly larger than . To show this, we make use of the equidistribution of geodesic spheres in the unit tangent bundles of closed hyperbolic manifolds, a result shown by Eskin-McMullen in [EM93].
Applying that comparison result to , we conclude that the map above is an isometry. This yields the intrinsic flat stability result, Theorem 3.7. We conclude the proof of Theorem 0.1 by applying a lemma of Portegies [Por15]: if a sequence of integral current spaces converges to a limit in the intrinsic flat topology without volume loss, then viewed as metric measure spaces the sequence converges to the limit in the Gromov-Prokhorov topology.
Organisation
Section 1 is about integral currents in metric spaces and maps between them. We prove a proposition answering in some cases the following question: given a sequence of uniformly Lipschitz, almost Riemannian isometries converging to a limit map, what can we say about that limit map?
In Section 2, we define the spherical Plateau problem for a closed oriented hyperbolic manifold. We introduce the barycenter map of Besson-Courtois-Gallot in our setting. Then we prove the intrinsic uniqueness of spherical Plateau solutions in Theorem 0.2.
In Section 3, we show a technical theorem whose proof is closely related to that of Theorem 0.2. We review an equidistribution result for geodesic spheres in the unit tangent bundle of hyperbolic manifolds, and how it implies a sharp comparaison theorem. Then, we apply the comparison theorem and the technical theorem to establish the volume entropy stability in terms of the intrinsic flat topology, which implies Theorem 0.1.
Acknowledgements
I am grateful to Gérard Besson, Gilles Courtois, Juan Souto, John Lott, Ursula Hamenstädt, Ben Lowe and Demetre Kazaras for insightful discussions during the writing of this article. I would especially like to thank Cosmin Manea, Hyun Chul Jang, Xingzhe Li and Dongming (Merrick) Hua for their careful reading, suggestions and for several corrections.
A.S. was partially supported by NSF grant DMS-2104254. This research was conducted during the period A.S. served as a Clay Research Fellow.
1. Limits of currents and limits of almost Riemannian isometries
1.1. Currents in metric spaces and Wenger’s compactness theorem
The theory of currents in metric spaces begins with works of De Giorgi, and Ambrosio-Kirchheim [AK00a]. It extends the theory of currents in finite dimensional manifolds due to De Giorgi, Federer-Fleming. For the most part, in this paper we will only stay in the standard framework of smooth maps and smooth manifolds. Nevertheless, a key reason for caring about metric currents is that this general theory enables to formulate powerful compactness results like Theorem 1.1 below. Besides, there is a profusion of standard tools (weak convergence, area/coarea formulae, slicing, push-forward…) for which the most natural language is given by geometric measure theory.
The main references we will need on the theory of metric currents are [AK00a, AK00b, Wen11, SW11]. We reviewed in some details the main definitions and results of the theory in Section 1 of [Son23]. In this paper, metrics on metric spaces assume only finite values. Integral currents in complete metric spaces are, roughly speaking, a countable union of Lipschitz push-forwards of Borel subsets in Euclidean spaces. They give a workable notion of “generalized oriented submanifolds” in complete metric spaces like Hilbert manifolds or Banach spaces. An -dimensional integral current has a well-defined notion of boundary which is an -dimensional integral current, a notion of volume measure denoted by and a notion of total volume called mass . Such a current is concentrated on a so-called canonical set , itself included in the support of the measure . The restriction of to a Borel set is denoted by , and its push-forward by a Lipschitz map is called . With those notations, See [AK00a, Section 3], see also [Son23, Subsections 1.1, 1.2] for a review.
The space of integral currents in a given complete metric space is endowed with the weak topology and flat topology, and the latter is finer than the former, see [Wen07, Subsection 1.1] [Son23, Subsection 1.3]. The mass is lower semicontinuous with respect to convergence in those topologies [AK00a].
The area formula expresses the mass of an integral current by its image under a Lipschitz map [AK00b, Section 8], [AK00a, Section 9], [Son23, Subsection 1.4]. The coarea formula, a kind of dual formula, expresses the mass of an integral current in terms of a double integral involving level sets of a Lipschitz map [AK00b, Section 9], [Son23, Subsection 1.4]. The slicing theorem is a kind of generalization of Sard’s theorem and tells us that almost all level sets of a Lipschitz map are integral currents themselves [AK00a, Theorems 5.6 and 5.7].
Following the notion of integral currents in complete metric spaces, one can define a more intrinsic notion of integral currents. That was achieved by Sormani-Wenger [SW11]. Basically an integral current space is a triple where is a metric space and is an integral current in the completion of , which we will usually denote by (one requires that is the “canonical set” of the current ) [Son23, Definition 1.3 and Subsection 1.1]. A simple example of integral current space is given by a closed, connected, oriented Riemannian -manifold : the metric space is endowed with the geodesic distance of , and the integral current structure is the natural integral current induced by the fundamental class .
There is also an intrinsic notion of flat topology, called intrinsic flat topology [SW11]. Similarly to the definition of Gromov-Hausdorff topology, two integral current spaces are close in the intrinsic flat topology whenever they can be isometrically embedded in a common complete metric space in which they are close in the usual flat topology [Son23, Definition 1.4] .
A key result is Wenger’s compactness theorem:
1.2. Limits of almost Riemannian isometries and intrinsic flat limit spaces
As usual, inside an -dimensional Riemannian manifold, we will denote by and the -dimensional and -dimensional Hausdorff measure. Sometimes, we also use to denote the -dimensional Hausdorff measure. Given a metric on a space, the standard notion of induced intrinsic metric is defined in [BBI22, Chapter 2, Section 2.3]. If is a Riemannian metric on a manifold , let be the metric on induced by . Sometimes we will make the identification
We will use a few times the following simple fact: if is a compact Riemannian -manifold with a piecewise smooth metric inducing , then for any metric whose induced intrinsic metric is , and any open subset , the mass of as an -dimensional current in is at most .
Lemma 1.2.
[Sor18, BCS23] Let be two complete metric spaces. Let be a sequence of integral currents in and let
be a sequence of -Lipschitz maps for some independent of . Suppose that (resp. ) converges in the flat topology to an integral current (resp. ) inside (resp. inside ), and that is compact.
Then there is a -Lipschitz map
such that:
-
(1)
after taking a subsequence if necessary, for any positive integer and any collection of points , there is a sequence of collections of points such that for each , as , converges to , and converges to ,
-
(2)
as currents inside .
Proof.
(1) is [Sor18, Theorem 6.1], and is proved using an Arzelà-Ascoli type argument.
(2) follows from a slight generalization of [BCS23, Lemma 7.3]. If is the Banach space of bounded real functions on endowed with the norm, then it is well-known that embeds isometrically inside by the Kuratowski embedding, and is an injective metric space in the following sense: given any other metric space , a subset , and a -Lipschitz map , there exists an extension of , called , which is still -Lipschitz. We can adapt the proof of [BCS23, Lemma 7.3] by using that extension theorem, instead of McShane’s extension theorem. ∎
For this subsection, we will make the following assumption.
Assumption 1.3.
Let be a connected, closed, oriented Riemannian -manifold. Let be a sequence of integral currents in a complete metric space , converging in the flat topology to an integral current inside . Suppose that
-
(a)
each support , endowed with the intrinsic metric induced by the metric , is a compact, oriented Riemannian manifold with a piecewise smooth metric (possibly with nonempty piecewise smooth boundary),
-
(b)
,
-
(c)
there is a sequence of maps
which are on the smooth part of and -Lipschitz for some independent of , such that converges to in the flat topology inside ,
-
(d)
there is a sequence of open subsets contained in the part of where is smooth, such that and ,
-
(e)
moreover, is almost a Riemannian isometry on in the sense that
where denotes any choice of orthonormal bases for the tangent spaces of .
Some of the conditions above are unnecessarily restrictive, but they will be convenient for our applications. Note that Lemma 1.2 applies under Assumption 1.3 and yields a limit map
The following proposition, while elementary, is technically important for us. It is related to, but different from Lipschitz-volume rigidity results like [BCG95, Proposition C.1], [BBCG12, Sections 3, 4, 5], [DP23, Theorem 1.1], [BCS23, Theorem 1.1] and [Züs23, Theorem 1.2].
Proposition 1.4.
Suppose that Assumption 1.3 above holds and let
be the limit map constructed in Lemma 1.2.
-
(1)
Then is a bi-Lipschitz bijection and its inverse is -Lipschitz with respect to the induced intrinsic metrics.
-
(2)
Suppose additionally that for any , there is such that if is large enough, then for any such that , we have
Then is an isometry with respect to the induced intrinsic metrics.
Remark 1.5.
The limit map in (1) is not -Lipschitz for the intrinsic metrics in general, which means that the additional condition in (2) is needed. Indeed consider for instance the standard round metric on the Euclidean unit sphere , and for each , consider the conformal metric where is outside the -neighborhood of the equator and in the -neighborhood of the equator. Let and be the corresponding intrinsic metrics. Then Assumption 1.3 is satisfied with being the identity map , etc. However, the intrinsic flat limit and Gromov-Hausdorff limit of are both determined by the length structure on induced by for curves not touching the equator, and with an equator of length instead of . The limit is still the identity map and it is not -Lipschitz for the intrinsic metrics.
Proof.
Property (2) follows directly from property (1) in the statement and Lemma 1.2 (1). Indeed, applying the additional assumption in (2) with arbitrarily small , together with Lemma 1.2 (1), we obtain that does not increase distances for the intrinsic metrics, in other words it is -Lipschitz for the intrinsic metrics. Since property (1) says that the inverse of is also -Lipschitz for the intrinsic metrics, it is an isometry.
It remains to prove property (1). Note that by Lemma 1.2 (1), is -Lipschitz. Let be the intrinsic metric on induced by the restricted metric (a priori is allowed to take as value). Note that by Assumption 1.3 (c) (d) (e), the area formula and the lower semincontinuity of mass under flat convergence, we can assume that is injective on without loss of generality by reducing that domain a bit.
For , set
Then for every ,
(3) |
Indeed, let us assume on the contrary that for some , . Then, by Assumption 1.3 (a) (d) (e), we should have
By a standard application of the slicing theorem, we can assume without loss of generality that the restricted current is an integral current converging to in the flat topology as . Thus still converges to in the flat topology. By Assumption 1.3 (c) and lower semicontinuity of the mass with respect to flat or weak convergence,
This contradicts the previous inequality and so (3) was true.
Given a Lipschitz curve in , let denote its length with respect to the metric . Next, it is convenient to show the following “curve lifting” property.
Curve lifting property: Let . Let and let
Then there exists a compact connected Lipschitz curve contained in , starting at , ending at , and moreover
Proof of the curve lifting property.
Let , be the metric balls in , of radius centered at . By Lemma 1.2 (1), if is chosen small enough then for all large and every (resp. ), we have
By lower semicontinuity of the mass, for each large,
for some depending on but independent of . For large, since we are assuming that is injective on without loss of generality, by Assumption 1.3 (d) (e) and the area formula, we have the following volume estimates:
(4) |
After applying the coarea formula, (3) and (4) as in the toy example, Example 1.6, at the the end of this subsection, we find, for each large enough, two points
and a smooth curve
with , joining to such that the restricted preimage
is a compact curve in avoiding , whose endpoints satisfies
Because and belong to (on which is assumed to be injective without loss of generality), in fact
(5) |
By Assumption 1.3 (c) (d) (e), the restriction of to the complement of has mass converging to as ; similarly, by (3), the restriction of to the complement of has mass converging to . Thus the coarea formula again implies that we could choose satisfying additionally:
(6) |
Together with Assumption 1.3 (e) and the area formula, these properties imply:
(7) |
where . By using (6) and the fact that (resp. ) is in (resp. ), we easily construct a new curve fully contained in joining to , with length at most for large. This proves the curve lifting property. ∎
The curve lifting property implies the following useful properties. Firstly, is compact. Suppose towards a contraction that is not compact, then for some , there is an infinite sequence of points such that those points are pairwise at distance at least in . By compactness of , for any there are such that
Then the curve lifting property implies that the distance between and is at most , a contradiction when .
Secondly is bijective. Indeed we verify that is injective by a direct application of the curve lifting property. Surjectivity follows from Lemma 1.2 (2) and the compactness of .
We are ready to prove property (1) of our proposition. Take two points and let . Let , . By applying the curve lifting property repeatedly and making , by compactness of we get a limit Lipschitz curve in joining to , with length at most . Thus the inverse is indeed -Lipschitz for the intrinsic metrics, and is bi-Lipschitz, as wanted.
∎
Below is a toy example illustrating how the standard coarea formula is applied in the proof of Proposition 1.4.
Example 1.6 (Toy example).
Let . Consider a sequence of compact Riemannian -manifolds with piecewise smooth metrics and with piecewise smooth boundaries. Suppose that there is a sequence of -Lipschitz smooth maps
and that for each , consider there are subsets so that
and for some independent of , for all :
Let be the projection on the last coordinates. By Fubini’s theorem, for each we can find a vector such that if we set
then we have
for some independent of . By composing each with a diffeomorphism with uniformly bounded Lipschitz constant if necessary, we can assume that . Applying the coarea formula and Sard’s theorem twice, first to the -Lipschitz maps , then to the map , we find for each some straight segment joining to , such that is a point in and such that
is a compact smooth curve (with possibly several connected components) which avoids for all large enough:
This example generalizes when replacing with a manifold.
2. The spherical Plateau problem for hyperbolic manifolds
2.1. The spherical Plateau problem
Let us define the spherical Plateau problem for closed oriented hyperbolic manifolds, which is part of a more general framework [Son23, Section 3]. Let be a closed oriented hyperbolic manifold, whose fundamental group is denoted by . Let be the unit sphere in . The -norm induces a Hilbert Riemannian metric on . The group acts isometrically on by the (left) regular representation : for all , , ,
Since is infinite and torsion-free, acts properly and freely on the infinite dimensional sphere . The quotient space is topologically a classifying space for . It is also a Hilbert manifold endowed with the induced Hilbert Riemannian metric . The diameter of is bounded from above by .
Given base points , , there is a smooth immersion inducing the identity map from to , which is unique up to homotopies sending to . Other choices of yield homotopic maps, so that determines a unique homotopy class of maps which we call “admissible”. Set
Any map defines the pull-back Riemannian metric on .
Besson-Courtois-Gallot introduced the spherical volume of in [BCG91]. It can be equivalently be defined as follows.
Definition 2.1.
The spherical volume of is defined as
The spherical volume of closed oriented hyperbolic manifolds was computed by Besson-Courtois-Gallot. See [Son23, Theorem 4.1] for the proof, adapted to our setting.
A sequence is said to be minimizing if
Denote by the integral current in induced by and its orientation. For a Lipschitz map , recall that denotes the push-forward integral current in . We can now define spherical Plateau solutions.
Definition 2.3.
We call spherical Plateau solution for any -dimensional integral current space which is the limit in the intrinsic flat topology of a sequence where is a minimizing sequence.
For any sequence such that
the mass and diameter of are uniformly bounded, so by Wenger’s compactness (Theorem 1.1) there is a subsequence of converging in the intrinsic flat topology. The need for an abstract compactness result like Theorem 1.1 is explained in [Son23, Remark 3.3].
Remark 2.4.
While for our present purpose, it is enough to consider the set of admissible smooth immersions from to , we believe that it is more natural to formulate the general spherical Plateau problem in terms of integral currents with compact support in representing a homology class . This is the point of view presented in [Son23, Section 3]. In fact, by [Bru08] and a standard polyhedral approximation result for integral currents in Hilbert manifolds [Son23, Lemma 1.6], it is possible to prove that these two setups lead to the same notions of spherical volume and spherical Plateau solutions, at least when the countable group is torsion-free.
2.2. The barycenter map and the Jacobian bound
The barycenter map played a crucial role in the work of Besson-Courtois-Gallot on the volume entropy inequality [BCG95, BCG96] (see also [BCG99, Sam99, CF03, Sou08] for a small sample of other uses of the barycenter map).
For the reader’s convenience, all the main properties of the barycenter map are proved in our setting in [Son23, Section 2] and the main Jacobian bound is recalled below. We choose to express the barycenter map using the -space on a group, instead of the -space on a boundary as in [BCG95]. The advantage is that only a minimal amount of knowledge is needed, and that it extends directly to other more general situations (3-manifolds, connected sums, Plateau Dehn fillings, see [Son23, Sections 4, 5, 6]).
Let be a closed oriented hyperbolic manifold. Let be its universal cover, namely the hyperbolic -space. Let . The latter acts properly cocompactly and freely on . Let be the unit sphere in the Hilbert space , on which acts freely and properly by isometries via the regular representation, so that is a smooth Hilbert manifold endowed with the standard round metric (see Subsection 2.1).
Set
where is a positive constant. When we fix large enough, the following holds: for any , the composition
is smooth everywhere and satisfies
(9) |
Definition 2.5.
Fix a basepoint . Let be the set of functions in with finite support. For , consider the functional
(10) |
The barycenter map is then defined as
The barycenter map is well-defined: the modified distance functions are strictly convex, moreover tends to infinity uniformly as , so that the point where attains its minimum exists and is unique. The subset is invariant by , and is -equivariant. The quotient map is also denoted by . For more details, see [Son23, Section 2].
We will avoid discussing regularity issues for the barycenter map by only considering its restriction to the supports of “polyhedral chains”, which will be enough in all our applications. A -dimensional polyhedral chain in is by definition a -dimensional integral current such that there are smoothly embedded totally geodesic -simplices endowed with an orientation, and integers so that
(see [Son23, Subsection 1.7]). Given a polyhedral chain in , we can check that the restriction
is indeed continuous and smooth on each simplex. For , given a smooth embedding with totally geodesic image , let be the tangent -plane at for some . The map
is smooth around , and its differential along is denoted by . For more details on those claims, see [Son23, Susbection 2.2].
The main result in this Subsection is the following (see [Son23, Lemma 2.4] for a proof):
Lemma 2.6.
[BCG95] Suppose that . Let and let be the tangent -plane at of a totally geodesic -simplex in passing through . Then
(11) |
Moreover for any small enough, there exists with , such that the following holds. If
then for any norm tangent vector ,
(12) |
and for any connected continuous piecewise geodesic curve of length less than starting at , we have
(13) |
where is computed using the standard round metric on .
2.3. Intrinsic uniqueness for hyperbolic manifolds
From a geometric point of view, a natural question is the uniqueness of spherical Plateau solutions for closed hyperbolic manifolds. We do not know if uniqueness holds, however we will prove uniqueness up to “intrinsic isomorphism”.
Consider an integral current space and an oriented, connected, closed Riemannian manifold , which induces the integral current space . The intrinsic metric on induced by is denoted by . Note that the identity map
is always -Lipschitz (on each path connected component).
Definition 2.7.
We say that is intrinsically isomorphic to if there is an isometry
such that
For clarity, we emphasize that “being intrinsically isomorphic” is weaker than “being at intrinsic flat distance from each other”.
Our main result in this section shows that in dimensions at least , the spherical Plateau solutions for closed hyperbolic manifolds are unique up to intrinsic isomorphism, see Definition 2.7.
Theorem 2.8.
Let be a closed oriented hyperbolic manifold of dimension at least . Then any spherical Plateau solution for is intrinsically isomorphic to .
Proof.
Let be a minimizing sequence, namely
(14) |
where the second equality follows from Theorem 2.2. We suppose that the integral currents
converge in the intrinsic flat topology to a spherical Plateau solution
Set . By a perturbation argument, we can assume without loss of generality that for all , for all , any lift of in has finite support. In particular, we can assume that
where is defined in Subsection 2.2. By a further perturbation of , we can even assume that is a polyhedral chain (a notion defined in Subsection 2.2), in particular that is a finite union of embedded totally geodesic -simplices in , see [Son23, Lemma 1.6].
From now on, we will use the notation
In the sequel, Jacobians, lengths and distances will be computed with respect to the metric on . Fix and let
be the barycenter map, see Section 2.2. By -equivariance, for any , is a Lipschitz homotopy equivalence, and
(15) |
By lower semicontinuity of the mass under intrinsic flat convergence [SW11]:
(16) |
(the equality above is Theorem 2.2).
The -dimensional Jacobian of along the tangent -plane of at any point in the interior of a “face” of is well-defined and is bounded from above by with respect to the metric on , by the main Jacobian bound (11) in Lemma 2.6. This implies by the area formula and (15) that
Since has mass converging to , by the area formula, the Jacobian of has to be close to on a larger and larger part of as , meaning that there are open subsets such that at every point , there is a well-defined tangent -plane of at , and
(17) |
where we recall that the Jacobian is computed with and denotes the Jacobian along the tangent -plane, see Section 1.1. For , set
By (17), the coarea formula and Sard’s theorem, after smoothing out the distance function from by a standard argument and still using the notation “” for the -sublevel set of the smoothed out distance function, there are such that for each ,
is an integral current, and is a compact piecewise smooth submanifold of satisfying the following:
-
•
the boundary of is piecewise smooth (this is where considering the smoothed out distance function is used) and we have
(18) -
•
after taking a subsequence, still converges to
in the intrinsic flat topology as , In particular, there are a Banach space and isometric embeddings
(with a slight abuse of notations we consider those sets as subsets of ), such that converges to in the flat topology inside .
Inequality (13) of Lemma 2.6 ensures that a Lipschitz bound holds uniformly in a neighborhood of : for any , there is , such that if is large enough, then for and joined to by a piecewise geodesic curve of length at most , we have
(20) |
Given and a curve in joining those two elements, after a small perturbation, that curve can be assumed to be inside . As a consequence of (20), we get the following local Lipschitz bound: for any , the restriction of to the subset is -Lipschitz for some independent of . In particular, the restriction
(21) |
We can now check that Assumption 1.3 is verified with , , , , , . In particular, in order to check Assumption 1.3 (e), observe that since the Jacobian of converges to on by (17), is forced to be almost a Riemannian isometry on by (12), (13) in Lemma 2.6. Furthermore, the additional assumption in Proposition 1.4 (2) is also satisfied by (20).
By Proposition 1.4 (2), we immediately conclude that there is a limit map which is an isometry for the intrinsic metrics induced on and . Moreover by Lemma 1.2 (2), preserves the current structures in the sense that
In other words, is intrinsically isomorphic to , as wanted.
∎
3. The entropy stability theorem
3.1. Technical preparation
As before, is the closed, connected, oriented hyperbolic manifold, is its fundamental group and is the unit sphere inside , which is acted upon by via the regular representation.
Let us define maps relating the volume entropy of a Riemannian metric on and the spherical volume of , introduced by Besson-Courtois-Gallot, see [BCG91, Proof of Lemma 3.1]. Let be a Riemannian metric on . The universal cover of is and its fundamental group is . Let be its volume entropy. Denote by a Borel fundamental domain in for the action of and let be its image by an element . Besson-Courtois-Gallot considered for maps similar to the following:
Those maps satisfy the following properties, which hold in any dimension :
Lemma 3.1 ([BCG91]).
For a Riemannian metric on , is a -equivariant Lipschitz map, and for almost any , it satisfies
(22) |
where is a -orthonormal basis of .
Proof.
For the reader’s convenience, let us outline the proof. Consider the unit sphere in . Set for :
and set
These maps are manifestly -equivariant, and note that . One easily checks that is -Lipschitz. To prove the lemma, it remains to study . By the Pythagorean theorem,
Taking the trace and using that the norm of the gradient of the distance function is well-defined almost everywhere and equal to , we get at almost every , in a -orthonormal basis of :
This proves the lemma.
∎
If is a Riemannian metric on , let be the geodesic distance on induced by . The definition of the standard notions of -isometry, -net can be found in [BBI22, Definition 7.3.27, Definition 1.6.1]. Given subset of a Riemannian manifold , denotes (by a slight abuse of notation) the intrinsic metric induced by the Riemannian metric using paths inside . In general is very different from , where is the restriction of the induced metric of to .
The set of admissible maps was defined in Subsection 2.1. The barycenter map was defined in Subsection 2.2. The following result is an intermediate step towards Theorem 0.1, and its proof is parallel to that of Theorem 2.8 but more technical.
Theorem 3.2.
Let be a closed oriented hyperbolic manifold of dimension . Let be a sequence of Riemannian metrics on of same volume as , and suppose that
Then, there are smooth open subsets such that the following holds after taking a subsequence:
-
(1)
and ,
-
(2)
converges in the intrinsic flat topology to an integral current space
-
(3)
there is a bi-Lipschitz bijection
which is -Lipschitz, and
-
(4)
converges to in the Gromov-Hausdorff topology. Moreover, for any , for all large enough, there is a homotopy equivalence
such that the restriction is an -isometry.
Proof.
Step 1: Finding good subsets
For technical convenience, set
Note that after rescaling,
By our assumptions, there is a sequence of positive numbers such that
(23) |
By Lemma 3.1, the maps
are -equivariant. The quotient maps
can be perturbed to be smooth immersions. Those new maps now belong to . After a further small perturbation, we obtain homotopic smooth immersions
sending inside , see [Son23, Lemma 1.6]. Moreover, by (22) and (23), it is not hard to ensure that after those standard smoothings, for all :
(24) |
for some positive (with respect to ), where is an orthonormal basis for . By (24) and the inequality of arithmetic and geometric means,
(25) |
where the Jacobian is computed with respect to . By Theorem 2.2,
on the other hand we have by assumption. Hence, by (25), converges to on an open region with
which by (24) forces
(26) |
where is the standard Hilbert Riemannian metric on the spherical quotient , and denotes any choice of orthonormal bases for the tangent spaces of .
Exactly as in the proof of Theorem 2.8 and using (26), we first find open subsets with
(27) |
which satisfy
Then we define smoothings of -neighborhoods of in , called , so that the closure of is a compact manifold with a smooth boundary whose area goes to as , and the restriction is uniformly Lipschitz.
Step 2: Constructing the limit map
We set
In order to apply Wenger’s compactness theorem, we need a uniform diameter bound. For that reason, if denotes the intrinsic metric induced by using paths contained in , we set
This defines a metric on with diameter at most , and it is locally isometric to the induced intrinsic metric . We then set
By Wenger’s compactness theorem, the integral current spaces converge to an integral current space
in the intrinsic flat topology, after picking a subsequence if necessary. In particular, there are a Banach space , and isometric embeddings
(28) |
with the usual abuse of notations, such that converges to in the flat topology inside .
Next, we check that Assumption 1.3 is satisfied for
(Note however that the additional condition of Proposition 1.4 (2) is a priori not satisfied, which accounts for the difference between the statements of Theorem 2.8 and Theorem 3.2.) Thus by Proposition 1.4 (1), there is a limit map
which is Lipschitz, bijective and whose inverse
is -Lipschitz with respect to the intrinsic metrics. Hence, is clearly -Lipschitz and bi-Lipschitz. By Lemma 1.2 (2), .
Step 3: Convergence for the original induced metric
We also need to check that , not just , subsequentially converges to the integral current space . Notice that for any and , the metric balls and are globally isometric. In particular, since has diameter at most that of by -Lipschitzness of , if denotes the -neighborhood of in , we have: whenever , for any and pair of points ,
(29) |
By the slicing theorem for metric currents, we can choose for each , some radius converging to , such that if we set
then are integral currents in converging to in the flat topology. By (29), this means that converges to in the intrinsic flat topology. We deduce in particular that the push-forward of by converges to as currents in . Then the liminf as of the mass of this push-forward is at least by lower semicontinuity of the mass. By the Jacobian bounds (25), (11), and since by (27) we have
(30) |
We conclude that converges to the same limit as in the intrinsic flat topology, which is , as desired.
Step 4: Gromov-Hausdorff convergence and -isometries
In general, does not converge in the Gromov-Hausdorff topology to . The end of the proof is about fixing this issue. By (28), there are finite subsets converging in the Hausdorff topology to in . For any , let
By lower semicontinuity of the mass and (28), for any and any sequence of points ,
(31) |
for some not depending on . We also have the following stronger property: for any and any sequence of points ,
(32) |
for some not depending on . Note that this is indeed a stronger inequality, since is the intrinsic metric on induced by using paths inside , and
To check this stronger property, recall that has been shown to be bi-Lipschitz to the closed Riemannian manifold via a map . For any two points , we can find approximating . Then, given a minimizing geodesic segment in between , we can approximate by a curve in between without increasing the length by more than a constant factor. Hence, for large,
By (30) and by the coarea formula, there is a , arbitrarily small, such that
This means that after taking a subsequence, we find converging to so that if we set
then for any and any sequence of points ,
(33) |
Now we can reapply all the arguments in Step 2 and Step 3 to a smoothing of instead of . Let us summarize what we have achieved so far: subsequentially, converges to an integral current space
and there are a Banach space , and isometric embeddings
(34) |
with the usual abuse of notations, such that converges to in the flat topology inside . Moreover, there is a bi-Lipschitz, -Lipschitz map
which is the inverse of a limit map constructed using Lemma 1.2 applied to . The following analogue of (30) holds: for any , if is the -neighborhood of in ,
(35) |
The key additional property we gained is that now converges to in the Hausdorff topology inside , by (33) and (35). Note that in general, and the previous space could be very different.
We can then set
which is a homotopy equivalence. By Lemma 1.2 (1), we conclude that for any ,
is an -isometry if is large. All of these complete the proof, after rescaling all the Riemannian metrics by . ∎
3.2. Equidistribution of geodesic spheres in hyperbolic manifolds
Consider a closed hyperbolic manifold, with universal cover . Fix and let be a lift of by the natural projection . Let denote the unit tangent bundle of . Let be the geodesic sphere of radius centered at in , and let be its lift to the unit tangent bundle by considering the outward unit normal vectors on . Let denote the projection of in , and let be the projection of to the unit tangent bundle . A measure on (resp. ) is called invariant if it is induced by a measure on invariant by isometries of (resp. induced by a measure on invariant by rotations of center in ).
As a corollary of the mixing property for the geodesic flow on closed hyperbolic manifolds, the lift of geodesic spheres equidistribute in the unit tangent bundle. This is for instance explained in [EM93, Section 2] for surfaces and generalized in [EM93, Theorem 1.2] 111I thank Ben Lowe for pointing out this reference.. With the above notations, the statement is the following:
Theorem 3.3.
For any continuous function ,
where is the unique invariant probability measure on and is the unique invariant probability measure on .
Below, areas (namely -dimensional Hausdorff measures) and lengths are computed using . Given an open subset , let denote the relative homotopy group. Consider a (not necessarily length minimizing) geodesic segment in with two different endpoints and let be two disjoint open geodesic balls centered at and . Fix as before. Let
be the natural projection.
Corollary 3.4.
There is depending on such that for all large enough, there is an open subset satisfying
and with the following property: for any , if denotes the length minimizing geodesic from to in parametrized by arclength, there are disjoint intervals
such that
-
(1)
,
-
(2)
for , the endpoints satisfy and ,
-
(3)
for , is a geodesic segment joining to , which is in the same class as in .
Proof.
Let be the length of . By continuity, there exist an open subset of the unit tangent bundle depending only on , such that for any tangent vector in , the basepoint of lies in , and the geodesic starting at with direction and length ends at a point , and satisfies the following:
Informally, geodesics of length starting at a vector in stay “close” to .
Let be the invariant probability measure on and set
Recall that is the lift of the sphere by its normal unit vector. Below, by abuse of notations, we will identify and . By applying Theorem 3.3 to the characteristic function of , for all large enough,
(36) |
for some independent of . If , let be the unique geodesic segment from the basepoint to in .
We claim that for some , for any large integer ,
Roughly speaking, this inequality means that for a uniformly positive fraction of the sphere , geodesics from to that portion of stay close to on a uniformly positive fraction of their length. Before proving the claim, note that is a sphere parametrized by via the exponential map with basepoint , and that the measure on corresponding to is just the standard uniform probability measure . We let be the characteristic function of the subset corresponding to and we compute for any large :
where the first inequality follows from (36). So there are , for any large , on some subset of of -measure at least ,
which is exactly the claim. ∎
3.3. From equidistribution of geodesic spheres to intrinsic flat stability
Let be a closed oriented hyperbolic manifold of dimension . One of the main technical tools in this section is the following volume entropy comparison, which roughly speaking says that if a sequence of metrics on approximates a metric space which is metrically dominated by , then the volume entropy of is eventually strictly larger than that of . Its proof relies on the equidistribution of geodesic spheres in hyperbolic manifolds, Theorem 3.3.
Theorem 3.5.
Let be a closed oriented hyperbolic manifold of dimension . Suppose that the following holds:
-
(1)
there is a metric on such that there is a bi-Lipschitz bijection
which is -Lipschitz,
-
(2)
there are Riemannian metrics () on so that for any , for all large enough, there are an open subset , and a homotopy equivalence
such that the restriction is an -isometry.
Then, if is not an isometry, we have
Remark 3.6.
We emphasize that denotes the metric space whose metric is induced by using paths in . In particular, it is not in general isometric to , where is the restriction of to the subset .
Proof.
Consider small positive numbers to be fixed later, and consider large enough so that there is and a homotopy equivalence whose restriction
is an -isometry, as in condition (2).
Let us then define “-lifts”. Given a point , we say that is an -lift of if is -close to with respect to the hyperbolic metric . Given a -geodesic segment in with endpoints (which is parametrized by arclength), we say that a curve with endpoints in is a -lift of if
-
•
are -lifts of ,
-
•
-
•
where means -geodesic ball.
By basic properties of -isometries [BBI22, Exercise 7.5.11] and since the bi-Lipschitz bijection is -Lipschitz, for any , whenever is small enough compared to and the injectivity radius of , any -geodesic segment in admits an -lift in .
Suppose now that the -Lipschitz map is not an isometry, which just means that there are two distinct points so that
(37) |
Choose and accordingly , so that
(38) |
Let be -lifts of . To fix ideas, let us assume without loss of generality that and . By the -isometry and (38),
(39) |
Let be a -length minimizing segment which realizes the -distance between and . Consider the compact curve with endpoints , and let us minimize its length among all curves homotopic to with same endpoints. This yields a -geodesic segment
parametrized by arclength, with endpoints . Note that, since is a homotopy equivalence, any -lift of with endpoints (that can always be ensured) is in fact homotopic (with fixed endpoints) to inside .
By (39), by continuity and uniqueness properties for geodesic loops in hyperbolic manifolds, there are small disjoint open -geodesic balls containing respectively and some with the following property: for any geodesic segment such that , , and
we can find a corresponding -lift in and a curve homotopic (with fixed endpoints) to such that
(40) |
An important remark is that, since is bi-Lipschitz, and since is an -isometry, the -length of is uniformly bounded independently of . By compactness, we can assume without loss of generality that is fixed and does not depend on . For that reason, we will assume that only depend on but not on . The notion of -lifts of curves and their properties extend naturally to curves in the universal covers and .
Given a Riemannian metric on and a point , let be the collection of homotopy classes of loops with fixed basepoint which contain at least one loop based at of -length at most . It is well-known that the volume entropy of is:
where denotes the cardinality of a set. In particular, it does not depend on the choice of base point .
Fix a base point and a lift (here the “lift” belongs to the universal cover, it is not to be confused with the notion of -lift). By uniqueness of geodesic loops in homotopy classes of loops inside hyperbolic manifolds, we identify with the set of geodesic loops based at with length at most . Classically, the volume entropy of the hyperbolic -plane is , meaning that
(41) |
The crux of the proof is that the equidistribution of lifts of geodesic spheres to the unit tangent bundle plus the distance comparison inequality (40) force the volume entropy of to be strictly larger than .
For all large, fix an -lift of the basepoint inside , and a lift of in the universal cover. As we saw earlier, we assume without loss of generality that does not depend on . By Corollary 3.4, inequality (40) and the properties of -lifts, we deduce that there are some small and (this is where the latter are fixed) depending on but independent of , such that the following holds for all large. In the geodesic spheres of universal cover , for any large enough, there is an open subset such that
and for any and any large enough, the minimizing geodesic from to admits an -lift joining to an -lift of in which in turn is homotopic (with fixed endpoints) to a curve of -length at most . In colloquial terms, a uniform fraction of points in at -distance from admit -lifts in which are at -distance significantly less than from .
By basic hyperbolic geometry (volume of geodesic spheres and balls, etc.) and properties of -lifts, the previous paragraph implies that for large enough, for all large enough there is a small depending only on such that for all large enough,
-
•
there are distinct points which are lifts of to , and their number satisfies
(42) -
•
there are curves joining to respectively, and they admit -lifts in , which are respectively homotopic (with fixed endpoints) to curves of -lengths at most ,
-
•
each of the curves joins to some other lift of inside the universal cover .
We conclude from (41) and (42) that for any large enough, for all large:
In particular,
Since does not depend on , the proof is complete.
∎
We are now ready to finish the proof of the intrinsic flat stability theorem.
Theorem 3.7.
Let be a closed oriented hyperbolic manifold of dimension at least . Let be a sequence of Riemannian metrics on with . If
then there is a sequence of smooth subsets such that
and converges to in the intrinsic flat topology and Gromov-Hausdorff topology.
Proof.
Under the assumptions of the theorem, by combining Theorem 3.2 and Theorem 3.5, we deduce that subsequentially, there are open subsets such that if
then after renumbering,
-
•
-
•
converges in the intrinsic flat topology to an integral current space
-
•
converges to in the Gromov-Hausdorff topology,
-
•
and there is an isometric bijection
such that
In particular, is isomorphic as an integral current space to the hyperbolic manifold . Since this integral current space is the only possible subsequential limit, there are with and converging to , and converges to in the intrinsic flat and Gromov-Hausdorff topologies (without the need to take subsequences). ∎
Recall that intrinsic flat convergence implies weak convergence (see Section 1.1). Given a Riemannian metric on , the mass measure of the integral current space is equal to the usual volume measure on . The proof of Theorem 0.1 is then completed by combining Theorem 3.7 and the following general lemma proved by Portegies, which yields that weak convergence plus volume convergence implies Gromov-Prokhorov convergence for Riemanian manifolds:
Lemma 3.8.
[Por15, Lemma 2.1] Suppose is a complete metric space, and is a sequence of integral currents in converging weakly to an integral current . Moreover, assume that converges to . Then the mass measure converges weakly to as measures on .
Remark 3.9.
Sometimes, as in Theorem 0.1, a sequence of -manifolds converges to a nice limit space in a given canonical topology only after removing negligible subsets from . For an example different from Theorem 0.1 and related to scalar curvature, see [DS23]. To quantify that phenomenon, we can look at the coarse dimension of . To measure the coarse dimension of a manifold , we propose the following notion of “Euclidean -area” :
where denotes the standard -dimensional Hausdorff measure. Let us declare that has coarse dimension if . 222A reason why we do not use the notion of Uryson width instead of Euclidean -area is that can usually be chosen to have small -dimensional Uryson width. As a corollary of Theorem 0.1, for the volume entropy inequality, hyperbolic manifolds of dimension are “codimension 2 stable” in the measured Gromov-Hausdorff topology. This is in general optimal. What about other stability and convergence problems?
References
- [AK00a] Luigi Ambrosio and Bernd Kirchheim. Currents in metric spaces. Acta Math., 185(1):1–80, 2000.
- [AK00b] Luigi Ambrosio and Bernd Kirchheim. Rectifiable sets in metric and Banach spaces. Math. Ann., 318(3):527–555, 2000.
- [All21] Brian Allen. Almost non-negative scalar curvature on Riemannian manifolds conformal to tori. The Journal of Geometric Analysis, 31(11):11190–11213, 2021.
- [Aub05] Erwann Aubry. Pincement sur le spectre et le volume en courbure de Ricci positive. In Annales scientifiques de l’Ecole normale supérieure, volume 38, pages 387–405, 2005.
- [BBCG12] Laurent Bessières, Gérard Besson, Gilles Courtois, and Sylvestre Gallot. Differentiable rigidity under Ricci curvature lower bound. Duke Math. J., 161(1):29–67, 2012.
- [BBI22] Dmitri Burago, Yuri Burago, and Sergei Ivanov. A course in metric geometry, volume 33. American Mathematical Society, 2022.
- [BCG91] Gérard Besson, Gilles Courtois, and Sylvestre Gallot. Volume et entropie minimale des espaces localement symétriques. Invent. Math., 103(2):417–445, 1991.
- [BCG95] Gérard Besson, Gilles Courtois, and Sylvestre Gallot. Entropies et rigidités des espaces localement symétriques de courbure strictement négative. Geom. Funct. Anal., 5(5):731–799, 1995.
- [BCG96] Gérard Besson, Gilles Courtois, and Sylvestre Gallot. Minimal entropy and Mostow’s rigidity theorems. Ergodic Theory Dynam. Systems, 16(4):623–649, 1996.
- [BCG99] Gérard Besson, Gilles Courtois, and Sylvestre Gallot. Lemme de schwarz réel et applications géométriques. Acta Mathematica, 183(2):145–169, 1999.
- [BCS23] Giuliano Basso, Paul Creutz, and Elefterios Soultanis. Filling minimality and Lipschitz-volume rigidity of convex bodies among integral current spaces. arxiv preprint arXiv:2209.12545v3, 2023.
- [Bru08] Michael Brunnbauer. Homological invariance for asymptotic invariants and systolic inequalities. Geom. Funct. Anal., 18(4):1087–1117, 2008.
- [But22] Karen Butt. Quantitative marked length spectrum rigidity. arXiv preprint arXiv:2203.12128, 2022.
- [CC97] Jeff Cheeger and Tobias H. Colding. On the structure of spaces with Ricci curvature bounded below. I. J. Differential Geom., 46(3):406–480, 1997.
- [CDNZ+21] Chris Connell, Xianzhe Dai, Jesús Núñez-Zimbrón, Raquel Perales, Pablo Suárez-Serrato, and Guofang Wei. Maximal volume entropy rigidity for spaces. Journal of the London Mathematical Society, 104(4):1615–1681, 2021.
- [CF03] Christopher Connell and Benson Farb. The degree theorem in higher rank. Journal of Differential Geometry, 65(1):19–59, 2003.
- [CL22] Jianchun Chu and Man-Chun Lee. Conformal tori with almost non-negative scalar curvature. Calculus of Variations and Partial Differential Equations, 61(3):114, 2022.
- [Col96a] Tobias H. Colding. Large manifolds with positive Ricci curvature. Invent. Math., 124(1-3):193–214, 1996.
- [Col96b] Tobias H. Colding. Shape of manifolds with positive Ricci curvature. Invent. Math., 124(1-3):175–191, 1996.
- [Cou98] Gilles Courtois. Des questions de stabilité. Séminaire de théorie spectrale et géométrie, 17:159–162, 1998.
- [CRX19] Lina Chen, Xiaochun Rong, and Shicheng Xu. Quantitative volume space form rigidity under lower Ricci curvature bound I. Journal of Differential Geometry, 113(2):227–272, 2019.
- [DP23] Giacomo Del Nin and Raquel Perales. Rigidity of mass-preserving -Lipschitz maps from integral current spaces into . arxiv preprint arXiv:2210.06406v3, 2023.
- [DS23] Conghan Dong and Antoine Song. Stability of Euclidean 3-space for the Positive Mass Theorem. arxiv preprint, 2023.
- [EM93] Alex Eskin and Curt McMullen. Mixing, counting, and equidistribution in Lie groups. Duke mathematical journal, 71(1):181–209, 1993.
- [GKL22] Colin Guillarmou, Gerhard Knieper, and Thibault Lefeuvre. Geodesic stretch, pressure metric and marked length spectrum rigidity. Ergodic Theory and Dynamical Systems, 42(3):974–1022, 2022.
- [GL19] Colin Guillarmou and Thibault Lefeuvre. The marked length spectrum of Anosov manifolds. Annals of Mathematics, 190(1):321–344, 2019.
- [HLS17] Lan-Hsuan Huang, Dan A Lee, and Christina Sormani. Intrinsic flat stability of the positive mass theorem for graphical hypersurfaces of Euclidean space. Journal für die reine und angewandte Mathematik (Crelles Journal), 2017(727):269–299, 2017.
- [KNPS21] Mikhail Karpukhin, Mickaël Nahon, Iosif Polterovich, and Daniel Stern. Stability of isoperimetric inequalities for Laplace eigenvalues on surfaces. arXiv preprint arXiv:2106.15043, 2021.
- [KSX23] Demetre Kazaras, Antoine Song, and Kai Xu. Scalar curvature and volume entropy of hyperbolic -manifolds. arxiv preprint, 2023.
- [Lan11] Urs Lang. Local currents in metric spaces. J. Geom. Anal., 21(3):683–742, 2011.
- [LNN20] Man-Chun Lee, Aaron Naber, and Robin Neumayer. convergence and -regularity theorems for entropy and scalar curvature lower bounds. arXiv preprint arXiv:2010.15663, 2020.
- [LS14] Dan A Lee and Christina Sormani. Stability of the positive mass theorem for rotationally symmetric riemannian manifolds. Journal für die reine und angewandte Mathematik (Crelles Journal), 2014(686):187–220, 2014.
- [LW11] François Ledrappier and Xiaodong Wang. Pinching theorems for the volume entropy. preprint, 2011.
- [Pet99] Peter Petersen. On eigenvalue pinching in positive Ricci curvature. Inventiones mathematicae, 138(1):1–21, 1999.
- [Por15] Jacobus W Portegies. Semicontinuity of eigenvalues under intrinsic flat convergence. Calculus of Variations and Partial Differential Equations, 54(2):1725–1766, 2015.
- [Rua22] Yuping Ruan. The Cayley hyperbolic space and volume entropy rigidity. arXiv preprint arXiv:2203.14418, 2022.
- [S+21] Christina Sormani et al. Conjectures on convergence and scalar curvature. arXiv preprint arXiv:2103.10093, 2021.
- [Sam99] Andrea Sambusetti. Minimal entropy and simplicial volume. Manuscripta Math., 99(4):541–560, 1999.
- [Son23] Antoine Song. Spherical volume and spherical Plateau problem. arXiv preprint, 2023.
- [Sor18] Christina Sormani. Intrinsic flat Arzela–Ascoli theorems. Communications in Analysis and Geometry, 26(6):1317–1373, 2018.
- [Sou08] Juan Souto. Two applications of the natural map. Geometriae Dedicata, 133(1):51–57, 2008.
- [SW11] Christina Sormani and Stefan Wenger. The intrinsic flat distance between Riemannian manifolds and other integral current spaces. J. Differential Geom., 87(1):117–199, 2011.
- [V+09] Cédric Villani et al. Optimal transport: old and new, volume 338. Springer, 2009.
- [Wen07] Stefan Wenger. Flat convergence for integral currents in metric spaces. Calc. Var. Partial Differential Equations, 28(2):139–160, 2007.
- [Wen11] Stefan Wenger. Compactness for manifolds and integral currents with bounded diameter and volume. Calc. Var. Partial Differential Equations, 40(3-4):423–448, 2011.
- [Züs23] Roger Züst. Lipschitz rigidigy of Lipschitz manifolds among integral current spaces. arxiv preprint arXiv:2302.07587v1, 2023.