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

Entropy and stability of hyperbolic manifolds

Antoine Song California Institute of Technology
177 Linde Hall, #1200 E. California Blvd., Pasadena, CA 91125
[email protected]
Abstract.

Let (M,g0)(M,g_{0}) be a closed oriented hyperbolic manifold of dimension at least 33. By the volume entropy inequality of G. Besson, G. Courtois and S. Gallot, for any Riemannian metric gg on MM with same volume as g0g_{0}, its volume entropy h(g)h(g) satisfies h(g)n1h(g)\geq n-1 with equality only when gg is isometric to g0g_{0}. We show that the hyperbolic metric g0g_{0} is stable in the following sense: if gig_{i} is a sequence of Riemaniann metrics on MM of same volume as g0g_{0} and if h(gi)h(g_{i}) converges to n1n-1, then there are smooth subsets ZiMZ_{i}\subset M such that both Vol(Zi,gi)\operatorname{Vol}(Z_{i},g_{i}) and Area(Zi,gi)\operatorname{Area}(\partial Z_{i},g_{i}) tend to 0, and (MZi,gi)(M\setminus Z_{i},g_{i}) converges to (M,g0)(M,g_{0}) in the measured Gromov-Hausdorff topology. The proof relies on showing that any spherical Plateau solution for MM is intrinsically isomorphic to (M,(n1)24ng0)(M,\frac{(n-1)^{2}}{4n}g_{0}).

Introduction

Let MM be a hyperbolic manifold of dimension at least 33 with hyperbolic metric g0g_{0}. If gg is a Riemannian metric on MM, let h(g)h(g) denote its volume entropy:

h(g):=limRlogVol(B~g(o,R),g)Rh(g):=\lim_{R\to\infty}\frac{\log\operatorname{Vol}(\tilde{B}_{g}(o,R),g)}{R}

where B~g(o,R)\tilde{B}_{g}(o,R) denotes the geodesic RR-ball centered at some point oo in the universal cover (M~,g)(\tilde{M},g) of (M,g)(M,g). The fundamental volume entropy inequality, proved by Besson-Courtois-Gallot in [BCG95, BCG96], asserts that for any Riemannian metric gg on MM of same volume as g0g_{0}, we have

(1) h(g)h(g0)=n1.h(g)\geq h(g_{0})=n-1.

Moreover, Besson-Courtois-Gallot showed that this inequality is rigid in the sense that if equality holds, then gg is isometric to g0g_{0}. How stable is the volume entropy inequality? We find that stability holds after removing negligible subsets:

Theorem 0.1.

Let (M,g0)(M,g_{0}) be a closed oriented hyperbolic manifold of dimension at least 33. Let {gi}i1\{g_{i}\}_{i\geq 1} be a sequence of Riemannian metrics on MM with Vol(M,gi)=Vol(M,g0)\operatorname{Vol}(M,g_{i})=\operatorname{Vol}(M,g_{0}). If

limih(gi)=n1,\lim_{i\to\infty}h(g_{i})=n-1,

then there is a sequence of smooth subsets ZiMZ_{i}\subset M such that

limiVol(Zi,gi)=limiArea(Zi,gi)=0\lim_{i\to\infty}\operatorname{Vol}(Z_{i},g_{i})=\lim_{i\to\infty}\operatorname{Area}(\partial Z_{i},g_{i})=0

and (MZi,gi)(M\setminus Z_{i},g_{i}) converges to (M,g0)(M,g_{0}) in the measured Gromov-Hausdorff topology.

In the statement of Theorem 0.1, (MZi,gi)(M\setminus Z_{i},g_{i}) is the metric space where the distance between two points a,bMZia,b\in M\setminus Z_{i} is given by the infimum of the gig_{i}-lengths of curves joining aa to bb inside MZiM\setminus Z_{i}. 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 limiVol(Zi,gi)=0\lim_{i\to\infty}\operatorname{Vol}(Z_{i},g_{i})=0. On the other hand, the conclusion that limiArea(Zi,gi)=0\lim_{i\to\infty}\operatorname{Area}(\partial Z_{i},g_{i})=0 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 g0g_{0}, we get a new metric gg whose volume and volume entropy are arbitrarily close to Vol(M,g0)\operatorname{Vol}(M,g_{0}) and n1n-1 respectively. In this example, (M,g)(M,g) is far from (M,g0)(M,g_{0}) in the Gromov-Hausdorff topology, although it is still close to (M,g0)(M,g_{0}) in the Gromov-Prokhorov topology. The following question remains open: under the assumptions of Theorem 0.1, does (M,gi)(M,g_{i}) converge to (M,g0)(M,g_{0}) 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].

These two parts together yield a stability result stronger than Theorem 0.1: under the same assumptions, (MZi,gi)(M\setminus Z_{i},g_{i}) actually converges to (M,g0)(M,g_{0}) with respect to the intrinsic flat topology, see Theorem 3.7.

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 (M,g0)(M,g_{0}) be a closed oriented hyperbolic manifold with its hyperbolic metric and let Γ:=π1(M)\Gamma:=\pi_{1}(M). Consider the unit sphere SS^{\infty} in the Hilbert space 2(Γ)\ell^{2}(\Gamma) and let Γ\Gamma act on SS^{\infty} by the regular representation λΓ\lambda_{\Gamma}. Denote by S/λΓ(Γ)S^{\infty}/\lambda_{\Gamma}(\Gamma) the corresponding quotient manifold, endowed with the standard round Hilbert Riemannian metric 𝐠Hil\mathbf{g}_{\text{Hil}}. There is a unique homotopy class M\mathscr{H}_{M} of smooth immersions from MM to S/λΓ(Γ)S^{\infty}/\lambda_{\Gamma}(\Gamma) inducing an isomorphism on the fundamental groups. Besson-Courtois-Gallot define the spherical volume of MM [BCG91] as follows

SphereVol(M):=inf{Vol(M,ϕ𝐠Hil);ϕM}.\operatorname{SphereVol}(M):=\inf\{\operatorname{Vol}(M,\phi^{*}\mathbf{g}_{\text{Hil}});\quad\phi\in\mathscr{H}_{M}\}.

A key step in Besson-Courtois-Gallot’s proof [BCG95, BCG96] of the entropy inequality (1) is to establish that

SphereVol(M)=Vol(M,(n1)24ng0).\operatorname{SphereVol}(M)=\operatorname{Vol}(M,\frac{(n-1)^{2}}{4n}g_{0}).

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 ϕiM\phi_{i}\in\mathscr{H}_{M}, namely a sequence such that

limiVol(M,ϕi𝐠Hil)=SphereVol(M).\lim_{i\to\infty}\operatorname{Vol}(M,\phi_{i}^{*}\mathbf{g}_{\text{Hil}})=\operatorname{SphereVol}(M).

Then by Wenger’s compactness theorem [Wen11], the images ϕi(M)\phi_{i}(M) subsequentially converge as integral current spaces to an integral current space

C=(X,d,T)C_{\infty}=(X_{\infty},d_{\infty},T_{\infty})

in the intrinsic flat topology, in the sense of Sormani-Wenger [SW11]. Here (X,d)(X_{\infty},d_{\infty}) is a metric space, TT_{\infty} is an integral current in the completion of (X,d)(X_{\infty},d_{\infty}), see Subsection 1.1. We call any such limit CC_{\infty} a spherical Plateau solution for MM.

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 (M,g0)(M,g_{0}) is a closed oriented hyperbolic manifold of dimension n3n\geq 3, then any spherical Plateau solution for MM is intrinsically isomorphic to (M,(n1)24ng0)(M,\frac{(n-1)^{2}}{4n}g_{0}).

Theorem 0.2 leads to a rigidity result with a representation theoretic flavor for π1(M)\pi_{1}(M), 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 33-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.

Remark 0.3.

The arguments in this paper extend to closed oriented manifolds which are locally symmetric of rank one due to [BCG96, Rua22], and so versions of the main theorems hold more generally for these spaces.

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

SphereVol(M)=Vol(M,(n1)24ng0).\operatorname{SphereVol}(M)=\operatorname{Vol}(M,\frac{(n-1)^{2}}{4n}g_{0}).

Their main tool was the barycenter map Bar{\mathrm{Bar}}. In our setting, this is a Lipschitz map which under some technical conditions sends cycles in S/λΓ(Γ)S^{\infty}/\lambda_{\Gamma}(\Gamma) of the form ϕ(M)\phi(M), where ϕ(M)\phi\in\mathcal{F}(M), to the rescaled hyperbolic manifold (M,(n1)24ng0)(M,\frac{(n-1)^{2}}{4n}g_{0}) with topological degree 11. Roughly speaking, the Jacobian of restriction of the barycenter map Bar:ϕ(M)(M,(n1)24ng0){\mathrm{Bar}}:\phi(M)\to(M,\frac{(n-1)^{2}}{4n}g_{0}) satisfies [BCG95, BCG96]

(2) |JacBar|1,|\operatorname{Jac}{\mathrm{Bar}}|\leq 1,

which in particular implies that SphereVol(M)Vol(M,(n1)24ng0)\operatorname{SphereVol}(M)\geq\operatorname{Vol}(M,\frac{(n-1)^{2}}{4n}g_{0}). Then the opposite inequality is checked by finding an explicit sequence of embeddings ϕi(M)\phi_{i}\in\mathcal{F}(M) such that limiVol(M,ϕi𝐠Hil)=Vol(M,(n1)24ng0).\lim_{i\to\infty}\operatorname{Vol}(M,\phi_{i}^{*}\mathbf{g}_{\text{Hil}})=\operatorname{Vol}(M,\frac{(n-1)^{2}}{4n}g_{0}).

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 ϕi(M)\phi_{i}\in\mathcal{F}(M), and denote by CiC_{i} the integral currents of S/λΓ(Γ)S^{\infty}/\lambda_{\Gamma}(\Gamma) induced by pushing forward the fundamental class of MM by ϕi\phi_{i}. The barycenter map Bar{\mathrm{Bar}} enjoys the Jacobian bound (2) which is almost achieved on a region Ωispt(Ci)\Omega_{i}\subset\operatorname{spt}(C_{i}) that covers almost all of spt(Ci)\operatorname{spt}(C_{i}) as ii\to\infty. Nontrivially, this implies a local Lipschitz bound for Bar{\mathrm{Bar}}, which holds on a whole neighborhood of Ωi\Omega_{i}, and the differential of Bar{\mathrm{Bar}} at points of Ωi\Omega_{i} can be shown to be close to a linear isometry. We can assume, by Wenger’s compactness theorem, that CiC_{i} converges to a spherical Plateau solution

C=(X,d,S)C_{\infty}=(X_{\infty},d_{\infty},S_{\infty})

(the fact that such a limit exists is crucial). We then construct a limit map from the support of SS_{\infty} to MM:

Bar:sptSM{\mathrm{Bar}}_{\infty}:\operatorname{spt}S_{\infty}\to M

sending the current structure SS_{\infty} to the natural current structure 1M\llbracket 1_{M}\rrbracket supported on MM. Heuristically, as ii goes to infinity, the Jacobian bound (2) for Bar:spt(Ci)(M,(n1)24ng0){\mathrm{Bar}}:\operatorname{spt}(C_{i})\to(M,\frac{(n-1)^{2}}{4n}g_{0}) should be almost saturated almost everywhere, which means that the differential of Bar{\mathrm{Bar}} should be close to a linear isometry almost everywhere. In other words, Bar:spt(Ci)(M,(n1)24ng0){\mathrm{Bar}}:\operatorname{spt}(C_{i})\to(M,\frac{(n-1)^{2}}{4n}g_{0}) are almost Riemannian isometries. Passing to the limit, we should be able to deduce that Bar{\mathrm{Bar}}_{\infty} 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 11-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 11-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 11-Lipschitz in our situation (even though it will a posteriori follow that it is 11-Lipschitz for the intrinsic metric on sptS\operatorname{spt}S_{\infty}). 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 gg on MM with same volume as g0g_{0} and with entropy close to n1n-1. Then, there is a uniformly Lipschitz map

𝒫:(M,(n1)24ng)(S/λΓ(Γ),𝐠Hil)\mathcal{P}:(M,\frac{(n-1)^{2}}{4n}g)\to(S^{\infty}/\lambda_{\Gamma}(\Gamma),\mathbf{g}_{\mathrm{Hil}})

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 Bar𝒫\mathrm{Bar}\circ\mathcal{P} instead of Bar\mathrm{Bar}. We deduce that, for smooth subset ZMZ\subset M,

  • Vol(Z,g)\operatorname{Vol}(Z,g) and Area(Z,g)\operatorname{Area}(\partial Z,g) are both small,

  • (MZ,g)(M\setminus Z,g) is close in the intrinsic flat topology to a space C=(X,d,S)C_{\infty}=(X_{\infty},d_{\infty},S_{\infty}),

  • there is a bi-Lipschitz, 11-Lipschitz map

    Ψ:(M,g0)(sptS,d),\Psi:(M,g_{0})\to(\operatorname{spt}S_{\infty},d_{\infty}),
  • (MZ,g)(M\setminus Z,g) is Gromov-Hausdorff close to (sptS,d)(\operatorname{spt}S_{\infty},d_{\infty}) via a topologically natural map.

The properties of BarΨ\mathrm{Bar}\circ\Psi are not as good as those of Bar\mathrm{Bar}, so unlike Theorem 0.2, we cannot readily conclude that Ψ\Psi is an isometry for the intrinsic metrics. We need to remove a small subset ZZ from MM to get the Gromov-Hausdorff closeness property above.

In order to prove that the map Ψ\Psi above is, in fact, an isometry, we rely on a volume entropy comparison result, Theorem 3.5. The latter roughly says that if (MZ,g)(M\setminus Z,g) is naturally Gromov-Hausdorff close to a metric space (M,d)(M,d) and if there is a 11-Lipschitz map Ψ\Psi from (M,g0)(M,g_{0}) to (M,d)(M,d), then either Ψ\Psi is an isometry or the volume entropy of (M,g)(M,g) is strictly larger than n1n-1. 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 (sptS,d)(\operatorname{spt}S_{\infty},d_{\infty}), we conclude that the map Ψ\Psi 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 nn-dimensional integral current SS has a well-defined notion of boundary S\partial S which is an (n1)(n-1)-dimensional integral current, a notion of volume measure denoted by S\|S\| and a notion of total volume called mass 𝐌(S)\mathbf{M}(S). Such a current SS is concentrated on a so-called canonical set set(S)\operatorname{set}(S), itself included in the support spt(S)\operatorname{spt}(S) of the measure S\|S\|. The restriction of SS to a Borel set AA is denoted by SAS\llcorner A, and its push-forward by a Lipschitz map ϕ\phi is called ϕS\phi_{\sharp}S. With those notations, 𝐌(SA)=S(A).\mathbf{M}(S\llcorner A)=\|S\|(A). 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 (X,d,S)(X,d,S) where (X,d)(X,d) is a metric space and SS is an integral current in the completion of (X,d)(X,d), which we will usually denote by sptS\operatorname{spt}S (one requires that XX is the “canonical set” of the current SS) [Son23, Definition 1.3 and Subsection 1.1]. A simple example of integral current space is given by a closed, connected, oriented Riemannian nn-manifold (N,h)(N,h): the metric space is NN endowed with the geodesic distance of hh, and the integral current structure 1N\llbracket 1_{N}\rrbracket is the natural integral current induced by the fundamental class [N]Hn(N;)[N]\in H_{n}(N;\mathbb{Z}).

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:

Theorem 1.1.

[Wen11][SW11, Theorem 4.19] Given a sequence of boundaryless integral current spaces

(Xm,dm,Sm)(X_{m},d_{m},S_{m})

with uniformly bounded mass and diameter, there is a subsequence converging to an integral current space in the intrinsic flat topology.

1.2. Limits of almost Riemannian isometries and intrinsic flat limit spaces

As usual, inside an nn-dimensional Riemannian manifold, we will denote by Vol\operatorname{Vol} and Area\operatorname{Area} the nn-dimensional and (n1)(n-1)-dimensional Hausdorff measure. Sometimes, we also use k\mathcal{H}^{k} to denote the kk-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 hh is a Riemannian metric on a manifold NN, let disth\operatorname{dist}_{h} be the metric on MM induced by gg. Sometimes we will make the identification

(N,h)=(N,disth).(N,h)=(N,\operatorname{dist}_{h}).

We will use a few times the following simple fact: if (M,g)(M,g) is a compact Riemannian nn-manifold with a piecewise smooth metric gg inducing distg\operatorname{dist}_{g}, then for any metric dd whose induced intrinsic metric is distg\operatorname{dist}_{g}, and any open subset ΩM\Omega\subset M, the mass of 1Ω\llbracket 1_{\Omega}\rrbracket as an nn-dimensional current in (M,d)(M,d) is at most Vol(Ω,g)\operatorname{Vol}(\Omega,g).

Lemma 1.2.

[Sor18, BCS23] Let (E1,d1),(E2,d2)(E_{1},d_{1}),(E_{2},d_{2}) be two complete metric spaces. Let SiS_{i} be a sequence of integral currents in (E1,d1)(E_{1},d_{1}) and let

φi:sptSi(E2,d2)\varphi_{i}:\operatorname{spt}S_{i}\to(E_{2},d_{2})

be a sequence of λ\lambda-Lipschitz maps for some λ>0\lambda>0 independent of ii. Suppose that SiS_{i} (resp. (φi)Si(\varphi_{i})_{\sharp}S_{i}) converges in the flat topology to an integral current SS_{\infty} (resp. TT_{\infty}) inside (E1,d1)(E_{1},d_{1}) (resp. inside (E2,d2)(E_{2},d_{2})), and that (E2,d2)(E_{2},d_{2}) is compact.

Then there is a λ\lambda-Lipschitz map

φ:(sptS,d)(E2,d2)\varphi_{\infty}:(\operatorname{spt}S_{\infty},d_{\infty})\to(E_{2},d_{2})

such that:

  1. (1)

    after taking a subsequence if necessary, for any positive integer mm and any collection of mm points {x,1,,x,m}sptS\{x_{\infty,1},...,x_{\infty,m}\}\subset\operatorname{spt}S_{\infty}, there is a sequence of collections of mm points {xi,1,,xi,m}Ni\{x_{i,1},...,x_{i,m}\}\subset N_{i} such that for each j{1,,m}j\in\{1,...,m\}, as ii\to\infty, xi,jx_{i,j} converges to x,jx_{\infty,j}, and φi(xi,j)\varphi_{i}(x_{i,j}) converges to φ(x,j)\varphi_{\infty}(x_{\infty,j}),

  2. (2)

    (φ)S=T(\varphi_{\infty})_{\sharp}S_{\infty}=T_{\infty} as currents inside (E2,d2)(E_{2},d_{2}).

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 L(E2)L^{\infty}(E_{2}) is the Banach space of bounded real functions on E2E_{2} endowed with the LL^{\infty} norm, then it is well-known that (E2,d2)(E_{2},d_{2}) embeds isometrically inside L(E2)L^{\infty}(E_{2}) by the Kuratowski embedding, and L(E2)L^{\infty}(E_{2}) is an injective metric space in the following sense: given any other metric space YY, a subset AYA\subset Y, and a λ\lambda-Lipschitz map ϕ:AL(E2)\phi:A\to L^{\infty}(E_{2}), there exists an extension of ϕ\phi, called ϕ~:YL(E2)\tilde{\phi}:Y\to L^{\infty}(E_{2}), which is still λ\lambda-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 (N,h)(N,h) be a connected, closed, oriented Riemannian nn-manifold. Let SiS_{i} be a sequence of integral currents in a complete metric space (E,d)(E,d), converging in the flat topology to an integral current SS_{\infty} inside (E,d)(E,d). Suppose that

  1. (a)

    each support Ni:=sptSiN_{i}:=\operatorname{spt}S_{i}, endowed with the intrinsic metric induced by the metric dd, is a compact, oriented Riemannian manifold (Ni,hi)(N_{i},h_{i}) with a piecewise smooth metric hih_{i} (possibly with nonempty piecewise smooth boundary),

  2. (b)

    limiArea(Ni,hi)=0\lim_{i\to\infty}\operatorname{Area}(\partial N_{i},h_{i})=0,

  3. (c)

    there is a sequence of maps

    φi:(Ni,d|Ni)(N,disth)\varphi_{i}:(N_{i},d|_{N_{i}})\to(N,\operatorname{dist}_{h})

    which are C1C^{1} on the smooth part of NiN_{i} and λ\lambda-Lipschitz for some λ>0\lambda>0 independent of ii, such that (φi)(Si)(\varphi_{i})_{\sharp}(S_{i}) converges to 1N\llbracket 1_{N}\rrbracket in the flat topology inside (N,h)(N,h),

  4. (d)

    there is a sequence of open subsets RiR_{i} contained in the part of NiN_{i} where hih_{i} is smooth, such that limiVol(NiRi,hi)=0\lim_{i\to\infty}\operatorname{Vol}(N_{i}\setminus R_{i},h_{i})=0 and limiVol(Ri,hi)=Vol(N,h)\lim_{i\to\infty}\operatorname{Vol}(R_{i},h_{i})=\operatorname{Vol}(N,h),

  5. (e)

    moreover, φi\varphi_{i} is almost a Riemannian isometry on RiR_{i} in the sense that

    limiu,v=1n|h(dφi(eu),dφi(ev))δuv|L(Ri)=0,\lim_{i\to\infty}\big{\|}\sum_{u,v=1}^{n}|h(d\varphi_{i}(e^{\prime}_{u}),d\varphi_{i}(e^{\prime}_{v}))-\delta_{uv}|\big{\|}_{L^{\infty}(R_{i})}=0,

    where {eu}u=1n\{e^{\prime}_{u}\}_{u=1}^{n} denotes any choice of orthonormal bases for the tangent spaces of (N,hi)(N,h_{i}).

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

φ:sptS(N,disth).\varphi_{\infty}:\operatorname{spt}S_{\infty}\to(N,\operatorname{dist}_{h}).

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

φ:sptS(N,disth)\varphi_{\infty}:\operatorname{spt}S_{\infty}\to(N,\operatorname{dist}_{h})

be the limit map constructed in Lemma 1.2.

  1. (1)

    Then φ\varphi_{\infty} is a bi-Lipschitz bijection and its inverse φ1:(N,disth)sptS\varphi_{\infty}^{-1}:(N,\operatorname{dist}_{h})\to\operatorname{spt}S_{\infty} is 11-Lipschitz with respect to the induced intrinsic metrics.

  2. (2)

    Suppose additionally that for any ϵ>0\epsilon>0, there is rϵ>0r_{\epsilon}>0 such that if ii is large enough, then for any x,yNix,y\in N_{i} such that d(x,y)<rϵd(x,y)<r_{\epsilon}, we have

    disth(φi(x),φi(y))(1+ϵ)d(x,y).\operatorname{dist}_{h}(\varphi_{i}(x),\varphi_{i}(y))\leq(1+\epsilon)d(x,y).

    Then φ\varphi_{\infty} is an isometry with respect to the induced intrinsic metrics.

Remark 1.5.

The limit map φ\varphi_{\infty} in (1) is not 11-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 gEuclg_{\mathrm{Eucl}} on the Euclidean unit sphere S2S^{2}, and for each i>0i>0, consider the conformal metric gi:=f2.gEuclg_{i}:=f^{2}.g_{\mathrm{Eucl}} where f:S2[12,1]f:S^{2}\to[\frac{1}{2},1] is 11 outside the 1i\frac{1}{i}-neighborhood of the equator and 12\frac{1}{2} in the 12i\frac{1}{2i}-neighborhood of the equator. Let distgEucl\operatorname{dist}_{g_{\mathrm{Eucl}}} and distgi\operatorname{dist}_{g_{i}} be the corresponding intrinsic metrics. Then Assumption 1.3 is satisfied with φi\varphi_{i} being the identity map id:(S2,distgi)(S2,distgEucl)\operatorname{id}:(S^{2},\operatorname{dist}_{g_{i}})\to(S^{2},\operatorname{dist}_{g_{\mathrm{Eucl}}}), etc. However, the intrinsic flat limit and Gromov-Hausdorff limit of (S2,gi)(S^{2},g_{i}) are both determined by the length structure LL on S2S^{2} induced by distgEucl\operatorname{dist}_{g_{\mathrm{Eucl}}} for curves not touching the equator, and with an equator of length π\pi instead of 2π2\pi. The limit φ\varphi_{\infty} is still the identity map id:(S2,L)(S2,distgEucl)\operatorname{id}:(S^{2},L)\to(S^{2},\operatorname{dist}_{g_{\mathrm{Eucl}}}) and it is not 11-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 ϵ\epsilon, together with Lemma 1.2 (1), we obtain that φ\varphi_{\infty} does not increase distances for the intrinsic metrics, in other words it is 11-Lipschitz for the intrinsic metrics. Since property (1) says that the inverse of φ\varphi_{\infty} is also 11-Lipschitz for the intrinsic metrics, it is an isometry.

It remains to prove property (1). Note that by Lemma 1.2 (1), φ\varphi_{\infty} is λ\lambda-Lipschitz. Let LdL_{d} be the intrinsic metric on sptS\operatorname{spt}S_{\infty} induced by the restricted metric d|Sd|_{S_{\infty}} (a priori LdL_{d} is allowed to take \infty 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 φi\varphi_{i} is injective on RiR_{i} without loss of generality by reducing that domain a bit.

For η>0\eta>0, set

Oη:=η-neighborhood of sptS inside (E,d).O_{\eta}:=\text{$\eta$-neighborhood of $\operatorname{spt}S_{\infty}$ inside $(E,d)$}.

Then for every η>0\eta>0,

(3) limiSi(EOη)=0,\lim_{i\to\infty}\|S_{i}\|(E\setminus O_{\eta})=0,

Indeed, let us assume on the contrary that for some η>0\eta>0, lim infiSi(EOη)>0\liminf_{i\to\infty}\|S_{i}\|(E\setminus O_{\eta})>0. Then, by Assumption 1.3 (a) (d) (e), we should have

lim infi𝐌((φi)(SiOη))\displaystyle\liminf_{i\to\infty}\mathbf{M}((\varphi_{i})_{\sharp}(S_{i}\llcorner O_{\eta})) =lim infi𝐌(SiOη)\displaystyle=\liminf_{i\to\infty}\mathbf{M}(S_{i}\llcorner O_{\eta})
<limi𝐌(Si)\displaystyle<\lim_{i\to\infty}\mathbf{M}(S_{i})
=limi𝐌((φi)Si)=Vol(N,h).\displaystyle=\lim_{i\to\infty}\mathbf{M}((\varphi_{i})_{\sharp}S_{i})=\operatorname{Vol}(N,h).

By a standard application of the slicing theorem, we can assume without loss of generality that the restricted current Si(EOη)S_{i}\llcorner(E\setminus O_{\eta}) is an integral current converging to 0 in the flat topology as ii\to\infty. Thus (φi)(SiOη)(\varphi_{i})_{\sharp}(S_{i}\llcorner O_{\eta}) still converges to 1N\llbracket 1_{N}\rrbracket in the flat topology. By Assumption 1.3 (c) and lower semicontinuity of the mass with respect to flat or weak convergence,

lim infi𝐌((φi)(SiOη))𝐌(1N)=Vol(N,h).\liminf_{i\to\infty}\mathbf{M}((\varphi_{i})_{\sharp}(S_{i}\llcorner O_{\eta}))\geq\mathbf{M}(\llbracket 1_{N}\rrbracket)=\operatorname{Vol}(N,h).

This contradicts the previous inequality and so (3) was true.

Given a Lipschitz curve ω\omega in (E,d)(E,d), let length(E,d)(ω)\operatorname{length}_{(E,d)}(\omega) denote its length with respect to the metric dd. Next, it is convenient to show the following “curve lifting” property.

Curve lifting property: Let η>0\eta>0. Let x,ysptSx,y\in\operatorname{spt}S_{\infty} and let

l:=disth(φ(x),φ(y)).l:=\operatorname{dist}_{h}(\varphi_{\infty}(x),\varphi_{\infty}(y)).

Then there exists a compact connected Lipschitz curve ω\omega contained in OηO_{\eta}, starting at xx, ending at yy, and moreover

length(E,d)(ω)l+η.\operatorname{length}_{(E,d)}(\omega)\leq l+\eta.
Proof of the curve lifting property.

Let bxb_{x}, byb_{y} be the metric balls in (E,d)(E,d), of radius r^(0,η100)\hat{r}\in(0,\frac{\eta}{100}) centered at x,ysptSx,y\in\operatorname{spt}S_{\infty}. By Lemma 1.2 (1), if r^\hat{r} is chosen small enough then for all ii large and every qbxsptSiq\in b_{x}\cap\operatorname{spt}S_{i} (resp. qbysptSiq\in b_{y}\cap\operatorname{spt}S_{i}), we have

disth(φi(q),φ(x))η/10(resp. disth(φi(q),φ(y))η/10).\operatorname{dist}_{h}(\varphi_{i}(q),\varphi_{\infty}(x))\leq\eta/10\quad\text{(resp. $\operatorname{dist}_{h}(\varphi_{i}(q),\varphi_{\infty}(y))\leq\eta/10$).}

By lower semicontinuity of the mass, for each ii large,

Si(bx)>2κandSi(by)>2κ\|S_{i}\|(b_{x})>2\kappa\quad\text{and}\quad\|S_{i}\|(b_{y})>2\kappa

for some κ>0\kappa>0 depending on r^,x,y\hat{r},x,y but independent of ii. For ii large, since we are assuming that φi\varphi_{i} is injective on RiR_{i} without loss of generality, by Assumption 1.3 (d) (e) and the area formula, we have the following volume estimates:

(4) n(φi(Ribx))κ,n(φi(Riby))κ.\begin{split}\mathcal{H}^{n}\big{(}\varphi_{i}(R_{i}\cap b_{x})\big{)}&\geq\kappa,\\ \quad\mathcal{H}^{n}\big{(}\varphi_{i}(R_{i}\cap b_{y})\big{)}&\geq\kappa.\end{split}

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 ii large enough, two points

y1,iRibx,y2,iRibyy_{1,i}\in R_{i}\cap b_{x},\quad y_{2,i}\in R_{i}\cap b_{y}

and a smooth curve

σiN\sigma_{i}\subset N

with lengthh(σi)l+η3\operatorname{length}_{h}(\sigma_{i})\leq l+\frac{\eta}{3}, joining φi(y1,i)\varphi_{i}(y_{1,i}) to φi(y2,i)\varphi_{i}(y_{2,i}) such that the restricted preimage

ϰi:=(φi)1(σi)\varkappa_{i}:=(\varphi_{i})^{-1}(\sigma_{i})

is a compact curve in NiN_{i} avoiding Ni\partial N_{i}, whose endpoints satisfies

φi(ϰi)φi(y1,i)φi(y2,i).\varphi_{i}(\partial\varkappa_{i})\subset\varphi_{i}(y_{1,i})\cup\varphi_{i}(y_{2,i}).

Because y1,iy_{1,i} and y2,iy_{2,i} belong to RiR_{i} (on which φi\varphi_{i} is assumed to be injective without loss of generality), in fact

(5) ϰi={y1,i,y2,i}.\partial\varkappa_{i}=\{y_{1,i},y_{2,i}\}.

By Assumption 1.3 (c) (d) (e), the restriction of SiS_{i} to the complement of RiR_{i} has mass converging to 0 as ii\to\infty; similarly, by (3), the restriction of SiS_{i} to the complement of Oη/2O_{\eta/2} has mass converging to 0. Thus the coarea formula again implies that we could choose ϰi\varkappa_{i} satisfying additionally:

(6) limi1(ϰiRi)=0,limi1(ji(ϰi)Oη/2)=0.\begin{split}\lim_{i\to\infty}\mathcal{H}^{1}(\varkappa_{i}\setminus R_{i})&=0,\\ \lim_{i\to\infty}\mathcal{H}^{1}(j_{i}(\varkappa_{i})\setminus O_{\eta/2})&=0.\end{split}

Together with Assumption 1.3 (e) and the area formula, these properties imply:

(7) 1(ϰi)(1+ϵi)(l+η/2)+ϵi\mathcal{H}^{1}(\varkappa_{i})\leq(1+\epsilon_{i})(l+\eta/2)+\epsilon_{i}

where limiϵi=0\lim_{i\to\infty}\epsilon_{i}=0. By using (6) and the fact that y1,iy_{1,i} (resp. y2,iy_{2,i}) is in bxb_{x} (resp. byb_{y}), we easily construct a new curve ωi\omega_{i} fully contained in OηO_{\eta} joining xx to yy, with length at most l+ηl+\eta for ii large. This proves the curve lifting property. ∎

The curve lifting property implies the following useful properties. Firstly, sptS\operatorname{spt}S_{\infty} is compact. Suppose towards a contraction that sptS\operatorname{spt}S_{\infty} is not compact, then for some r(0,1)r^{\prime}\in(0,1), there is an infinite sequence of points {xm}m0sptS\{x_{m}\}_{m\geq 0}\subset\operatorname{spt}S_{\infty} such that those points are pairwise at distance at least rr^{\prime} in (E,d)(E,d). By compactness of NN, for any ϵ>0\epsilon>0 there are m1m2m_{1}\neq m_{2} such that

disth(φ(xm1),φ(xm2))ϵ.\operatorname{dist}_{h}(\varphi_{\infty}(x_{m_{1}}),\varphi_{\infty}(x_{m_{2}}))\leq\epsilon.

Then the curve lifting property implies that the distance between xm1x_{m_{1}} and xm2x_{m_{2}} is at most ϵ\epsilon, a contradiction when ϵ<r/2\epsilon<r^{\prime}/2.

Secondly φ:sptSN\varphi_{\infty}:\operatorname{spt}S_{\infty}\to N is bijective. Indeed we verify that φ\varphi_{\infty} is injective by a direct application of the curve lifting property. Surjectivity follows from Lemma 1.2 (2) and the compactness of sptS\operatorname{spt}S_{\infty}.

We are ready to prove property (1) of our proposition. Take two points u,vNu,v\in N and let η>0\eta>0. Let x:=φ1(u)x:=\varphi_{\infty}^{-1}(u), y:=φ1(v)y:=\varphi_{\infty}^{-1}(v). By applying the curve lifting property repeatedly and making η0\eta\to 0, by compactness of sptS\operatorname{spt}S_{\infty} we get a limit Lipschitz curve in sptS\operatorname{spt}S_{\infty} joining xx to yy, with length at most disth(u,v)\operatorname{dist}_{h}(u,v). Thus the inverse φ\varphi_{\infty} is indeed 11-Lipschitz for the intrinsic metrics, and φ\varphi_{\infty} 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 λ>0\lambda>0. Consider a sequence of compact Riemannian nn-manifolds {Σi}i0\{\Sigma_{i}\}_{i\geq 0} with piecewise smooth metrics and with piecewise smooth boundaries. Suppose that there is a sequence of λ\lambda-Lipschitz smooth maps

{πi:Σin}i0,\{\pi^{\prime}_{i}:\Sigma_{i}\to\mathbb{R}^{n}\}_{i\geq 0},

and that for each ii, consider there are subsets Ai,BiΣiA_{i},B_{i}\subset\Sigma_{i} so that

limin1(πi(Σi)[0,3]×[0,1]n1)=0,\lim_{i\to\infty}\mathcal{H}^{n-1}(\pi^{\prime}_{i}(\partial\Sigma_{i})\cap[0,3]\times[0,1]^{n-1})=0,
πi(Ai)[0,1]×[0,1]n1,\pi^{\prime}_{i}(A_{i})\subset[0,1]\times[0,1]^{n-1},
πi(Bi)[2,3]×[0,1]n1,\pi^{\prime}_{i}(B_{i})\subset[2,3]\times[0,1]^{n-1},

and for some κ>0\kappa>0 independent of ii, for all ii:

n(πi(Ai))κ,\quad\mathcal{H}^{n}(\pi^{\prime}_{i}(A_{i}))\geq\kappa,
n(πi(Bi))κ.\mathcal{H}^{n}(\pi^{\prime}_{i}(B_{i}))\geq\kappa.

Let proj:nn1\mathrm{proj}:\mathbb{R}^{n}\to\mathbb{R}^{n-1} be the projection on the last n1n-1 coordinates. By Fubini’s theorem, for each ii we can find a vector ti{0}×[1,1]n1\overrightarrow{t_{i}}\in\{0\}\times[-1,1]^{n-1} such that if we set

𝒲i:={x[0,1]n1;\displaystyle\mathcal{W}_{i}:=\{x\in[0,1]^{n-1};\quad 1(proj1(x)πi(Ai))>0 and\displaystyle\mathcal{H}^{1}(\mathrm{proj}^{-1}(x)\cap\pi^{\prime}_{i}(A_{i}))>0\text{ and}
1(proj1(x)(πi(Bi)+ti))>0}\displaystyle\mathcal{H}^{1}(\mathrm{proj}^{-1}(x)\cap(\pi^{\prime}_{i}(B_{i})+\overrightarrow{t_{i}}))>0\}

then we have

n1(𝒲i)>κ\mathcal{H}^{n-1}(\mathcal{W}_{i})>\kappa^{\prime}

for some κ>0\kappa^{\prime}>0 independent of ii. By composing each πi\pi^{\prime}_{i} with a diffeomorphism with uniformly bounded Lipschitz constant if necessary, we can assume that ti=0\overrightarrow{t_{i}}=0. Applying the coarea formula and Sard’s theorem twice, first to the λ\lambda-Lipschitz maps πi:=projπi\pi_{i}:=\mathrm{proj}\circ\pi^{\prime}_{i}, then to the map (x1,xn)x1(x_{1},...x_{n})\mapsto x_{1}, we find for each ii some straight segment σi\sigma_{i} joining aiπi(Ai)a_{i}\in\pi^{\prime}_{i}(A_{i}) to biπi(Bi)b_{i}\in\pi^{\prime}_{i}(B_{i}), such that proj(σi)\mathrm{proj}(\sigma_{i}) is a point in [0,1]n1[0,1]^{n-1} and such that

ϰi:=(πi)1(σi)Σi\varkappa_{i}:=(\pi^{\prime}_{i})^{-1}(\sigma_{i})\subset\Sigma_{i}

is a compact smooth curve (with possibly several connected components) which avoids Σi\partial\Sigma_{i} for all ii large enough:

ϰiΣi=.\varkappa_{i}\cap\partial\Sigma_{i}=\varnothing.

This example generalizes when replacing n\mathbb{R}^{n} 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 MM be a closed oriented hyperbolic manifold, whose fundamental group is denoted by Γ\Gamma. Let SS^{\infty} be the unit sphere in 2(Γ)\ell^{2}(\Gamma). The 2\ell^{2}-norm induces a Hilbert Riemannian metric 𝐠Hil\mathbf{g}_{\mathrm{Hil}} on SS^{\infty}. The group Γ\Gamma acts isometrically on SS^{\infty} by the (left) regular representation λΓ:ΓEnd(2(Γ))\lambda_{\Gamma}:\Gamma\to\operatorname{End}(\ell^{2}(\Gamma)): for all γΓ\gamma\in\Gamma, xΓx\in\Gamma, fSf\in S^{\infty},

(λΓ(γ).f)(x):=f(γ1x).(\lambda_{\Gamma}(\gamma).f)(x):=f(\gamma^{-1}x).

Since Γ\Gamma is infinite and torsion-free, Γ\Gamma acts properly and freely on the infinite dimensional sphere SS^{\infty}. The quotient space S/λΓ(Γ)S^{\infty}/\lambda_{\Gamma}(\Gamma) is topologically a classifying space for Γ\Gamma. It is also a Hilbert manifold endowed with the induced Hilbert Riemannian metric 𝐠Hil\mathbf{g}_{\mathrm{Hil}}. The diameter of (S/λΓ(Γ),𝐠Hil)(S^{\infty}/\lambda_{\Gamma}(\Gamma),\mathbf{g}_{\mathrm{Hil}}) is bounded from above by π\pi.

Given base points p0Mp_{0}\in M, q0S/λΓ(Γ)q_{0}\in S^{\infty}/\lambda_{\Gamma}(\Gamma), there is a smooth immersion MS/λΓ(Γ)M\to S^{\infty}/\lambda_{\Gamma}(\Gamma) inducing the identity map from π1(M,p0)\pi_{1}(M,p_{0}) to π1(S/λΓ(Γ),q0)\pi_{1}(S^{\infty}/\lambda_{\Gamma}(\Gamma),q_{0}), which is unique up to homotopies sending p0p_{0} to q0q_{0}. Other choices of p0,q0p_{0},q_{0} yield homotopic maps, so that determines a unique homotopy class of maps which we call “admissible”. Set

M:={ϕ:MS/λΓ(Γ); ϕ is an admissible smooth immersion}.\mathscr{H}_{M}:=\{\phi:M\to S^{\infty}/\lambda_{\Gamma}(\Gamma);\quad\text{ $\phi$ is an admissible smooth immersion}\}.

Any map ϕM\phi\in\mathscr{H}_{M} defines the pull-back Riemannian metric ϕ(𝐠Hil)\phi^{*}(\mathbf{g}_{\mathrm{Hil}}) on MM.

Besson-Courtois-Gallot introduced the spherical volume of MM in [BCG91]. It can be equivalently be defined as follows.

Definition 2.1.

The spherical volume of MM is defined as

SphereVol(M):=inf{Vol(M,ϕ(𝐠Hil));ϕM}.\operatorname{SphereVol}(M):=\inf\{\operatorname{Vol}(M,\phi^{*}(\mathbf{g}_{\mathrm{Hil}}));\quad\phi\in\mathscr{H}_{M}\}.

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.

Theorem 2.2.

[BCG95, BCG96] Let (M,g0)(M,g_{0}) be a closed oriented hyperbolic manifold. Then

(8) SphereVol(M)=Vol(M,(n1)24ng0).\operatorname{SphereVol}(M)=\operatorname{Vol}(M,\frac{(n-1)^{2}}{{4n}}g_{0}).

A sequence ϕiM\phi_{i}\in\mathscr{H}_{M} is said to be minimizing if

limiVol(M,ϕi(𝐠Hil))=SphereVol(M).\lim_{i\to\infty}\operatorname{Vol}(M,\phi_{i}^{*}(\mathbf{g}_{\mathrm{Hil}}))=\operatorname{SphereVol}(M).

Denote by 1M\llbracket 1_{M}\rrbracket the integral current in (M,g0)(M,g_{0}) induced by MM and its orientation. For a Lipschitz map ϕ:MS/λΓ(Γ)\phi:M\to S^{\infty}/\lambda_{\Gamma}(\Gamma), recall that ϕ(1M)\phi_{\sharp}(\llbracket 1_{M}\rrbracket) denotes the push-forward integral current in S/λΓ(Γ)S^{\infty}/\lambda_{\Gamma}(\Gamma). We can now define spherical Plateau solutions.

Definition 2.3.

We call spherical Plateau solution for MM any nn-dimensional integral current space CC_{\infty} which is the limit in the intrinsic flat topology of a sequence Ci:=(ϕi)1MC_{i}:=(\phi_{i})_{\sharp}\llbracket 1_{M}\rrbracket where ϕiM\phi_{i}\in\mathscr{H}_{M} is a minimizing sequence.

For any sequence ϕiM\phi_{i}\in\mathscr{H}_{M} such that

limiVol(M,ϕi(𝐠Hil))=SphereVol(M),\lim_{i\to\infty}\operatorname{Vol}(M,\phi_{i}^{*}(\mathbf{g}_{\mathrm{Hil}}))=\operatorname{SphereVol}(M),

the mass and diameter of (ϕi)1M(\phi_{i})_{\sharp}\llbracket 1_{M}\rrbracket are uniformly bounded, so by Wenger’s compactness (Theorem 1.1) there is a subsequence of (ϕi)1M(\phi_{i})_{\sharp}\llbracket 1_{M}\rrbracket 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 M\mathscr{H}_{M} of admissible smooth immersions from MM to S/λΓ(Γ)S^{\infty}/\lambda_{\Gamma}(\Gamma), we believe that it is more natural to formulate the general spherical Plateau problem in terms of integral currents with compact support in S/λΓ(Γ)S^{\infty}/\lambda_{\Gamma}(\Gamma) representing a homology class hHn(Γ;)h\in H_{n}(\Gamma;\mathbb{Z}). 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 Γ\Gamma 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 2\ell^{2}-space on a group, instead of the L2L^{2}-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 (M,g0)(M,g_{0}) be a closed oriented hyperbolic manifold. Let (M~,g0)(\tilde{M},g_{0}) be its universal cover, namely the hyperbolic nn-space. Let Γ:=π1(M)\Gamma:=\pi_{1}(M). The latter acts properly cocompactly and freely on (M~,g0)(\tilde{M},g_{0}). Let SS^{\infty} be the unit sphere in the Hilbert space 2(Γ)\ell^{2}(\Gamma), on which Γ\Gamma acts freely and properly by isometries via the regular representation, so that S/λΓ(Γ)S^{\infty}/\lambda_{\Gamma}(\Gamma) is a smooth Hilbert manifold endowed with the standard round metric (see Subsection 2.1).

Set

ϰ(t):=1clog(cosh(ct))\varkappa(t):=\frac{1}{c}\log(\cosh(ct))

where cc is a positive constant. When we fix cc large enough, the following holds: for any wM~w\in\tilde{M}, the composition

ρw(.):=ϰ(distg0(w,.))\rho_{w}(.):=\varkappa(\operatorname{dist}_{g_{0}}(w,.))

is smooth everywhere and satisfies

(9) DdρwIddρwdρw.Dd\rho_{w}\geq\operatorname{Id}-d\rho_{w}\otimes d\rho_{w}.
Definition 2.5.

Fix a basepoint oM~o\in\tilde{M}. Let 𝕊+\mathbb{S}^{+} be the set of functions in SS^{\infty} with finite support. For f𝕊+f\in\mathbb{S}^{+}, consider the functional

(10) f:M~[0,]f(x):=γΓ|f(γ)|2ργ.o(x).\begin{split}\mathcal{B}_{f}&:\tilde{M}\to[0,\infty]\\ \mathcal{B}_{f}(x)&:=\sum_{\gamma\in\Gamma}|f(\gamma)|^{2}\rho_{\gamma.o}(x).\end{split}

The barycenter map is then defined as

Bar:𝕊+M~\mathrm{Bar}:\mathbb{S}^{+}\to\tilde{M}
Bar(f):= the unique point minimizing f.\mathrm{Bar}(f):=\text{ the unique point minimizing $\mathcal{B}_{f}$}.

The barycenter map is well-defined: the modified distance functions ργ.o\rho_{\gamma.o} are strictly convex, moreover f\mathcal{B}_{f} tends to infinity uniformly as xx\to\infty, so that the point where f\mathcal{B}_{f} attains its minimum exists and is unique. The subset 𝕊+S\mathbb{S}^{+}\subset S^{\infty} is invariant by Γ\Gamma, and Bar\mathrm{Bar} is Γ\Gamma-equivariant. The quotient map 𝕊+/ΓM\mathbb{S}^{+}/\Gamma\to M is also denoted by Bar\mathrm{Bar}. For more details, see [Son23, Section 2].

We will avoid discussing regularity issues for the barycenter map Bar:𝕊+/ΓM\mathrm{Bar}:\mathbb{S}^{+}/\Gamma\to M by only considering its restriction to the supports of “polyhedral chains”, which will be enough in all our applications. A kk-dimensional polyhedral chain PP in S/λΓ(Γ)S^{\infty}/\lambda_{\Gamma}(\Gamma) is by definition a kk-dimensional integral current PP such that there are smoothly embedded totally geodesic kk-simplices S1,,SmS/λΓ(Γ)S_{1},...,S_{m}\subset S^{\infty}/\lambda_{\Gamma}(\Gamma) endowed with an orientation, and integers aja_{j} so that

P=j=1maj1SjP=\sum_{j=1}^{m}a_{j}\llbracket 1_{S_{j}}\rrbracket

(see [Son23, Subsection 1.7]). Given a polyhedral chain PP in 𝕊+/Γ\mathbb{S}^{+}/\Gamma, we can check that the restriction

Bar:spt(P)M\mathrm{Bar}:\operatorname{spt}(P)\to M

is indeed continuous and smooth on each simplex. For 1kn1\leq k\leq n, given a smooth embedding with totally geodesic image φ:k𝕊+S\varphi:\mathbb{R}^{k}\to\mathbb{S}^{+}\subset S^{\infty}, let QQ be the tangent kk-plane at p:=φ(y)p:=\varphi(y) for some yky\in\mathbb{R}^{k}. The map

Bar:φ(k)M~\mathrm{Bar}:\varphi(\mathbb{R}^{k})\to\tilde{M}

is smooth around pp, and its differential along QQ is denoted by dBar|Q:QTBar(p)M~d\mathrm{Bar}\big{|}_{Q}:Q\to T_{\mathrm{Bar}(p)}\tilde{M}. 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 n3n\geq 3. Let f𝕊+f\in\mathbb{S}^{+} and let QQ be the tangent nn-plane at ff of a totally geodesic nn-simplex in 𝕊+\mathbb{S}^{+} passing through ff. Then

(11) |JacBar|Q|(4n(n1)2)n/2.|\operatorname{Jac}\mathrm{Bar}\big{|}_{Q}|\leq\big{(}\frac{{4n}}{(n-1)^{2}}\big{)}^{n/2}.

Moreover for any η>0\eta>0 small enough, there exists cη>0c_{\eta}>0 with limη0cη=0\lim_{\eta\to 0}c_{\eta}=0, such that the following holds. If

|JacBar|Q|(4n(n1)2)n/2η,|\operatorname{Jac}\mathrm{Bar}\big{|}_{Q}|\geq\big{(}\frac{{4n}}{(n-1)^{2}}\big{)}^{n/2}-\eta,

then for any norm 11 tangent vector uQ\vec{u}\in Q,

(12) |dBar|Q(u)|(4n(n1)2)1/2cη|d\mathrm{Bar}\big{|}_{Q}(\vec{u})|\geq\big{(}\frac{4n}{(n-1)^{2}}\big{)}^{1/2}-c_{\eta}

and for any connected continuous piecewise geodesic curve α𝕊+\alpha\subset\mathbb{S}^{+} of length less than η\eta starting at ff, we have

(13) lengthg0(Bar(α))((4n(n1)2)1/2+cη)length(α)\operatorname{length}_{g_{0}}(\mathrm{Bar}(\alpha))\leq(\big{(}\frac{4n}{(n-1)^{2}}\big{)}^{1/2}+c_{\eta})\operatorname{length}(\alpha)

where length(α)\operatorname{length}(\alpha) is computed using the standard round metric on SS^{\infty}.

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 C=(X,d,T)C=(X,d,T) and an oriented, connected, closed Riemannian manifold (N,gN)(N,g_{N}), which induces the integral current space (N,distgN,1N)(N,\operatorname{dist}_{g_{N}},\llbracket 1_{N}\rrbracket). The intrinsic metric on XX induced by dd is denoted by LdL_{d}. Note that the identity map

id:(X,Ld)(X,d)\operatorname{id}:(X,L_{d})\to(X,d)

is always 11-Lipschitz (on each path connected component).

Definition 2.7.

We say that C=(X,d,T)C=(X,d,T) is intrinsically isomorphic to (N,gN)(N,g_{N}) if there is an isometry

φ:(N,distgN)(X,Ld)\varphi:(N,\operatorname{dist}_{g_{N}})\to(X,L_{d})

such that

(idφ)1N=T.(\operatorname{id}\circ\varphi)_{\sharp}\llbracket 1_{N}\rrbracket=T.

For clarity, we emphasize that “being intrinsically isomorphic” is weaker than “being at intrinsic flat distance 0 from each other”.

Our main result in this section shows that in dimensions at least 33, the spherical Plateau solutions for closed hyperbolic manifolds are unique up to intrinsic isomorphism, see Definition 2.7.

Theorem 2.8.

Let (M,g0)(M,g_{0}) be a closed oriented hyperbolic manifold of dimension at least 33. Then any spherical Plateau solution for MM is intrinsically isomorphic to (M,(n1)24ng0)(M,\frac{(n-1)^{2}}{{4n}}g_{0}).

Proof.

Let ϕiM\phi_{i}\in\mathscr{H}_{M} be a minimizing sequence, namely

(14) limiVol(M,ϕi(𝐠Hil))=SphereVol(M)=Vol(M,(n1)24ng0),\lim_{i\to\infty}\operatorname{Vol}(M,\phi_{i}^{*}(\mathbf{g}_{\mathrm{Hil}}))=\operatorname{SphereVol}(M)=\operatorname{Vol}(M,\frac{(n-1)^{2}}{4n}g_{0}),

where the second equality follows from Theorem 2.2. We suppose that the integral currents

Ci:=(ϕi)1MC_{i}:=(\phi_{i})_{\sharp}\llbracket 1_{M}\rrbracket

converge in the intrinsic flat topology to a spherical Plateau solution

C=(X,d,S).C_{\infty}=(X_{\infty},d_{\infty},S_{\infty}).

Set Γ:=π1(M)\Gamma:=\pi_{1}(M). By a perturbation argument, we can assume without loss of generality that for all ii, for all yMy\in M, any lift of ϕi(y)S/λΓ(Γ)\phi_{i}(y)\in S^{\infty}/\lambda_{\Gamma}(\Gamma) in S2(Γ)S^{\infty}\subset\ell^{2}(\Gamma) has finite support. In particular, we can assume that

spt(Ci)𝕊+/Γ\operatorname{spt}(C_{i})\subset\mathbb{S}^{+}/\Gamma

where 𝕊+\mathbb{S}^{+} is defined in Subsection 2.2. By a further perturbation of ϕi\phi_{i}, we can even assume that CiC_{i} is a polyhedral chain (a notion defined in Subsection 2.2), in particular that spt(Ci)\operatorname{spt}(C_{i}) is a finite union of embedded totally geodesic nn-simplices in S/λΓ(Γ)S^{\infty}/\lambda_{\Gamma}(\Gamma), see [Son23, Lemma 1.6].

From now on, we will use the notation

g:=(n1)24ng0.g^{\prime}:=\frac{(n-1)^{2}}{4n}g_{0}.

In the sequel, Jacobians, lengths and distances will be computed with respect to the metric gg^{\prime} on MM. Fix oM~o\in\tilde{M} and let

Bar:𝕊+/ΓM\mathrm{Bar}:\mathbb{S}^{+}/\Gamma\to M

be the barycenter map, see Section 2.2. By Γ\Gamma-equivariance, for any ii, Bar:spt(Ci)M\mathrm{Bar}:\operatorname{spt}(C_{i})\to M is a Lipschitz homotopy equivalence, and

(15) Bar(Ci)=1M.{\mathrm{Bar}}_{\sharp}(C_{i})=\llbracket 1_{M}\rrbracket.

By lower semicontinuity of the mass under intrinsic flat convergence [SW11]:

(16) 𝐌(C)SphereVol(M)=Vol(M,g)\mathbf{M}(C_{\infty})\leq\operatorname{SphereVol}(M)=\operatorname{Vol}(M,g^{\prime})

(the equality above is Theorem 2.2).

The nn-dimensional Jacobian of Bar{\mathrm{Bar}} along the tangent nn-plane of spt(Ci)\operatorname{spt}(C_{i}) at any point qq in the interior of a “face” of spt(Ci)\operatorname{spt}(C_{i}) is well-defined and is bounded from above by 11 with respect to the metric gg^{\prime} on MM, by the main Jacobian bound (11) in Lemma 2.6. This implies by the area formula and (15) that

𝐌(Ci)Vol(M,g)=SphereVol(M).\mathbf{M}(C_{i})\geq\operatorname{Vol}(M,g^{\prime})=\operatorname{SphereVol}(M).

Since CiC_{i} has mass converging to SphereVol(M)\operatorname{SphereVol}(M), by the area formula, the Jacobian of Bar{\mathrm{Bar}} has to be close to 11 on a larger and larger part of spt(Ci)\operatorname{spt}(C_{i}) as ii\to\infty, meaning that there are open subsets Ωi in the smooth part of spt(Ci)\Omega_{i}\text{ in the smooth part of }\operatorname{spt}(C_{i}) such that at every point qΩiq\in\Omega_{i}, there is a well-defined tangent nn-plane of sptCi\operatorname{spt}C_{i} at qq, and

(17) limi𝐌(CiΩi)=limi𝐌(Ci)=SphereVol(M), limiJacBar1L(Ωi)=0,\begin{split}\lim_{i\to\infty}\mathbf{M}(C_{i}\llcorner\Omega_{i})=\lim_{i\to\infty}\mathbf{M}(C_{i})&=\operatorname{SphereVol}(M), \\ \lim_{i\to\infty}\|\operatorname{Jac}{{\mathrm{Bar}}}-1\|_{L^{\infty}(\Omega_{i})}&=0,\\ \end{split}

where we recall that the Jacobian is computed with gg^{\prime} and JacBar\operatorname{Jac}{{\mathrm{Bar}}} denotes the Jacobian along the tangent nn-plane, see Section 1.1. For r>0r>0, set

Ωi,r:=the r-neighborhood of Ωi in 𝕊+/ΓS/λΓ(Γ).\Omega_{i,r}:=\text{the $r$-neighborhood of $\Omega_{i}$ in $\mathbb{S}^{+}/\Gamma\subset S^{\infty}/\lambda_{\Gamma}(\Gamma)$}.

By (17), the coarea formula and Sard’s theorem, after smoothing out the distance function from Ωi\Omega_{i} by a standard argument and still using the notation “Ωi,r\Omega_{i,r}” for the rr-sublevel set of the smoothed out distance function, there are r(i)(0,1)r^{(i)}\in(0,1) such that for each ii,

Di:=CiΩi,r(i)D_{i}:=C_{i}\llcorner\Omega_{i,r^{(i)}}

is an integral current, and spt(Di)\operatorname{spt}(D_{i}) is a compact piecewise smooth submanifold of S/λΓ(Γ)S^{\infty}/\lambda_{\Gamma}(\Gamma) satisfying the following:

  • the boundary of spt(Di)\operatorname{spt}(D_{i}) is piecewise smooth (this is where considering the smoothed out distance function is used) and we have

    (18) limi𝐌(Di)=0.\lim_{i\to\infty}\mathbf{M}(\partial D_{i})=0.
  • after taking a subsequence, DiD_{i} still converges to

    C=(X,d,S)C_{\infty}=(X_{\infty},d_{\infty},S_{\infty})

    in the intrinsic flat topology as ii\to\infty, In particular, there are a Banach space 𝐙\mathbf{Z}^{\prime} and isometric embeddings

    spt(Di)𝐙,sptS𝐙\operatorname{spt}(D_{i})\hookrightarrow\mathbf{Z}^{\prime},\quad\operatorname{spt}S_{\infty}\hookrightarrow\mathbf{Z}^{\prime}

    (with a slight abuse of notations we consider those sets as subsets of 𝐙\mathbf{Z}^{\prime}), such that DiD_{i} converges to SS_{\infty} in the flat topology inside 𝐙\mathbf{Z}^{\prime}.

By (17) and (15),

(19) Bar(Di)converges in the flat topology to 1M. inside (M,g){\mathrm{Bar}}_{\sharp}(D_{i})\quad\text{converges in the flat topology to $\llbracket 1_{M}\rrbracket.$ inside $(M,g^{\prime})$}

Inequality (13) of Lemma 2.6 ensures that a Lipschitz bound holds uniformly in a neighborhood of Ωi\Omega_{i}: for any ϵ>0\epsilon>0, there is rϵ>0r_{\epsilon}>0, such that if ii is large enough, then for fΩif\in\Omega_{i} and f𝕊+/Γf^{\prime}\in\mathbb{S}^{+}/\Gamma joined to ff by a piecewise geodesic curve α𝕊+/Γ\alpha\subset\mathbb{S}^{+}/\Gamma of length at most rϵ>0r_{\epsilon}>0, we have

(20) lengthg(Bar(α))(1+ϵ)length(α).\operatorname{length}_{g^{\prime}}({\mathrm{Bar}}(\alpha))\leq(1+\epsilon)\operatorname{length}(\alpha).

Given f,f𝕊+/Γf,f^{\prime}\in\mathbb{S}^{+}/\Gamma and a curve in S/λΓ(Γ)S^{\infty}/\lambda_{\Gamma}(\Gamma) joining those two elements, after a small perturbation, that curve can be assumed to be inside 𝕊+/Γ\mathbb{S}^{+}/\Gamma. As a consequence of (20), we get the following local Lipschitz bound: for any r~(0,1)\tilde{r}\in(0,1), the restriction of Bar{\mathrm{Bar}} to the subset Ωi,r~\Omega_{i,\tilde{r}} is λ\lambda-Lipschitz for some λ>0\lambda>0 independent of ii. In particular, the restriction

(21) Bar:spt(Di)Mis λ-Lipschitz.{\mathrm{Bar}}:\operatorname{spt}(D_{i})\to M\quad\text{is $\lambda$-Lipschitz.}

We can now check that Assumption 1.3 is verified with (N,h)=(M,g)(N,h)=(M,g^{\prime}), (E,d)=𝐙(E,d)=\mathbf{Z}^{\prime}, Si=DiS_{i}=D_{i}, Ni=sptDiN_{i}=\operatorname{spt}D_{i}, φi=Bar\varphi_{i}=\mathrm{Bar}, Ri=ΩiR_{i}=\Omega_{i}. In particular, in order to check Assumption 1.3 (e), observe that since the Jacobian of Bar\mathrm{Bar} converges to 11 on Ωi\Omega_{i} by (17), Bar\mathrm{Bar} is forced to be almost a Riemannian isometry on Ωi\Omega_{i} 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 Bar:sptS(M,g){\mathrm{Bar}}_{\infty}:\operatorname{spt}S_{\infty}\to(M,g^{\prime}) which is an isometry for the intrinsic metrics induced on sptS\operatorname{spt}S_{\infty} and MM. Moreover by Lemma 1.2 (2), Bar{\mathrm{Bar}}_{\infty} preserves the current structures in the sense that

(Bar)(S)=1M.({\mathrm{Bar}}_{\infty})_{\sharp}(S_{\infty})=\llbracket 1_{M}\rrbracket.

In other words, C=(X,d,S)C_{\infty}=(X_{\infty},d_{\infty},S_{\infty}) is intrinsically isomorphic to (M,g)(M,g^{\prime}), as wanted.

3. The entropy stability theorem

3.1. Technical preparation

As before, MM is the closed, connected, oriented hyperbolic manifold, Γ\Gamma is its fundamental group and SS^{\infty} is the unit sphere inside 2(Γ)\ell^{2}(\Gamma), which is acted upon by Γ\Gamma via the regular representation.

Let us define maps 𝒫c\mathcal{P}_{c} relating the volume entropy of a Riemannian metric on MM and the spherical volume of MM, introduced by Besson-Courtois-Gallot, see [BCG91, Proof of Lemma 3.1]. Let gg be a Riemannian metric on MM. The universal cover of MM is M~\tilde{M} and its fundamental group is Γ\Gamma. Let h(g)h(g) be its volume entropy. Denote by DMD_{M} a Borel fundamental domain in M~\tilde{M} for the action of Γ\Gamma and let γ.DM\gamma.D_{M} be its image by an element γΓ\gamma\in\Gamma. Besson-Courtois-Gallot considered for c>h(g)c>h(g) maps similar to the following:

𝒫c\displaystyle\mathcal{P}_{c} :M~S\displaystyle:\tilde{M}\to S^{\infty}
x{γ1ec2distg(x,.)L2(M~,g)[γ.DMecdistg(x,u)dvolg(u)]1/2}.\displaystyle x\mapsto\{\gamma\mapsto\frac{1}{\|e^{-\frac{c}{2}\operatorname{dist}_{g}(x,.)}\|_{L^{2}(\tilde{M},g)}}\big{[}\int_{\gamma.D_{M}}e^{-c\operatorname{dist}_{g}(x,u)}\operatorname{dvol}_{g}(u)\big{]}^{1/2}\}.

Those maps satisfy the following properties, which hold in any dimension n2n\geq 2:

Lemma 3.1 ([BCG91]).

For a Riemannian metric gg on MM, 𝒫c\mathcal{P}_{c} is a Γ\Gamma-equivariant Lipschitz map, and for almost any xM~x\in\tilde{M}, it satisfies

(22) j=1n|dx𝒫c(ej)|2c24,\sum_{j=1}^{n}|d_{x}{\mathcal{P}}_{c}(e_{j})|^{2}\leq\frac{c^{2}}{4},

where {ej}\{e_{j}\} is a gg-orthonormal basis of TxM~T_{x}\tilde{M}.

Proof.

For the reader’s convenience, let us outline the proof. Consider S2(M~,g)S_{2}(\tilde{M},g) the unit sphere in L2(M~,g)L^{2}(\tilde{M},g). Set for c>h(g)c>h(g):

𝒫¯c:M~S2(M~,g)\overline{\mathcal{P}}_{c}:\tilde{M}\to S_{2}(\tilde{M},g)
𝒫¯c:x{y1ec2distg(x,.)L2(M~,g)ec2distg(x,y)},\overline{\mathcal{P}}_{c}:x\mapsto\{y\mapsto\frac{1}{\|e^{-\frac{c}{2}\operatorname{dist}_{g}(x,.)}\|_{L^{2}(\tilde{M},g)}}e^{-\frac{c}{2}\operatorname{dist}_{g}(x,y)}\},

and set

:S2(M~,g)S\mathcal{I}:S_{2}(\tilde{M},g)\to S^{\infty}
:f{γ[γ.DMf2(u)dvolg(u)]1/2}.\mathcal{I}:f\mapsto\{\gamma\mapsto\big{[}\int_{\gamma.D_{M}}f^{2}(u)\operatorname{dvol}_{g}(u)\big{]}^{1/2}\}.

These maps are manifestly Γ\Gamma-equivariant, and note that 𝒫c=𝒫¯c\mathcal{P}_{c}=\mathcal{I}\circ\overline{\mathcal{P}}_{c}. One easily checks that \mathcal{I} is 11-Lipschitz. To prove the lemma, it remains to study 𝒫¯c\overline{\mathcal{P}}_{c}. By the Pythagorean theorem,

dx𝒫¯cL2(M~,g)2\displaystyle\|d_{x}\overline{\mathcal{P}}_{c}\|^{2}_{L^{2}(\tilde{M},g)} 1ec2distg(x,.)L2(M~,g)2M~dxec2distg(x,y)2dvolg(y)\displaystyle\leq\frac{1}{\|e^{-\frac{c}{2}\operatorname{dist}_{g}(x,.)}\|^{2}_{L^{2}(\tilde{M},g)}}\int_{\tilde{M}}\|d_{x}e^{-\frac{c}{2}\operatorname{dist}_{g}(x,y)}\|^{2}\operatorname{dvol}_{g}(y)
c2/4ec2distg(x,.)L2(M~,g)2M~dxdistg(.,y)2ecdistg(x,y)dvolg(y).\displaystyle\leq\frac{c^{2}/4}{\|e^{-\frac{c}{2}\operatorname{dist}_{g}(x,.)}\|^{2}_{L^{2}(\tilde{M},g)}}\int_{\tilde{M}}\|d_{x}\operatorname{dist}_{g}(.,y)\|^{2}e^{-c\operatorname{dist}_{g}(x,y)}\operatorname{dvol}_{g}(y).

Taking the trace and using that the norm of the gradient of the distance function is well-defined almost everywhere and equal to 11, we get at almost every xMx\in M, in a gg-orthonormal basis {ej}\{e_{j}\} of TxMT_{x}{M}:

j=1ndx𝒫¯c(ej)L2(M~,g)2c24.\sum_{j=1}^{n}\|d_{x}\overline{\mathcal{P}}_{c}(e_{j})\|^{2}_{L^{2}(\tilde{M},g)}\leq\frac{c^{2}}{4}.

This proves the lemma.

If gg is a Riemannian metric on MM, let distg\operatorname{dist}_{g} be the geodesic distance on MM induced by gg. The definition of the standard notions of ϵ\epsilon-isometry, ϵ\epsilon-net can be found in [BBI22, Definition 7.3.27, Definition 1.6.1]. Given Ω\Omega subset of a Riemannian manifold (M,g)(M,g), g|Ωg|_{\Omega} denotes (by a slight abuse of notation) the intrinsic metric induced by the Riemannian metric gg using paths inside Ω\Omega. In general (Ω,g|Ω)(\Omega,g|_{\Omega}) is very different from (Ω,distg|Ω)(\Omega,\operatorname{dist}_{g}|_{\Omega}), where distg|Ω\operatorname{dist}_{g}|_{\Omega} is the restriction of the induced metric distg\operatorname{dist}_{g} of (M,g)(M,g) to Ω\Omega.

The set of admissible maps M\mathscr{H}_{M} was defined in Subsection 2.1. The barycenter map Bar:𝕊+/ΓM\mathrm{Bar}:\mathbb{S}^{+}/\Gamma\to M 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 (M,g0)(M,g_{0}) be a closed oriented hyperbolic manifold of dimension n3n\geq 3. Let gig_{i} (i1)(i\geq 1) be a sequence of Riemannian metrics on MM of same volume as g0g_{0}, and suppose that

limih(gi)=h(g0)=n1.\lim_{i\to\infty}h(g_{i})=h(g_{0})=n-1.

Then, there are smooth open subsets AiMA_{i}\subset M such that the following holds after taking a subsequence:

  1. (1)

    limiVol(Ai,gi)=Vol(M,g0)\lim_{i\to\infty}\operatorname{Vol}(A_{i},g_{i})=\operatorname{Vol}(M,g_{0}) and limiArea(Ai,gi)=0\lim_{i\to\infty}\operatorname{Area}(\partial A_{i},g_{i})=0,

  2. (2)

    (Ai,gi|Ai)(A_{i},g_{i}|_{A_{i}}) converges in the intrinsic flat topology to an integral current space

    C=(X,d,S),C_{\infty}=(X_{\infty},d_{\infty},S_{\infty}),
  3. (3)

    there is a bi-Lipschitz bijection

    Ψ:(M,g0)(sptS,d)\Psi:(M,g_{0})\to(\operatorname{spt}S_{\infty},d_{\infty})

    which is 11-Lipschitz, and

    Ψ(1M)=S.\Psi_{\sharp}(\llbracket 1_{M}\rrbracket)=S_{\infty}.
  4. (4)

    (Ai,gi|Ai)(A_{i},g_{i}|_{A_{i}}) converges to (sptS,d)(\operatorname{spt}S_{\infty},d_{\infty}) in the Gromov-Hausdorff topology. Moreover, for any ε>0\varepsilon>0, for all ii large enough, there is a homotopy equivalence

    fi:MsptSf_{i}:M\to\operatorname{spt}S_{\infty}

    such that the restriction fi:(Ai,gi|Ai)(sptS,d)f_{i}:(A_{i},{g_{i}}|_{A_{i}})\to(\operatorname{spt}S_{\infty},d_{\infty}) is an ε\varepsilon-isometry.

Proof.

Step 1: Finding good subsets

For technical convenience, set

g:=(n1)24ng0,gi:=(n1)24ng0.g^{\prime}:=\frac{(n-1)^{2}}{4n}g_{0},\quad g^{\prime}_{i}:=\frac{(n-1)^{2}}{4n}g_{0}.

Note that after rescaling,

h(g)=2n.h(g^{\prime})=2\sqrt{n}.

By our assumptions, there is a sequence ci>h(gi)c^{\prime}_{i}>h(g^{\prime}_{i}) of positive numbers such that

(23) limici=2n.\lim_{i\to\infty}c^{\prime}_{i}=2\sqrt{n}.

By Lemma 3.1, the maps

𝒫ci:M~S\mathcal{P}_{c^{\prime}_{i}}:\tilde{M}\to S^{\infty}

are Γ\Gamma-equivariant. The quotient maps

𝒫ci:(M,gi)S/λΓ(Γ)\mathcal{P}_{c^{\prime}_{i}}:(M,g^{\prime}_{i})\to S^{\infty}/\lambda_{\Gamma}(\Gamma)

can be perturbed to be smooth immersions. Those new maps now belong to M\mathscr{H}_{M}. After a further small perturbation, we obtain homotopic smooth immersions

ϕiM\phi_{i}\in\mathscr{H}_{M}

sending (M,gi)(M,g^{\prime}_{i}) inside 𝕊+/Γ\mathbb{S}^{+}/\Gamma, see [Son23, Lemma 1.6]. Moreover, by (22) and (23), it is not hard to ensure that after those standard smoothings, for all xMx\in M:

(24) j=1n|dxϕi(ej)|2n+νi\sum_{j=1}^{n}|d_{x}\phi_{i}(e^{\prime}_{j})|^{2}\leq n+\nu_{i}

for some positive νi0\nu_{i}\to 0 (with respect to gig^{\prime}_{i}), where {ej}\{e^{\prime}_{j}\} is an orthonormal basis for gig^{\prime}_{i}. By (24) and the inequality of arithmetic and geometric means,

(25) |Jacϕi|(1+νin)n/2|\operatorname{Jac}\phi_{i}|\leq(1+\frac{\nu_{i}}{n})^{n/2}

where the Jacobian is computed with respect to gig^{\prime}_{i}. By Theorem 2.2,

SphereVol(M)=Vol(M,g),\operatorname{SphereVol}(M)=\operatorname{Vol}(M,g^{\prime}),

on the other hand we have Vol(M,gi)=Vol(M,g)\operatorname{Vol}(M,g^{\prime}_{i})=\operatorname{Vol}(M,g^{\prime}) by assumption. Hence, by (25), |Jacϕi|gi|\operatorname{Jac}\phi_{i}|_{g^{\prime}_{i}} converges to 11 on an open region Ω^iM\hat{\Omega}_{i}\subset M with

limiVol(Ω^i,gi)=Vol(M,g),\lim_{i\to\infty}\operatorname{Vol}(\hat{\Omega}_{i},g^{\prime}_{i})=\operatorname{Vol}(M,g^{\prime}),

which by (24) forces

(26) limiu,v=1n|𝐠Hil(dϕi(eu),dϕi(ev))δuv|L(Ω^i)=0\lim_{i\to\infty}\big{\|}\sum_{u,v=1}^{n}|\mathbf{g}_{\mathrm{Hil}}(d\phi_{i}(e^{\prime}_{u}),d\phi_{i}(e^{\prime}_{v}))-\delta_{uv}|\big{\|}_{L^{\infty}(\hat{\Omega}_{i})}=0

where 𝐠Hil\mathbf{g}_{\mathrm{Hil}} is the standard Hilbert Riemannian metric on the spherical quotient S/λΓ(Γ)S^{\infty}/\lambda_{\Gamma}(\Gamma), and {eu}u=1n\{e^{\prime}_{u}\}_{u=1}^{n} denotes any choice of orthonormal bases for the tangent spaces of (M,gi)(M,g^{\prime}_{i}).

Exactly as in the proof of Theorem 2.8 and using (26), we first find open subsets ΩiM\Omega_{i}\subset M with

(27) limiVol(Ωi,gi)=Vol(M,g),\lim_{i\to\infty}\operatorname{Vol}(\Omega_{i},g^{\prime}_{i})=\operatorname{Vol}(M,g^{\prime}),

which satisfy

limiu,v=1n|g(d(Barϕi)(eu),d(Barϕi)(ev))δuv|L(Ωi)=0.\lim_{i\to\infty}\big{\|}\sum_{u,v=1}^{n}|g^{\prime}(d(\mathrm{Bar}\circ\phi_{i})(e^{\prime}_{u}),d(\mathrm{Bar}\circ\phi_{i})(e^{\prime}_{v}))-\delta_{uv}|\big{\|}_{L^{\infty}(\Omega_{i})}=0.

Then we define smoothings of r(i)r^{(i)}-neighborhoods of Ωi\Omega_{i} in (M,gi)(M,g^{\prime}_{i}), called Ωi,r(i)\Omega_{i,r^{(i)}}, so that the closure of Ωi,r(i)\Omega_{i,r^{(i)}} is a compact manifold with a smooth boundary whose area Area(Ωi,r(i),gi)\operatorname{Area}(\partial\Omega_{i,r^{(i)}},g^{\prime}_{i}) goes to 0 as ii\to\infty, and the restriction Barϕi|Ωi,r(i)\mathrm{Bar}\circ\phi_{i}|_{\Omega_{i,r^{(i)}}} is uniformly Lipschitz.

Step 2: Constructing the limit map

We set

(Ni,hi):=(Ωi,r(i),gi|Ωi,r(i)).(N_{i},h_{i}):=(\Omega_{i,r^{(i)}},g^{\prime}_{i}|_{\Omega_{i,r^{(i)}}}).

In order to apply Wenger’s compactness theorem, we need a uniform diameter bound. For that reason, if disthi\operatorname{dist}_{h_{i}} denotes the intrinsic metric induced by gig^{\prime}_{i} using paths contained in NiN_{i}, we set

d^i:=min{disthi,6diam(M,g)}.\hat{d}_{i}:=\min\{\operatorname{dist}_{h_{i}},6\operatorname{diam}(M,g^{\prime})\}.

This defines a metric on NiN_{i} with diameter at most 6diam(M,g)6\operatorname{diam}(M,g^{\prime}), and it is locally isometric to the induced intrinsic metric hih_{i}. We then set

Di:=1Ni.D_{i}:=\llbracket 1_{N_{i}}\rrbracket.

By Wenger’s compactness theorem, the integral current spaces DiD_{i} converge to an integral current space

C^=(X^,d^,S^)\hat{C}_{\infty}=(\hat{X}_{\infty},\hat{d}_{\infty},\hat{S}_{\infty})

in the intrinsic flat topology, after picking a subsequence if necessary. In particular, there are a Banach space 𝐙^\hat{\mathbf{Z}}, and isometric embeddings

(28) (Ni,d^i)𝐙^,sptS^𝐙^,(N_{i},\hat{d}_{i})\hookrightarrow\hat{\mathbf{Z}},\quad\operatorname{spt}\hat{S}_{\infty}\hookrightarrow\hat{\mathbf{Z}},

with the usual abuse of notations, such that 1Ni\llbracket 1_{N_{i}}\rrbracket converges to S^\hat{S}_{\infty} in the flat topology inside 𝐙^\hat{\mathbf{Z}}.

Next, we check that Assumption 1.3 is satisfied for

(N,h)=(M,g),Si=Di,(Ni,hi)=(Ωi,r(i),gi|Ωi,r(i)),(N,h)=(M,g^{\prime}),\quad S_{i}=D_{i},\quad(N_{i},h_{i})=(\Omega_{i,r^{(i)}},g^{\prime}_{i}|_{\Omega_{i,r^{(i)}}}),
φi=Barϕi,Ri=Ωi\varphi_{i}=\mathrm{Bar}\circ\phi_{i},\quad R_{i}=\Omega_{i}...

(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

φ:(sptS^,d^)(M,g)\varphi_{\infty}:(\operatorname{spt}\hat{S}_{\infty},\hat{d}_{\infty})\to(M,g^{\prime})

which is Lipschitz, bijective and whose inverse

Ψ^:=φ1\hat{\Psi}:=\varphi_{\infty}^{-1}

is 11-Lipschitz with respect to the intrinsic metrics. Hence, Ψ^\hat{\Psi} is clearly 11-Lipschitz and bi-Lipschitz. By Lemma 1.2 (2), Ψ^(1M)=S^\hat{\Psi}_{\sharp}(\llbracket 1_{M}\rrbracket)=\hat{S}_{\infty}.

Step 3: Convergence for the original induced metric

We also need to check that (Ni,disthi,1Ni)(N_{i},\operatorname{dist}_{h_{i}},\llbracket 1_{N_{i}}\rrbracket), not just (Ni,d^i,1Ni)(N_{i},\hat{d}_{i},\llbracket 1_{N_{i}}\rrbracket), subsequentially converges to the integral current space C^\hat{C}_{\infty}. Notice that for any xNix\in N_{i} and R(0,3diam(M,g))R\in(0,3\operatorname{diam}(M,g^{\prime})), the metric balls Bdisthi(x,R)(Ni,disthi)B_{\operatorname{dist}_{h_{i}}}(x,R)\subset(N_{i},\operatorname{dist}_{h^{\prime}_{i}}) and Bd^i(x,R)(Ni,d^i)B_{\hat{d}_{i}}(x,R)\subset(N_{i},{\hat{d}_{i}}) are globally isometric. In particular, since (sptS^,d)(\operatorname{spt}\hat{S}_{\infty},d_{\infty}) has diameter at most that of (M,g)(M,g^{\prime}) by 11-Lipschitzness of Ψ\Psi, if OrO_{r} denotes the rr-neighborhood of sptS^\operatorname{spt}\hat{S}_{\infty} in 𝐙^\hat{\mathbf{Z}}, we have: whenever r(0,diam(M,g))r\in(0,\operatorname{diam}(M,g^{\prime})), for any ii and pair of points x,yNiOrx,y\in N_{i}\cap O_{r},

(29) disthi(x,y)=d^i(x,y).\operatorname{dist}_{h_{i}}(x,y)=\hat{d}_{i}(x,y).

By the slicing theorem for metric currents, we can choose for each ii, some radius ri(0,diam(M,g))r_{i}\in(0,\operatorname{diam}(M,g^{\prime})) converging to 0, such that if we set

O~i:=OriNi𝐙^,\tilde{O}_{i}:=O_{r_{i}}\cap N_{i}\subset\hat{\mathbf{Z}},

then 1O~i\llbracket 1_{\tilde{O}_{i}}\rrbracket are integral currents in 𝐙^\hat{\mathbf{Z}} converging to S^\hat{S}_{\infty} in the flat topology. By (29), this means that (O~i,disthi|O~i,1O~i)(\tilde{O}_{i},\operatorname{dist}_{h_{i}}|_{\tilde{O}_{i}},\llbracket 1_{\tilde{O}_{i}}\rrbracket) converges to C^\hat{C}_{\infty} in the intrinsic flat topology. We deduce in particular that the push-forward of 1O~i\llbracket 1_{\tilde{O}_{i}}\rrbracket by Barϕi\mathrm{Bar}\circ\phi_{i} converges to 1M\llbracket 1_{M}\rrbracket as currents in (M,g)(M,g^{\prime}). Then the liminf as ii\to\infty of the mass of this push-forward is at least Vol(M,g)\operatorname{Vol}(M,g^{\prime}) by lower semicontinuity of the mass. By the Jacobian bounds (25), (11), and since by (27) we have limiVol(Ni,hi)=Vol(M,g),\lim_{i\to\infty}\operatorname{Vol}(N_{i},h_{i})=\operatorname{Vol}(M,g^{\prime}),

(30) limiVol(O~i,hi)=Vol(M,g),limiVol(NiO~i,hi)=0.\lim_{i\to\infty}\operatorname{Vol}(\tilde{O}_{i},h_{i})=\operatorname{Vol}(M,g^{\prime}),\quad\lim_{i\to\infty}\operatorname{Vol}(N_{i}\setminus\tilde{O}_{i},h_{i})=0.

We conclude that (Ni,disthi,1Ni)(N_{i},\operatorname{dist}_{h_{i}},\llbracket 1_{N_{i}}\rrbracket) converges to the same limit as (O~i,disthi|O~i,1O~i)(\tilde{O}_{i},\operatorname{dist}_{h_{i}}|_{\tilde{O}_{i}},\llbracket 1_{\tilde{O}_{i}}\rrbracket) in the intrinsic flat topology, which is C^\hat{C}_{\infty}, as desired.

Step 4: Gromov-Hausdorff convergence and ε\varepsilon-isometries

In general, (Ni,disthi)(N_{i},\operatorname{dist}_{h_{i}}) does not converge in the Gromov-Hausdorff topology to (sptS^,d)(\operatorname{spt}\hat{S}_{\infty},d_{\infty}). The end of the proof is about fixing this issue. By (28), there are finite subsets ΣiNi\Sigma_{i}\subset N_{i} converging in the Hausdorff topology to sptS^\operatorname{spt}\hat{S}_{\infty} in 𝐙^\hat{\mathbf{Z}}. For any t>0t>0, let

Σi,t:=t-neighborhood of Σi in (Ni,disthi).\Sigma_{i,t}:=\text{$t$-neighborhood of $\Sigma_{i}$ in $(N_{i},\operatorname{dist}_{h_{i}})$}.

By lower semicontinuity of the mass and (28), for any s1>2t>0s_{1}>2t>0 and any sequence of points piΣi,tp_{i}\in\Sigma_{i,t},

(31) lim infiVol(Bdisthi|Σi,t(pi,s1),hi)>κ0(s1)>0\liminf_{i\to\infty}\operatorname{Vol}(B_{\operatorname{dist}_{h_{i}}|_{\Sigma_{i,t}}}(p_{i},s_{1}),h_{i})>\kappa_{0}(s_{1})>0

for some κ0(s1)\kappa_{0}(s_{1}) not depending on tt. We also have the following stronger property: for any s1>2t>0s_{1}>2t>0 and any sequence of points piΣi,tp_{i}\in\Sigma_{i,t},

(32) lim infiVol(Bhi|Σi,t(pi,s1),hi)>κ(s1)>0\liminf_{i\to\infty}\operatorname{Vol}(B_{h_{i}|_{\Sigma_{i,t}}}(p_{i},s_{1}),h_{i})>\kappa(s_{1})>0

for some κ(s1)\kappa(s_{1}) not depending on tt. Note that this is indeed a stronger inequality, since hi|Σi,t{h_{i}|_{\Sigma_{i,t}}} is the intrinsic metric on Σi,t\Sigma_{i,t} induced by hih_{i} using paths inside Σi,t\Sigma_{i,t}, and

Bhi|Σi,t(pi,s1)Bdisthi|Σi,t(pi,s1).B_{h_{i}|_{\Sigma_{i,t}}}(p_{i},s_{1})\subset B_{\operatorname{dist}_{h_{i}}|_{\Sigma_{i,t}}}(p_{i},s_{1}).

To check this stronger property, recall that sptS^\operatorname{spt}\hat{S}_{\infty} has been shown to be bi-Lipschitz to the closed Riemannian manifold (M,g)(M,g^{\prime}) via a map φ\varphi_{\infty}. For any two points a,bΣi,ta,b\in\Sigma_{i,t}, we can find a,bsptS^𝐙^a^{\prime},b^{\prime}\in\operatorname{spt}\hat{S}_{\infty}\subset\hat{\mathbf{Z}} approximating a,ba,b. Then, given a minimizing geodesic segment γ\gamma in (M,g)(M,g^{\prime}) between φ(a),φ(b)\varphi_{\infty}(a^{\prime}),\varphi_{\infty}(b^{\prime}), we can approximate (φ)1(γ)(\varphi_{\infty})^{-1}(\gamma) by a curve in Σi,t\Sigma_{i,t} between a,ba,b without increasing the length by more than a constant factor. Hence, for ii large,

Bdisthi|Σi,t(pi,λ0s1)Bhi|Σi,t(pi,s1)B_{\operatorname{dist}_{h_{i}}|_{\Sigma_{i,t}}}(p_{i},\lambda_{0}s_{1})\subset B_{h_{i}|_{\Sigma_{i,t}}}(p_{i},s_{1})

for some λ0(0,1)\lambda_{0}\in(0,1) independent of tt. This and (31) explain (32).

By (30) and by the coarea formula, there is a t>0t>0, arbitrarily small, such that

limiVol(Σi,t,hi)=Vol(M,g),limiArea(Σi,t,hi)=0.\lim_{i\to\infty}\operatorname{Vol}(\Sigma_{i,t},h_{i})=\operatorname{Vol}(M,g^{\prime}),\quad\lim_{i\to\infty}\operatorname{Area}(\partial\Sigma_{i,t},h_{i})=0.

This means that after taking a subsequence, we find ti>0t_{i}>0 converging to 0 so that if we set

Ai:=Σi,tiA_{i}:=\Sigma_{i,t_{i}}

then for any s1>0s_{1}>0 and any sequence of points piAip_{i}\in A_{i},

(33) lim infiVol(Bhi|Ai(pi,s1),hi)>0,\liminf_{i\to\infty}\operatorname{Vol}(B_{h_{i}|_{A_{i}}}(p_{i},s_{1}),h_{i})>0,
limiVol(Ai,hi)=Vol(M,g),\lim_{i\to\infty}\operatorname{Vol}(A_{i},h_{i})=\operatorname{Vol}(M,g^{\prime}),
limiArea(Ai,hi)=0.\lim_{i\to\infty}\operatorname{Area}(\partial A_{i},h_{i})=0.

Now we can reapply all the arguments in Step 2 and Step 3 to a smoothing of Ai=Σi,tiA_{i}=\Sigma_{i,{t_{i}}} instead of NiN_{i}. Let us summarize what we have achieved so far: subsequentially, (Ai,gi|Ai)(A_{i},g^{\prime}_{i}|_{A_{i}}) converges to an integral current space

C=(X,d,S),C_{\infty}=(X_{\infty},d_{\infty},S_{\infty}),

and there are a Banach space 𝐙\mathbf{Z}, and isometric embeddings

(34) (Ai,gi|Ai)𝐙,sptS𝐙,(A_{i},g^{\prime}_{i}|_{A_{i}})\hookrightarrow\mathbf{Z},\quad\operatorname{spt}S_{\infty}\hookrightarrow\mathbf{Z},

with the usual abuse of notations, such that 1Ai\llbracket 1_{A_{i}}\rrbracket converges to SS_{\infty} in the flat topology inside 𝐙\mathbf{Z}. Moreover, there is a bi-Lipschitz, 11-Lipschitz map

Ψ:(M,g)(sptS,d)\Psi:(M,g^{\prime})\to(\operatorname{spt}S_{\infty},d_{\infty})

which is the inverse of a limit map constructed using Lemma 1.2 applied to Barϕi\mathrm{Bar}\circ\phi_{i}. The following analogue of (30) holds: for any r>0r>0, if OrO_{r} is the rr-neighborhood of sptS\operatorname{spt}S_{\infty} in 𝐙\mathbf{Z},

(35) limiVol(AiOr,gi|Ai)=0.\lim_{i\to\infty}\operatorname{Vol}(A_{i}\setminus O_{r},g^{\prime}_{i}|_{A_{i}})=0.

The key additional property we gained is that (Ai,gi|Ai)(A_{i},g^{\prime}_{i}|_{A_{i}}) now converges to sptS\operatorname{spt}S_{\infty} in the Hausdorff topology inside 𝐙\mathbf{Z}, by (33) and (35). Note that in general, sptS\operatorname{spt}S_{\infty} and the previous space sptS^\operatorname{spt}\hat{S}_{\infty} could be very different.

We can then set

fi:=ΨBarϕi:MsptS,f_{i}:=\Psi\circ\mathrm{Bar}\circ\phi_{i}:M\to\operatorname{spt}S_{\infty},

which is a homotopy equivalence. By Lemma 1.2 (1), we conclude that for any ε>0\varepsilon>0,

fi:(Ai,gi|Ai)(sptS,d)f_{i}:(A_{i},g^{\prime}_{i}|_{A_{i}})\to(\operatorname{spt}S_{\infty},d_{\infty})

is an ε\varepsilon-isometry if ii is large. All of these complete the proof, after rescaling all the Riemannian metrics by 4n(n1)2\frac{4n}{(n-1)^{2}}. ∎

3.2. Equidistribution of geodesic spheres in hyperbolic manifolds

Consider (M,g0)(M,g_{0}) a closed hyperbolic manifold, with universal cover M~\tilde{M}. Fix 𝐱M\mathbf{x}\in M and let 𝐱~\tilde{\mathbf{x}} be a lift of 𝐱\mathbf{x} by the natural projection M~M\tilde{M}\to M. Let T1MT^{1}M denote the unit tangent bundle of MM. Let S~(𝐱~,t)\tilde{S}(\tilde{\mathbf{x}},t) be the geodesic sphere of radius tt centered at 𝐱~\tilde{\mathbf{x}} in M~\tilde{M}, and let S~1(𝐱~,t)\tilde{S}_{1}(\tilde{\mathbf{x}},t) be its lift to the unit tangent bundle T1M~T^{1}\tilde{M} by considering the outward unit normal vectors on S~(𝐱~,t)\tilde{S}(\tilde{\mathbf{x}},t). Let S(𝐱,t)S(\mathbf{x},t) denote the projection of S~(x~,t)\tilde{S}(\tilde{x},t) in MM, and let S1(𝐱,t)S_{1}(\mathbf{x},t) be the projection of S~1(𝐱~,t)\tilde{S}_{1}(\tilde{\mathbf{x}},t) to the unit tangent bundle T1MT^{1}M. A measure on T1MT^{1}M (resp. S1(x0,t){S}_{1}({x}_{0},t)) is called invariant if it is induced by a measure on T1M~T^{1}\tilde{M} invariant by isometries of M~\tilde{M} (resp. induced by a measure on S~1(𝐱~,t)\tilde{S}_{1}(\tilde{\mathbf{x}},t) invariant by rotations of center 𝐱~\tilde{\mathbf{x}} in M~\tilde{M}).

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 f:T1Mf:T^{1}M\to\mathbb{R},

limtS1(x0,t)f(y)𝑑μt(y)=T1Mf(y)𝑑vT1M(y),\lim_{t\to\infty}\int_{{S}_{1}({x}_{0},t)}f(y)d\mu_{t}(y)=\int_{T^{1}M}f(y)dv_{T^{1}M}(y),

where dμtd\mu_{t} is the unique invariant probability measure on S1(x0,t){S}_{1}({x}_{0},t) and dvT1Mdv_{T^{1}M} is the unique invariant probability measure on T1MT^{1}M.

Below, areas (namely (n1)(n-1)-dimensional Hausdorff measures) and lengths are computed using g0g_{0}. Given an open subset UMU\subset M, let π1(M,U)\pi_{1}(M,U) denote the relative homotopy group. Consider a (not necessarily length minimizing) geodesic segment σ\sigma in (M,g0)(M,g_{0}) with two different endpoints x,yMx,y\in M and let Ux,UyU_{x},U_{y} be two disjoint open geodesic balls centered at xx and yy. Fix 𝐱~M~\tilde{\mathbf{x}}\in\tilde{M} as before. Let

π:M~M\pi:\tilde{M}\to M

be the natural projection.

Corollary 3.4.

There is θ(0,1)\theta\in(0,1) depending on M,σ,Ux,UyM,\sigma,U_{x},U_{y} such that for all tt large enough, there is an open subset WtS~(𝐱~,t)W_{t}\subset\tilde{S}(\tilde{\mathbf{x}},t) satisfying

Area(Wt)θArea(S~(𝐱~,t))\operatorname{Area}(W_{t})\geq\theta\operatorname{Area}(\tilde{S}(\tilde{\mathbf{x}},t))

and with the following property: for any zWtz\in W_{t}, if l:[0,t]M~l:[0,t]\to\tilde{M} denotes the length minimizing geodesic from 𝐱~\tilde{\mathbf{x}} to zz in (M~,g0)(\tilde{M},g_{0}) parametrized by arclength, there are disjoint intervals

[a1,b1],,[am,bm][0,t][a_{1},b_{1}],...,[a_{m},b_{m}]\subset[0,t]

such that

  1. (1)

    j=1m|bjaj|θt\sum_{j=1}^{m}|b_{j}-a_{j}|\geq\theta t,

  2. (2)

    for j{1,,m}j\in\{1,...,m\}, the endpoints satisfy π(l(aj))Ux\pi(l(a_{j}))\in U_{x} and π(l(bj))Uy\pi(l(b_{j}))\in U_{y},

  3. (3)

    for j{1,,m}j\in\{1,...,m\}, πl:[aj,bj]M\pi\circ l:[a_{j},b_{j}]\to M is a geodesic segment joining π(l(aj))\pi(l(a_{j})) to π(l(bj))\pi(l(b_{j})), which is in the same class as σ\sigma in π1(M,UxUy)\pi_{1}(M,U_{x}\cup U_{y}).

Proof.

Let t0t_{0} be the length of σ\sigma. By continuity, there exist an open subset OO of the unit tangent bundle T1MT^{1}M depending only on M,Ux,UyM,U_{x},U_{y}, such that for any tangent vector vv in OO, the basepoint pp of vv lies in UxU_{x}, and the geodesic γ\gamma starting at pp with direction vv and length t0t_{0} ends at a point qUyq\in U_{y}, and satisfies the following:

γ[σ]π1(M,UxUy).\gamma\in[\sigma]\in\pi_{1}(M,U_{x}\cup U_{y}).

Informally, geodesics of length t0t_{0} starting at a vector in OO stay “close” to σ\sigma.

Let μ~t\tilde{\mu}_{t} be the invariant probability measure on S~1(𝐱~,t)\tilde{S}_{1}(\tilde{\mathbf{x}},t) and set

O~:=π1(O)T1M~,O~t:=O~S~1(𝐱~,t).\tilde{O}:=\pi^{-1}(O)\subset T^{1}\tilde{M},\quad\tilde{O}_{t}:=\tilde{O}\cap\tilde{S}_{1}(\tilde{\mathbf{x}},t).

Recall that S~1(𝐱~,t)\tilde{S}_{1}(\tilde{\mathbf{x}},t) is the lift of the sphere S~(𝐱~,t)\tilde{S}(\tilde{\mathbf{x}},t) by its normal unit vector. Below, by abuse of notations, we will identify S~1(𝐱~,t)\tilde{S}_{1}(\tilde{\mathbf{x}},t) and S~(𝐱~,t)\tilde{S}(\tilde{\mathbf{x}},t). By applying Theorem 3.3 to the characteristic function of OO, for all tt large enough,

(36) μ~t(O~t)>c1>0\tilde{\mu}_{t}(\tilde{O}_{t})>c_{1}>0

for some c1c_{1} independent of tt. If xS~1(𝐱~,t)x\in\tilde{S}_{1}(\tilde{\mathbf{x}},t), let τx\tau_{x} be the unique geodesic segment from the basepoint 𝐱~\tilde{\mathbf{x}} to xx in M~\tilde{M}.

We claim that for some c2,c3>0c_{2},c_{3}>0, for any large integer NN,

μ~Nt0(\displaystyle\tilde{\mu}_{Nt_{0}}( {xS~1(𝐱~,Nt0): for at least c2N distinct k{1,,N}τxO~kt0})>c3.\displaystyle\{x\in\tilde{S}_{1}(\tilde{\mathbf{x}},Nt_{0}):\text{ for at least $c_{2}N$ distinct $k\in\{1,...,N\}$, $\tau_{x}\cap\tilde{O}_{kt_{0}}\neq\varnothing$}\})>c_{3}.

Roughly speaking, this inequality means that for a uniformly positive fraction of the sphere S~1(𝐱~,Nt0)\tilde{S}_{1}(\tilde{\mathbf{x}},Nt_{0}), geodesics from 𝐱~\tilde{\mathbf{x}} to that portion of S~1(𝐱~,Nt0)\tilde{S}_{1}(\tilde{\mathbf{x}},Nt_{0}) stay close to σ\sigma on a uniformly positive fraction of their length. Before proving the claim, note that S~1(𝐱~,t)\tilde{S}_{1}(\tilde{\mathbf{x}},t) is a sphere parametrized by S2S^{2} via the exponential map with basepoint x~M~\tilde{x}\in\tilde{M}, and that the measure on S2S^{2} corresponding to μ~t\tilde{\mu}_{t} is just the standard uniform probability measure dνS2d\nu_{S^{2}}. We let χO~t:S2{0,1}\chi_{\tilde{O}_{t}}:S^{2}\to\{0,1\} be the characteristic function of the subset corresponding to O~t\tilde{O}_{t} and we compute for any large NN:

c1<1Nk=1Nμ~kt0(O~kt0)=1Nk=1NS2χO~kt0𝑑νS2=S21Nk=1NχO~kt0dνS2c_{1}<\frac{1}{N}\sum_{k=1}^{N}\tilde{\mu}_{kt_{0}}(\tilde{O}_{kt_{0}})=\frac{1}{N}\sum_{k=1}^{N}\int_{S^{2}}\chi_{\tilde{O}_{kt_{0}}}d\nu_{S^{2}}=\int_{S^{2}}\frac{1}{N}\sum_{k=1}^{N}\chi_{\tilde{O}_{kt_{0}}}d\nu_{S^{2}}

where the first inequality follows from (36). So there are c2,c3>0c_{2},c_{3}>0, for any large NN, on some subset of S2S^{2} of dνS2d\nu_{S^{2}}-measure at least c3c_{3},

1Nk=1NχO~kt0>c2\frac{1}{N}\sum_{k=1}^{N}\chi_{\tilde{O}_{kt_{0}}}>c_{2}

which is exactly the claim. ∎

3.3. From equidistribution of geodesic spheres to intrinsic flat stability

Let (M,g0)(M,g_{0}) be a closed oriented hyperbolic manifold of dimension nn. 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 gig_{i} on MM approximates a metric space which is metrically dominated by (M,g0)(M,g_{0}), then the volume entropy of gig_{i} is eventually strictly larger than that of g0g_{0}. Its proof relies on the equidistribution of geodesic spheres in hyperbolic manifolds, Theorem 3.3.

Theorem 3.5.

Let (M,g0)(M,g_{0}) be a closed oriented hyperbolic manifold of dimension n2n\geq 2. Suppose that the following holds:

  1. (1)

    there is a metric dd on MM such that there is a bi-Lipschitz bijection

    Ψ:(M,g0)(M,d)\Psi:(M,g_{0})\to(M,d)

    which is 11-Lipschitz,

  2. (2)

    there are Riemannian metrics gig_{i} (i1i\geq 1) on MM so that for any ε>0\varepsilon>0, for all ii large enough, there are an open subset AiMA_{i}\subset M, and a homotopy equivalence

    fi:MMf_{i}:M\to M

    such that the restriction fi:(Ai,gi|Ai)(M,d)f_{i}:(A_{i},g_{i}|_{A_{i}})\to(M,d) is an ε\varepsilon-isometry.

Then, if Ψ\Psi is not an isometry, we have

lim infih(gi)>h(g0).\liminf_{i\to\infty}h(g_{i})>h(g_{0}).
Remark 3.6.

We emphasize that (Ai,gi|Ai)(A_{i},g_{i}|_{A_{i}}) denotes the metric space whose metric is induced by gig_{i} using paths in AiA_{i}. In particular, it is not in general isometric to (Ai,distgi|Ai)(A_{i},\operatorname{dist}_{g_{i}}|_{A_{i}}), where distgi|Ai\operatorname{dist}_{g_{i}}|_{A_{i}} is the restriction of distgi\operatorname{dist}_{g_{i}} to the subset AiA_{i}.

Proof.

Consider small positive numbers η,ε(0,1)\eta,\varepsilon\in(0,1) to be fixed later, and consider ii large enough so that there is AiMA_{i}\subset M and a homotopy equivalence fi:MMf_{i}:M\to M whose restriction

fi:(Ai,gi|Ai)(M,d)f_{i}:(A_{i},g_{i}|_{A_{i}})\to(M,d)

is an ε\varepsilon-isometry, as in condition (2).

Let us then define “fif_{i}-lifts”. Given a point p(M,g0)p\in(M,g_{0}), we say that piAip_{i}\in A_{i} is an fif_{i}-lift of pp if Ψ1(fi(pi))\Psi^{-1}(f_{i}(p_{i})) is η\eta-close to pp with respect to the hyperbolic metric g0g_{0}. Given a g0g_{0}-geodesic segment σ\sigma in (M,g0)(M,g_{0}) with endpoints s,ts,t (which is parametrized by arclength), we say that a curve σi\sigma_{i} with endpoints si,tis_{i},t_{i} in (Ai,gi|Ai)(A_{i},g_{i}|_{A_{i}}) is a fif_{i}-lift of σ\sigma if

  • si,tis_{i},t_{i} are fif_{i}-lifts of s,ts,t,

  • lengthgi(σi)(1+η)lengthg0(σ),\operatorname{length}_{g_{i}}(\sigma_{i})\leq(1+\eta)\operatorname{length}_{g_{0}}(\sigma),

  • Ψ1(fi(σi))[σ]π1(M,Bg0(s,η)Bg0(t,η))\Psi^{-1}(f_{i}(\sigma_{i}))\in[\sigma]\in\pi_{1}(M,B_{g_{0}}(s,\eta)\cup B_{g_{0}}(t,\eta)) where Bg0B_{g_{0}} means g0g_{0}-geodesic ball.

By basic properties of ε\varepsilon-isometries [BBI22, Exercise 7.5.11] and since the bi-Lipschitz bijection Ψ\Psi is 11-Lipschitz, for any η\eta, whenever ε\varepsilon is small enough compared to η\eta and the injectivity radius of (M,g0)(M,g_{0}), any g0g_{0}-geodesic segment σ\sigma in (M,g0)(M,g_{0}) admits an fif_{i}-lift σi\sigma_{i} in (Ai,gi|Ai)(A_{i},g_{i}|_{A_{i}}).

Suppose now that the 11-Lipschitz map Ψ\Psi is not an isometry, which just means that there are two distinct points x,yMx,y\in M so that

(37) d(Ψ(x),Ψ(y))<distg0(x,y).d(\Psi(x),\Psi(y))<\operatorname{dist}_{g_{0}}(x,y).

Choose ε,η\varepsilon,\eta and accordingly ii, so that

(38) 0<εηdistg0(x,y)d(Ψ(x),Ψ(y))100.0<\varepsilon\ll\eta\ll\frac{\operatorname{dist}_{g_{0}}(x,y)-d(\Psi(x),\Psi(y))}{100}.

Let xi,yix_{i},y_{i} be fif_{i}-lifts of x,yx,y. To fix ideas, let us assume without loss of generality that Ψ1(fi(xi))=x\Psi^{-1}(f_{i}(x_{i}))=x and Ψ1(f(yi))=y\Psi^{-1}(f(y_{i}))=y. By the ε\varepsilon-isometry fif_{i} and (38),

(39) distgi(xi,yi)d(Ψ(x),Ψ(y))+ε<distg0(x,y).\operatorname{dist}_{g_{i}}(x_{i},y_{i})\leq d(\Psi(x),\Psi(y))+\varepsilon<\operatorname{dist}_{g_{0}}(x,y).

Let σi(Ai,gi|Ai)\sigma_{i}\subset(A_{i},g_{i}|_{A_{i}}) be a gig_{i}-length minimizing segment which realizes the gi|Aig_{i}|_{A_{i}}-distance between xix_{i} and yiy_{i}. Consider the compact curve Ψ1(fi(σi))\Psi^{-1}(f_{i}(\sigma_{i})) with endpoints x,yx,y, and let us minimize its length among all curves homotopic to Ψ1(fi(σi))\Psi^{-1}(f_{i}(\sigma_{i})) with same endpoints. This yields a g0g_{0}-geodesic segment

σ:[0,lengthg0(σ)](M,g0)\sigma:[0,\operatorname{length}_{g_{0}}(\sigma)]\to(M,g_{0})

parametrized by arclength, with endpoints x,yx,y. Note that, since fif_{i} is a homotopy equivalence, any fif_{i}-lift of σ\sigma with endpoints xi,yix_{i},y_{i} (that can always be ensured) is in fact homotopic (with fixed endpoints) to σi\sigma_{i} inside (M,gi)(M,g_{i}).

By (39), by continuity and uniqueness properties for geodesic loops in hyperbolic manifolds, there are small disjoint open g0g_{0}-geodesic balls Ux,UyU_{x},U_{y} containing respectively x,yx,y and some θ0(0,1)\theta_{0}\in(0,1) with the following property: for any geodesic segment ω:[0,L](M,g0)\omega:[0,L]\to(M,g_{0}) such that ω(0)Ux\omega(0)\in U_{x}, ω(L)Uy\omega(L)\in U_{y}, and

ω[σ]π1(M,UxUy),\omega\in[\sigma]\in\pi_{1}(M,U_{x}\cup U_{y}),

we can find a corresponding fif_{i}-lift ωi\omega_{i} in (M,gi)(M,g_{i}) and a curve ω^i\hat{\omega}_{i} homotopic (with fixed endpoints) to ωi\omega_{i} such that

(40) lengthgi(ω^i)θ0lengthg0(ω).\operatorname{length}_{g_{i}}(\hat{\omega}_{i})\leq\theta_{0}\operatorname{length}_{g_{0}}(\omega).

An important remark is that, since Ψ\Psi is bi-Lipschitz, and since fi:(Ai,gi|Ai)(M,d)f_{i}:(A_{i},g_{i}|_{A_{i}})\to(M,d) is an ε\varepsilon-isometry, the g0g_{0}-length of σ\sigma is uniformly bounded independently of ii. By compactness, we can assume without loss of generality that σ\sigma is fixed and does not depend on ii. For that reason, we will assume that Ux,Uy,θ0U_{x},U_{y},\theta_{0} only depend on M,x,y,σM,x,y,\sigma but not on ii. The notion of fif_{i}-lifts of curves and their properties extend naturally to curves in the universal covers (M~,g0)(\tilde{M},g_{0}) and (M~,gi)(\tilde{M},g_{i}).

Given a Riemannian metric gg on MM and a point 𝐱M\mathbf{x}\in M, let L(g,𝐱)\mathcal{L}_{\leq L}(g,\mathbf{x}) be the collection of homotopy classes of loops with fixed basepoint 𝐱\mathbf{x} which contain at least one loop based at 𝐱M\mathbf{x}\in M of gg-length at most LL. It is well-known that the volume entropy of gg is:

h(g)=limLlog(𝐜𝐚𝐫𝐝L(g,𝐱))Lh(g)=\lim_{L\to\infty}\frac{\log(\mathbf{card}\mathcal{L}_{\leq L}(g,\mathbf{x}))}{L}

where 𝐜𝐚𝐫𝐝\mathbf{card} denotes the cardinality of a set. In particular, it does not depend on the choice of base point 𝐱\mathbf{x}.

Fix a base point 𝐱(M,g0)\mathbf{x}\in(M,g_{0}) and a lift 𝐱~M~\tilde{\mathbf{x}}\in\tilde{M} (here the “lift” belongs to the universal cover, it is not to be confused with the notion of fif_{i}-lift). By uniqueness of geodesic loops in homotopy classes of loops inside hyperbolic manifolds, we identify L(g0,𝐱)\mathcal{L}_{\leq L}(g_{0},\mathbf{x}) with the set of geodesic loops based at 𝐱\mathbf{x} with length at most LL. Classically, the volume entropy of the hyperbolic nn-plane (M~,g0)(\tilde{M},g_{0}) is n1n-1, meaning that

(41) limLlog(𝐜𝐚𝐫𝐝L(g0,𝐱))L=n1.\lim_{L\to\infty}\frac{\log(\mathbf{card}\mathcal{L}_{\leq L}(g_{0},\mathbf{x}))}{L}=n-1.

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 (M,gi)(M,g_{i}) to be strictly larger than n1n-1.

For all ii large, fix an fif_{i}-lift 𝐱i\mathbf{x}_{i} of the basepoint 𝐱\mathbf{x} inside (M,gi)(M,g_{i}), and a lift 𝐱~i(M~,gi)\tilde{\mathbf{x}}_{i}\in(\tilde{M},g_{i}) of 𝐱i\mathbf{x}_{i} in the universal cover. As we saw earlier, we assume without loss of generality that σ\sigma does not depend on ii. By Corollary 3.4, inequality (40) and the properties of fif_{i}-lifts, we deduce that there are some small θ1(0,1)\theta_{1}\in(0,1) and ε,η\varepsilon,\eta (this is where the latter are fixed) depending on M,σ,Ux,UyM,\sigma,U_{x},U_{y} but independent of ii, such that the following holds for all ii large. In the geodesic spheres S~(𝐱~,L)\tilde{S}(\tilde{\mathbf{x}},L) of universal cover (M~,g0)(\tilde{M},g_{0}), for any LL large enough, there is an open subset WLS~(𝐱~,L)W_{L}\subset\tilde{S}(\tilde{\mathbf{x}},L) such that

Area(WL,g0)θ1Area(S~(𝐱~,L),g0),\operatorname{Area}(W_{L},g_{0})\geq\theta_{1}\operatorname{Area}(\tilde{S}(\tilde{\mathbf{x}},L),g_{0}),

and for any qWLq\in W_{L} and any ii large enough, the minimizing geodesic ll from 𝐱~\tilde{\mathbf{x}} to qq admits an fif_{i}-lift joining 𝐱~i\tilde{\mathbf{x}}_{i} to an fif_{i}-lift of qq in (M~,gi)(\tilde{M},g_{i}) which in turn is homotopic (with fixed endpoints) to a curve of gig_{i}-length at most (1θ1)L(1-\theta_{1})L. In colloquial terms, a uniform fraction of points in (M~,g0)(\tilde{M},g_{0}) at g0g_{0}-distance LL from 𝐱~\tilde{\mathbf{x}} admit fif_{i}-lifts in (M~,gi)(\tilde{M},g_{i}) which are at gig_{i}-distance significantly less than LL from 𝐱~i\tilde{\mathbf{x}}_{i}.

By basic hyperbolic geometry (volume of geodesic spheres and balls, etc.) and properties of fif_{i}-lifts, the previous paragraph implies that for ii large enough, for all LL large enough there is a small θ2(0,1)\theta_{2}\in(0,1) depending only on M,σ,Ux,UyM,\sigma,U_{x},U_{y} such that for all i,Li,L large enough,

  • there are distinct points p1,,pK(M~,g0)p_{1},...,p_{K}\in(\tilde{M},g_{0}) which are lifts of 𝐱\mathbf{x} to M~\tilde{M}, and their number satisfies

    (42) Kθ2exp((1+θ2)(n1)L),K\geq\theta_{2}\exp((1+\theta_{2})(n-1)L),
  • there are curves c1,,cK(M~,g0)c_{1},...,c_{K}\subset(\tilde{M},g_{0}) joining 𝐱~\tilde{\mathbf{x}} to p1,,pKp_{1},...,p_{K} respectively, and they admit fif_{i}-lifts in (M~,gi)(\tilde{M},g_{i}), which are respectively homotopic (with fixed endpoints) to curves ci,1,,ci,K(M~,gi)c_{i,1},...,c_{i,K}\subset(\tilde{M},g_{i}) of gig_{i}-lengths at most LL,

  • each of the curves ci,1,,ci,Kc_{i,1},...,c_{i,K} joins 𝐱~i\tilde{\mathbf{x}}_{i} to some other lift of 𝐱i\mathbf{x}_{i} inside the universal cover (M~,gi)(\tilde{M},g_{i}).

We conclude from (41) and (42) that for any ii large enough, for all LL large:

log(𝐜𝐚𝐫𝐝L(gi,𝐱i))log(𝐜𝐚𝐫𝐝(1+θ22)L(g0,𝐱)).\log(\mathbf{card}\mathcal{L}_{\leq L}(g_{i},{\mathbf{x}}_{i}))\geq\log(\mathbf{card}\mathcal{L}_{\leq(1+\frac{\theta_{2}}{2})L}(g_{0},\mathbf{x})).

In particular,

h(gi)\displaystyle h(g_{i}) lim infLlog(𝐜𝐚𝐫𝐝(1+θ22)L(g0,𝐱))L=(1+θ22)h(g0).\displaystyle\geq\liminf_{L\to\infty}\frac{\log(\mathbf{card}\mathcal{L}_{\leq(1+\frac{\theta_{2}}{2})L}(g_{0},\mathbf{x}))}{L}=(1+\frac{\theta_{2}}{2})h(g_{0}).

Since θ2>0\theta_{2}>0 does not depend on ii, the proof is complete.

We are now ready to finish the proof of the intrinsic flat stability theorem.

Theorem 3.7.

Let (M,g0)(M,g_{0}) be a closed oriented hyperbolic manifold of dimension at least 33. Let {gi}i1\{g_{i}\}_{i\geq 1} be a sequence of Riemannian metrics on MM with Vol(M,gi)=Vol(M,g0)\operatorname{Vol}(M,g_{i})=\operatorname{Vol}(M,g_{0}). If

limih(gi)=h(g0)=n1,\lim_{i\to\infty}h(g_{i})=h(g_{0})=n-1,

then there is a sequence of smooth subsets ZiMZ_{i}\subset M such that

limiVol(Zi,gi)=limiArea(Zi,gi)=0\lim_{i\to\infty}\operatorname{Vol}(Z_{i},g_{i})=\lim_{i\to\infty}\operatorname{Area}(\partial Z_{i},g_{i})=0

and (MZi,gi|MZi)(M\setminus Z_{i},g_{i}|_{M\setminus Z_{i}}) converges to (M,g0)(M,g_{0}) 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 AiMA_{i}\subset M such that if

Zi:=MAiZ_{i}:=M\setminus A_{i}

then after renumbering,

  • limiVol(Zi,gi)=limiArea(Zi,gi)=0,\lim_{i\to\infty}\operatorname{Vol}(Z_{i},g_{i})=\lim_{i\to\infty}\operatorname{Area}(\partial Z_{i},g_{i})=0,

  • (MZi,gi|MZi)(M\setminus Z_{i},g_{i}|_{M\setminus Z_{i}}) converges in the intrinsic flat topology to an integral current space C=(X,d,S),C_{\infty}=(X_{\infty},d_{\infty},S_{\infty}),

  • (MZi,gi|MZi)(M\setminus Z_{i},g_{i}|_{M\setminus Z_{i}}) converges to (sptS,d)(\operatorname{spt}S_{\infty},d_{\infty}) in the Gromov-Hausdorff topology,

  • and there is an isometric bijection

    Ψ:(M,g0)(sptS,d)\Psi:(M,g_{0})\to(\operatorname{spt}S_{\infty},d_{\infty})

    such that

    Ψ(1M)=S.\Psi_{\sharp}(\llbracket 1_{M}\rrbracket)=S_{\infty}.

In particular, CC_{\infty} is isomorphic as an integral current space to the hyperbolic manifold (M,g0)(M,g_{0}). Since this integral current space is the only possible subsequential limit, there are ZiMiZ_{i}\subset M_{i} with Vol(Zi,gi)\operatorname{Vol}(Z_{i},g_{i}) and Area(Zi,gi)\operatorname{Area}(\partial Z_{i},g_{i}) converging to 0, and (MZi,gi|MZi)(M\setminus Z_{i},g_{i}|_{M\setminus Z_{i}}) converges to (M,g0)(M,g_{0}) 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 gg on MM, the mass measure of the integral current space (M,distg,1M)(M,\operatorname{dist}_{g},\llbracket 1_{M}\rrbracket) is equal to the usual volume measure dvolg\operatorname{dvol}_{g} on MM. 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 ZZ is a complete metric space, and {Ti}\{T_{i}\} is a sequence of integral currents in ZZ converging weakly to an integral current TT. Moreover, assume that 𝐌(Ti)\mathbf{M}(T_{i}) converges to 𝐌(T)\mathbf{M}(T). Then the mass measure Ti\|T_{i}\| converges weakly to T\|T\| as measures on ZZ.

Remark 3.9.

Sometimes, as in Theorem 0.1, a sequence of nn-manifolds (Mi,gi)(M_{i},g_{i}) converges to a nice limit space XX in a given canonical topology 𝒯\mathscr{T} only after removing negligible subsets ZiZ_{i} from MiM_{i}. 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 Zi\partial Z_{i}. To measure the coarse dimension of a manifold (N,h)(N,h), we propose the following notion of “Euclidean qq-area” 𝒜q(N,h)\mathcal{A}_{q}(N,h):

𝒜q(N,h):=sup{q(π(N));π:(N,h)q is a 1-Lipschitz map}\mathcal{A}_{q}(N,h):=\sup\{\mathcal{H}^{q}(\pi(N));\quad\pi:(N,h)\to\mathbb{R}^{q}\text{ is a $1$-Lipschitz map}\}

where q\mathcal{H}^{q} denotes the standard qq-dimensional Hausdorff measure. Let us declare that (Zi,gi)(\partial Z_{i},g_{i}) has coarse dimension q1q-1 if limi𝒜q(Zi,gi)=0\lim_{i\to\infty}\mathcal{A}_{q}(\partial Z_{i},g_{i})=0. 222A reason why we do not use the notion of Uryson width instead of Euclidean qq-area is that ZiZ_{i} can usually be chosen to have small 0-dimensional Uryson width. As a corollary of Theorem 0.1, for the volume entropy inequality, hyperbolic manifolds of dimension n3n\geq 3 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 RCD((N1),N)RCD*(-(N-1),N) 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 11-Lipschitz maps from integral current spaces into 𝐑n\mathbf{R}^{n}. 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 33-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. dpd_{p} convergence and ϵ\epsilon-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.