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

Spherical volume and spherical Plateau problem

Antoine Song California Institute of Technology
177 Linde Hall, #1200 E. California Blvd., Pasadena, CA 91125
aysong@caltech.edu
Abstract.

Given a closed oriented manifold or more generally a group homology class, we introduce the spherical Plateau problem, which is a variational problem corresponding to a topological invariant called the spherical volume. In principle, its solutions should be realized by minimal surfaces in quotients of spheres. We explain that in many geometrically interesting cases, those solutions are essentially unique. We start with a review of the Ambrosio-Kirchheim theory of metric currents, and the barycenter map method developed by Besson-Courtois-Gallot. Then, we outline the following applications:

  1. (1)

    the intrinsic uniqueness of spherical Plateau solutions for negatively curved, locally symmetric, closed oriented manifolds,

  2. (2)

    the intrinsic uniqueness of spherical Plateau solutions for all 3-dimensional closed oriented manifolds,

  3. (3)

    the construction of higher-dimensional analogues of hyperbolic Dehn fillings.

We also propose some open questions.

Introduction

The classical Plateau problem, a fundamental question in Differential Geometry, concerns the existence and regularity of surfaces of least area spanning a given boundary contour inside the 33-dimensional Euclidean space. The term “Plateau problem” has been extended to encompass any situation where the objective is to construct and study “minimal surfaces”, which are minimizers of a volume or area functional subject to topological or geometric constraints. As a concrete example, consider a bounded Riemannian manifold (N,g)(N,g) of finite or infinite dimension, and fix an nn-dimensional integer homology class hHn(N;)h\in H_{n}(N;\mathbb{Z}) where nn is a positive integer. Roughly speaking, the “volume” or “area” of this homology class is defined as follows:

Area(h):=inf{Area(C);C is a cycle in N representing h}.\operatorname{Area}(h):=\inf\{\operatorname{Area}(C);\quad\text{$C$ is a cycle in $N$ representing $h$}\}.

Here cycles can be intuitively understood as generalized nn-submanifolds of NN. Let CiC_{i} be a sequence of cycles in NN representing hh, which is minimizing in the sense that

limiArea(Ci)=Area(h).\lim_{i\to\infty}\operatorname{Area}(C_{i})=\operatorname{Area}(h).

Thanks to adequate compactness results, a subsequence CijC_{i_{j}} converges to some limit space CC_{\infty} called Plateau solution. The properties of this limit space, such as regularity, uniqueness, and geometric structure, are the main focus of the Plateau problem. In some way, this space CC_{\infty} can be considered as an “optimal geometric representative” of the homology class hh of (N,g)(N,g). Uniqueness of Plateau solutions holds only in exceptional situations, and this paper is about a natural Plateau problem in infinite dimension, for which interesting uniqueness results hold or are conjectured to be true.

A central question in the study of Plateau problems is to define the concepts of “cycle”, “area”, “volume” and “convergence”. We will rely on the framework of integral currents in Geometric Measure Theory, as it provides the most far-reaching existence and regularity results so far, though there are other possible choices. In this context, the Plateau problem has been thoroughly studied for finite dimensional manifolds NN: in particular if the Riemannian manifold (N,g)(N,g) is finite dimensional and closed, any kk-dimensional integer homology class hh of NN admits an area-minimizing integral current representative CC_{\infty} in NN with area equal to Area(h)\operatorname{Area}(h), which is smooth outside of a codimension two subset (this object is a “generalized kk-dimensional minimal surface”). This major result follows from the successive works of De Giorgi, Federer-Fleming, Allard, Almgren, and many others. See [FF60, Theorem 9.6] for the existence statement and the surveys [Amb16, DL16] for the partial regularity problem.

However when the underlying manifold NN is infinite-dimensional and hence not locally compact, the situation becomes more challenging. Efforts have been made to extend the theory of currents to arbitrary complete metric spaces, which started with Ambrosio-Kirchheim’s article [AK00a]. This theory has since been applied, extended, and revisited by many researchers, including Lang [Lan11], Ambrosio-Schmidt [AS13], Wenger [Wen07, Wen11, Wen14], Sormani-Wenger [SW11], Ambrosio-De Lellis-Schmidt [ADLS18] etc. It is this extended theory which allows us to define and construct Plateau solutions in general. More concretely, if (N,g)(N,g) is an infinite-dimensional complete Riemannian manifold with finite diameter and hHn(N;)h\in H_{n}(N;\mathbb{Z}), a cycle in (N,g)(N,g) is by definition a boundaryless integral current with compact support in (N,g)(N,g) in the sense of [AK00a]. Its area is given by the notion of mass of an integral current [AK00a]. Given a minimizing sequence of cycles CiC_{i} representing hh, by Wenger’s compactness theorem [Wen11], CiC_{i} subsequentially converges in the intrinsic flat topology to an integral current space CC_{\infty} [SW11]. An integral current space is roughly speaking determined by an underlying metric space (X,d)(X,d) and an integral current SS in the completion of (X,d)(X,d). Any such CC_{\infty} is called a Plateau solution for hh.

Inspired by earlier works of Besson-Courtois-Gallot [BCG91, BCG95, BCG96], we consider the spherical Plateau problem, a natural infinite-dimensional Plateau problem at the interface of geometric measure theory, geometric group theory and topology. The ambient manifold (N,g)(N,g) is of the following form:

(N,g)=(S/λΓ(Γ),𝐠Hil)(N,g)=({S^{\infty}}/\lambda_{\Gamma}(\Gamma),\mathbf{g}_{\mathrm{Hil}})

where Γ\Gamma is a countable group, S{S^{\infty}} is the unit sphere in 2(Γ)\ell^{2}(\Gamma), Γ\Gamma acts on S{S^{\infty}} by the left regular representation λΓ:ΓEnd(2(Γ))\lambda_{\Gamma}:\Gamma\to\operatorname{End}(\ell^{2}(\Gamma)) and S/λΓ(Γ){S^{\infty}}/\lambda_{\Gamma}(\Gamma) is endowed with the quotient of the round metric 𝐠Hil\mathbf{g}_{\mathrm{Hil}}. Given hHn(S/λΓ(Γ);)h\in H_{n}({S^{\infty}}/\lambda_{\Gamma}(\Gamma);\mathbb{Z}), the invariant Area(h)\operatorname{Area}(h) is called the spherical volume of hh and denoted by

SphereVol(h):=inf{Area(C);C is a cycle in S/λΓ(Γ) representing h}.\operatorname{SphereVol}(h):=\inf\{\operatorname{Area}(C);\quad\text{$C$ is a cycle in ${S^{\infty}}/\lambda_{\Gamma}(\Gamma)$ representing $h$}\}.

Versions of that invariant were first defined by Besson-Courtois-Gallot [BCG91, BCG95, BCG96]. A more detailed definition is given in Section 3. The integral current spaces CC_{\infty} corresponding to hh, obtained by the recipe described in the previous paragraph, are called spherical Plateau solutions for hh.

Aim. The purpose of this note is to introduce the spherical Plateau problem, and show that for many special choices of Γ\Gamma and hHn(S/λΓ(Γ);)h\in H_{n}({S^{\infty}}/\lambda_{\Gamma}(\Gamma);\mathbb{Z}), the spherical Plateau solutions are in a sense unique and geometrically meaningful. Throughout the text, we propose several questions related to the spherical volume and the spherical Plateau problem.

The style of this article is rather informal, nevertheless we try to provide a rather comprehensive review of the main technical tools.

We will outline the proofs of the three results stated below111Most of the content in this survey is based on an earlier unpublished preprint [Son22]. Theorem 0.1 is proved in details in [Son23]. See also Cosmin Manea’s master’s thesis [Man23]., which we view as a proof-of-concept justifying a general study of the spherical Plateau problem.

Locally symmetric manifolds of negative curvature. The first result pertains to locally symmetric spaces of rank one. Let (M,g0)(M,g_{0}) be a closed oriented locally symmetric nn-manifold with negative curvature, and let Γ\Gamma be its fundamental group. Let h(g0)h(g_{0}) be the volume entropy of (M,g0)(M,g_{0}), whose definition is recalled in (31). The quotient space S/λΓ(Γ){S^{\infty}}/\lambda_{\Gamma}(\Gamma) is a classifying space for Γ\Gamma and the fundamental class [M]Hn(M;)[M]\in H_{n}(M;\mathbb{Z}) determines a unique homology class hMHn(S/λΓ(Γ);)=Hn(M;)h_{M}\in H_{n}({S^{\infty}}/\lambda_{\Gamma}(\Gamma);\mathbb{Z})=H_{n}(M;\mathbb{Z}). A key step in the proof of the celebrated volume entropy inequality of Besson-Courtois-Gallot [BCG95] was the computation of SphereVol(hM)\operatorname{SphereVol}(h_{M}) with the barycenter map method, see Section 4. The barycenter map will be a central tool in this paper too.

A closed oriented Riemannian nn-manifold (W,gW)(W,g_{W}) admits a natural integral current structure 1W\llbracket 1_{W}\rrbracket induced by its fundamental class [W]Hn(W;)[W]\in H_{n}(W;\mathbb{Z}). A spherical Plateau solution CC_{\infty} for hMh_{M} is called “intrinsically isomorphic” to (W,gW)(W,g_{W}) if the underlying metric space of CC_{\infty} is intrinsically isometric to (W,gW)(W,g_{W}) via a map sending the current structure of CC_{\infty} to 1W\llbracket 1_{W}\rrbracket, see Definition 3.5.

Theorem 0.1.

If (M,g0)(M,g_{0}) is a closed oriented locally symmetric manifold of dimension n3n\geq 3, with negative curvature between 4-4 and 1-1, then any spherical Plateau solution for hMh_{M} is intrinsically isomorphic to (M,h(g0)24ng0)(M,\frac{h(g_{0})^{2}}{4n}g_{0}).

This intrinsic uniqueness result leads to the formulation of a new kind of rigidity property for the regular representation of π1(M)\pi_{1}(M), see Corollary 4.3. The proof of Theorem 0.1 can also be applied to show that the entropy inequality of Besson-Courtois-Gallot [BCG95] is stable [Son23]. Similar questions for higher rank locally symmetric manifolds (even the computation of the spherical volume) are still wide open, because the barycenter map method does not work so well there [CF02]. Motivated by Theorem 0.1, it is suggested in [CWN23] to inspect the spherical Plateau problem in the context of Cannon’s conjecture.

Closed oriented 33-manifolds. Our next result gives another large pool of examples where the spherical Plateau solutions are almost explicit. Let MM be a closed oriented 33-manifold and Γ\Gamma its fundamental group. The fundamental class of MM induces a homology class hMH3(S/λΓ(Γ);)h_{M}\in H_{3}({S^{\infty}}/\lambda_{\Gamma}(\Gamma);\mathbb{Z}). By the Geometrization theorem [KL08, MT14], MM can be uniquely written as a connected sum M=M1##MkM=M_{1}\#...\#M_{k} where each MjM_{j} is a closed oriented prime 33-manifold which is canonically divided into two pieces along some tori: Mj=Mj,hypMj,SeifM_{j}=M_{j,\mathrm{hyp}}\cup M_{j,\mathrm{Seif}} where Mj,hypM_{j,\mathrm{hyp}} carries a finite-volume complete hyperbolic metric, and Mj,SeifM_{j,\mathrm{Seif}} is a union of Seifert manifolds. The disjoint union of the hyperbolic pieces Mj,hypM_{j,\mathrm{hyp}} endowed with their hyperbolic metrics is denoted (Mhyp,ghyp)(M_{\mathrm{hyp}},g_{\mathrm{hyp}}).

Theorem 0.2.

If MM is closed oriented 33-manifold with hyperbolic part denoted by (Mhyp,ghyp)(M_{\mathrm{hyp}},g_{\mathrm{hyp}}), then any spherical Plateau solution for hMh_{M} is intrinsically isomorphic to (Mhyp,13ghyp)(M_{\mathrm{hyp}},\frac{1}{3}g_{\mathrm{hyp}}).

Hence the hyperbolic part (Mhyp,ghyp)(M_{\mathrm{hyp}},g_{\mathrm{hyp}}), which is canonically determined by MM, emerges as the solution of the spherical Plateau problem. The fundamental nature of the hyperbolic part (Mhyp,ghyp)(M_{\mathrm{hyp}},g_{\mathrm{hyp}}) had previously appeared in other contexts including the Ricci flow, the Yamabe invariant, the simplicial volume, the volume entropy, etc. For instance, the normalized Ricci flow starting at any Riemannian metric on MM converges as time goes to infinity to (Mhyp,ghyp)(M_{\mathrm{hyp}},g_{\mathrm{hyp}}) in a multi-pointed Gromov-Hausdorff sense [KL08].

Plateau Dehn fillings. The two previous theorems show that, to some extent, spherical Plateau solutions form a class of spaces which naturally generalizes locally symmetric manifolds of negative curvature. Our last result explores that interpretation further in the context of Dehn fillings. Recall that in dimension 33, there is a fundamental feature of hyperbolic geometry discovered by Thurston, called hyperbolic Dehn fillings [Thu97]: in its simplest version, it states that given any finite volume non-compact hyperbolic 33-manifold MM, there is a sequence of finite volume hyperbolic manifolds MiM_{i} with volume strictly less than that of MM, and converging geometrically to MM. In higher dimensions, due to the finiteness theorem of H.C. Wang [Wan72] for locally symmetric negatively curved manifolds, such a phenomenon is impossible in the smooth setting. Nevertheless, as we will see below, by enlarging the set of locally symmetric negatively curved manifolds to the set of spherical Plateau solutions, we do have such an accumulation phenomenon in all dimensions higher than 22. We start with the higher dimensional CAT(0)\mathrm{CAT}(0) Dehn fillings constructed by Fujiwara-Manning in [FM10]. These fillings are denoted by M(T1,,Tm)M(T_{1},...,T_{m}) and are obtained by closing the cusps of a finite volume non-compact hyperbolic nn-manifold (M,ghyp)(M,g_{\mathrm{hyp}}) with toral cusps. The TiT_{i} denote certain subtori in the mm cusps of MM, which are assumed to have injectivity radius greater than π\pi. Each M(T1,,Tm)M(T_{1},...,T_{m}) determines a unique homology class hM(T1,,Tm)h_{M(T_{1},...,T_{m})} in the corresponding spherical quotient S/λΓ(Γ){S^{\infty}}/\lambda_{\Gamma}(\Gamma) where Γ:=π1(M(T1,,Tm))\Gamma:=\pi_{1}(M(T_{1},...,T_{m})). The behavior of Plateau Dehn fillings, namely the spherical Plateau solutions for hM(T1,,Tm)h_{M(T_{1},...,T_{m})}, is completely analogous to the 33-dimensional case of hyperbolic Dehn fillings [Gro81]:

Theorem 0.3.

We have

SphereVol(hM(T1,,Tm))<Vol(M,(n1)24nghyp).\operatorname{SphereVol}(h_{M(T_{1},...,T_{m})})<\operatorname{Vol}(M,\frac{(n-1)^{2}}{4n}g_{\mathrm{hyp}}).

Moreover for any ϵ>0\epsilon>0, if the injectivity radii of T1,,TmT_{1},...,T_{m} are sufficiently large, then

SphereVol(hM(T1,,Tm))>Vol(M,(n1)24nghyp)ϵ\operatorname{SphereVol}(h_{M(T_{1},...,T_{m})})>\operatorname{Vol}(M,\frac{(n-1)^{2}}{4n}g_{\mathrm{hyp}})-\epsilon

and any spherical Plateau solution for hM(T1,,Tm)h_{M(T_{1},...,T_{m})} is ϵ\epsilon-close, in the intrinsic flat topology, to an integral current space intrinsically isomorphic to (M,(n1)24nghyp)(M,\frac{(n-1)^{2}}{4n}g_{\mathrm{hyp}}).

Fujiwara-Manning previously proved that the fillings M(T1,,Tm)M(T_{1},...,T_{m}) have simplicial volumes satisfying

M(T1,,Tm)M\|M(T_{1},...,T_{m})\|\leq\|M\|

where M\|M\| is the simplicial volume of MM, and conjectured that M(T1,,Tm)\|M(T_{1},...,T_{m})\| converges to M\|M\| as the injectivity radius of the tori TiT_{i} goes to infinity, but never attains the limit [FM11, Conjecture 1.8, Question 1.9]. The previous theorem settles the spherical volume version of this conjecture. For the construction of Einstein manifolds analogous to Dehn fillings, see [And06, Bam12].

Outline of the paper. In Section 1, we give an overview of the theory of integral currents in metric spaces developed by Ambrosio-Kirchheim and others. Results are stated without proofs.

In Section 2, we explain in details the barycenter map and reproduce the proofs of some important properties. For clarity, we focus only on the hyperbolic case.

In Section 3, the spherical Plateau problem is defined, and basic properties of spherical cycles are discussed.

In Section 4, we sketch Besson-Courtois-Gallot’s computation of the spherical volume for hyperbolic manifolds. We briefly explain its close relation to the volume entropy inequality. We then outline the proof of our first uniqueness result, Theorem 4.2.

In Section 5, we indicate how to apply the Geometrization theorem for 33-manifolds, and extend the arguments of Theorem 4.2 to prove the uniqueness statement for 33-manifolds, Theorem 5.2.

In Section 6, we describe the idea behind Plateau Dehn fillings and Theorem 6.2, which can be viewed as an asymptotic uniqueness result.

Acknowledgments

I am grateful to Gérard Besson, Gilles Courtois, John Lott, Ian Agol, Jason Manning, Camillo De Lellis, Xin Zhou, Hyun Chul Jang, Luca Spolaor, Tamunonye Cheetham-West, Alexander Nolte, Luca Di Cerbo, Richard Bamler, Song Sun, Stéphane Sabourau, Alexander Nabutovsky, Ben Lowe, Shi Wang, Cosmin Manea and Zhenhua Liu for many insightful and stimulating discussions during the writing of this article.

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. Preliminaries on metric integral currents

The classical notion of integral currents in finite dimensional manifolds [Fed69] exhibits simultaneously several properties explaining the success of the theory: they are mild generalizations of submanifolds, they satisfy strong compactness results and area-minimizers are smooth submanifolds outside of a small singular set [Alm00, DLS14]. Building on earlier ideas of De Giorgi and Gromov, Ambrosio-Kirchheim [AK00a, AK00b] initiated an extension of the theory to complete metric spaces, including infinite-dimensional Riemannian manifolds. Further developments led to Wenger’s compactness result [Wen11] which will be essential to define spherical Plateau solutions (Subsection 3.1). In this section, we review the definitions and results developed in [AK00a, AK00b, Wen07, Wen11, SW11].

1.1. Basic definitions

Let (E,d)(E,d) be a complete metric space222The metric dd is allowed to take \infty as value.. Let n0n\geq 0, and let 𝒟n(E)\mathcal{D}^{n}(E) be the set of (n+1)(n+1)-tuples (f,π1,,πn)(f,\pi_{1},...,\pi_{n}) of Lipschitz functions on EE with ff bounded. As a suggestive reference to the finite dimensional theory of currents, (f,π1,,πn)(f,\pi_{1},...,\pi_{n}) is also usually denoted by fdπ1,dπnfd\pi_{1}\wedge...,\wedge d\pi_{n}. Metric currents in the sense of Ambrosio-Kirchheim are a flexible generalization of oriented submanifolds:

Definition 1.1.

[AK00a] An nn-dimensional metric current in (E,d)(E,d) is a multi-linear functional on 𝒟n(E)\mathcal{D}^{n}(E) such that

  1. (1)

    If πij\pi_{i}^{j} converges pointwise to πi\pi_{i} as jj\to\infty, and if supi,jLip(πij)<\sup_{i,j}\operatorname{{Lip}}(\pi_{i}^{j})<\infty, then

    limjT(fdπ1j,dπnj)=T(fdπ1,dπn).\lim_{j\to\infty}T(fd\pi_{1}^{j}\wedge...,\wedge d\pi_{n}^{j})=T(fd\pi_{1}\wedge...,\wedge d\pi_{n}).
  2. (2)

    If {xE;f(x)0}\{x\in E;f(x)\neq 0\} is contained in the union i=1nBi\bigcup_{i=1}^{n}B_{i} of Borel sets BiB_{i} and if πi\pi_{i} is constant on BiB_{i} then

    T(fdπ1,dπn)=0.T(fd\pi_{1}\wedge...,\wedge d\pi_{n})=0.
  3. (3)

    There exists a finite Borel measure μ\mu on EE such that

    |T(fdπ1,dπn)|i=1kLip(πi)E|f|𝑑μ|T(fd\pi_{1}\wedge...,\wedge d\pi_{n})|\leq\prod_{i=1}^{k}\operatorname{{Lip}}(\pi_{i})\int_{E}|f|d\mu

    for all fdπ1,dπn𝒟n(E)fd\pi_{1}\wedge...,\wedge d\pi_{n}\in\mathcal{D}^{n}(E).

The minimal Borel measure μ\mu satisfying the above inequality is called the mass of TT and denoted by T\|T\|. The total mass of TT is defined as

𝐌(T):=T(E),\mathbf{M}(T):=\|T\|(E),

and should be thought of as the “nn-dimensional area” of TT. Currents in the sense of Ambrosio-Kirchheim [AK00a] have finite mass by definition (and we will only consider such currents here), but there is a variant of this theory due to Lang [Lan11] which avoids the finite mass condition. The support spt(T)\operatorname{spt}(T) of TT is the support of the measure T\|T\| in the usual sense. The canonical set of TT, called set(T)\operatorname{set}(T), is the collection of points in EE with a positive lower density:

set(T):={pE;limr0+T(B(p,r))rn>0}.\operatorname{set}(T):=\{p\in E;\lim_{r\to 0^{+}}\|T\|(B(p,r))r^{-n}>0\}.

In general,

set(T)spt(T)\operatorname{set}(T)\subset\operatorname{spt}(T)

and the inclusion is usually strict. The boundary T\partial T is defined by

T(fdπ1,dπn1):=T(1dfdπ1,dπn)\partial T(fd\pi_{1}\wedge...,\wedge d\pi_{n-1}):=T(1df\wedge d\pi_{1}\wedge...,\wedge d\pi_{n})

for all fdπ1,dπn1𝒟n1(E)fd\pi_{1}\wedge...,\wedge d\pi_{n-1}\in\mathcal{D}^{n-1}(E). The push-forward of TT by a Lipschitz map ψ\psi from EE to another complete metric space EE^{\prime} is given by

ψT(fdπ1,dπn):=T(fψd(π1ψ)d(πnψ))\psi_{\sharp}T(fd\pi_{1}\wedge...,\wedge d\pi_{n}):=T(f\circ\psi d(\pi_{1}\circ\psi)\wedge...\wedge d(\pi_{n}\circ\psi))

for all fdπ1,dπn𝒟n(E)fd\pi_{1}\wedge...,\wedge d\pi_{n}\in\mathcal{D}^{n}(E^{\prime}). There is also a notion of restriction of TT to a Borel subset AEA\subset E:

(TA)(fdπ1,dπn):=T(fχAdπ1,dπn)(T\llcorner A)(fd\pi_{1}\wedge...,\wedge d\pi_{n}):=T(f\chi_{A}d\pi_{1}\wedge...,\wedge d\pi_{n})

where χA\chi_{A} is the characteristic function of AA (the above is well-defined by an extension of the functional TT).

1.2. Rectifiable sets and integral currents

We are mainly interested in integral currents, which roughly speaking are currents TT such that both TT and T\partial T are the push-forward of a countable union of elementary currents by Lipschitz maps. An elementary current is obtained as follows: consider θL1(A;)\theta\in L^{1}(A;\mathbb{N}) where AnA\subset\mathbb{R}^{n}, then define the following current in n\mathbb{R}^{n}: for all fdπ1,dπn𝒟n(n)fd\pi_{1}\wedge...,\wedge d\pi_{n}\in\mathcal{D}^{n}(\mathbb{R}^{n}),

θ(fdπ1,dπn):=Aθfdπ1,dπn.\llbracket\theta\rrbracket(fd\pi_{1}\wedge...,\wedge d\pi_{n}):=\int_{A}\theta fd\pi_{1}\wedge...,\wedge d\pi_{n}.

Integer rectifiable currents and integral currents enjoy the following characterizations [AK00a, Theorem 4.5, Theorem 8.6], which we will take as definitions:

Definition 1.2.

[AK00a] A current TT in EE is an nn-dimensional integer rectifiable current if and only if there are Lipschitz maps φi:AiE\varphi_{i}:A_{i}\to E where AinA_{i}\subset\mathbb{R}^{n} are precompact Borel measurable and have disjoint images by φi\varphi_{i}, and there are θiL1(Ai;)\theta_{i}\in L^{1}(A_{i};\mathbb{N}) such that

T=i=1(φi)θi,andT=i=1(φi)θi.T=\sum_{i=1}^{\infty}(\varphi_{i})_{\sharp}\llbracket\theta_{i}\rrbracket,\quad\text{and}\quad\|T\|=\sum_{i=1}^{\infty}\|(\varphi_{i})_{\sharp}\llbracket\theta_{i}\rrbracket\|.

The pair ({φi:AiE},{θi})(\{\varphi_{i}:A_{i}\to E\},\{\theta_{i}\}) is called a parametrization.

The current TT is an integral current if and only if both TT and T\partial T are integer rectifiable currents.

For instance, if (M,g)(M,g) is a complete oriented Riemannian manifold with compact boundary and finite volume, then it carries a natural nn-dimensional integral current usually denoted by 1M\llbracket 1_{M}\rrbracket, induced by “integration on MM”.

Recall that a Borel set SES\subset E is countable n\mathcal{H}^{n}-rectifiable if there is a sequence of Lipschitz functions φi:AiE\varphi_{i}:A_{i}\to E where AinA_{i}\subset\mathbb{R}^{n} is Borel, such that

(1) n(Siφi(Ai))=0\mathcal{H}^{n}(S\setminus\bigcup_{i}\varphi_{i}(A_{i}))=0

where n\mathcal{H}^{n} denotes the nn-dimensional Hausdorff measure. It is proved in [AK00a, Theorem 4.6] that if TT is an nn-dimensional integral current with a parametrization ({φi},{θi})(\{\varphi_{i}\},\{\theta_{i}\}), then

n(set(T)i=1φi(Ai)i=1φi(Ai)set(T))=0.\mathcal{H}^{n}\big{(}\operatorname{set}(T)\setminus\bigcup_{i=1}^{\infty}\varphi_{i}(A_{i})\cup\bigcup_{i=1}^{\infty}\varphi_{i}(A_{i})\setminus\operatorname{set}(T)\big{)}=0.

The canonical set set(T)\operatorname{set}(T) is in particular a countably n\mathcal{H}^{n}-rectifiable set in EE (contrarily to the support spt(T)\operatorname{spt}(T), in general). The functions θi\theta_{i} determine a Borel function called the multiplicity function θT:E\theta_{T}:E\to\mathbb{N}, which is well-defined n\mathcal{H}^{n}-almost everywhere. In the special case where EE is an infinite Riemannian dimensional manifold (i.e. locally modelled on a Hilbert space), it is shown in [AK00a, Theorem 9.5] that

T=θTnset(T).\|T\|=\theta_{T}\mathcal{H}^{n}\llcorner\operatorname{set}(T).

That intuitive formula needs a correction factor in the case of general Banach spaces.

1.3. Weak and flat topology

There are two fundamental notions of convergence for integral currents in a metric space: the weak and flat convergences. A sequence {Tm}\{T_{m}\} of nn-dimensional integral currents in EE is said to converge weakly to some current TT if for all fdπ1,dπn𝒟n(E)fd\pi_{1}\wedge...,\wedge d\pi_{n}\in\mathcal{D}^{n}(E),

limmTm(fdπ1,dπn)=T(fdπ1,dπn).\lim_{m\to\infty}T_{m}(fd\pi_{1}\wedge...,\wedge d\pi_{n})=T(fd\pi_{1}\wedge...,\wedge d\pi_{n}).

In that case, an important result states that such a limit TT is also an integral current [AK00a, Theorem 8.5]. Besides, a fundamental property of the mass which makes it so useful in Plateau problems is that it is lower semicontinuous with respect to weak converge, see paragraph below [AK00a, Definition 3.6]: if TmT_{m} converges weakly to TT then

𝐌(T)lim infm𝐌(Tm).\mathbf{M}(T)\leq\liminf_{m\to\infty}\mathbf{M}(T_{m}).

The sequence {Tm}\{T_{m}\} is said to converge to TT in the flat topology if there are sequences {Um},{Vm}\{U_{m}\},\{V_{m}\} of integral currents such that

TmT=Um+Vm,limm𝐌(Um)=limm𝐌(Vm)=0.T_{m}-T=U_{m}+\partial V_{m},\quad\lim_{m\to\infty}\mathbf{M}(U_{m})=\lim_{m\to\infty}\mathbf{M}(V_{m})=0.

Convergence in the flat topology implies convergence in the weak topology. A partial converse is proved in [Wen07].

1.4. The area and coarea formulas

Next we recall some versions of two particularly useful tools, the area and coarea formulas for countably n\mathcal{H}^{n}-rectifiable sets. In this paragraph we assume for simplicity that (E,d)(E,d) is a separable complete infinite-dimensional Riemannian manifold, embedded isometrically inside an \ell^{\infty} space YY by a Kuratowski embedding, since only that case is needed here. Note that an \ell^{\infty} space is a ww^{*}-separable dual space in the sense of [AK00a]. Given a Lipschitz map ψ:AE\psi:A\to E where AnA\subset\mathbb{R}^{n} is Borel, for almost every yAy\in A, ψ\psi is ww^{*}-differentiable at yy with a ww^{*}-differential called wdyψwd_{y}\psi in the sense of [AK00b][AK00a, Section 9]. The latter is a linear map from TynT_{y}\mathbb{R}^{n} to YY. For almost every yAy\in A, wdyψwd_{y}\psi is of full rank; in that case the image Q:=wdyψ(Tyn)Q:=wd_{y}\psi(T_{y}\mathbb{R}^{n}) is called an approximate tangent nn-plane at p:=ψ(y)p:=\psi(y). Such QQ is a linear nn-plane inside YY, and in our case it is also a tangent nn-plane of the manifold EE. More generally, let SES\subset E be a countably n\mathcal{H}^{n}-rectifiable set, with {φi:AiE}\{\varphi_{i}:A_{i}\to E\} as in (1). At n\mathcal{H}^{n}-almost every pSp\in S, there are ii and yAiy\in A_{i} such that φi(y)=p\varphi_{i}(y)=p and an approximate tangent nn-plane QQ at pp exists in the sense above.

Let SS be a countably n\mathcal{H}^{n}-rectifiable set in EE. Given a Lispchitz map g:SEg:S\to E^{\prime} where EE^{\prime} is another separable complete infinite-dimensional Riemannian manifold embedded isometrically in an \ell^{\infty} space YY^{\prime}, there is a well-defined nonnegative number for n\mathcal{H}^{n}-almost every zSz\in S, called the Jacobian of gg and denoted by 𝐉n(dSgz)\mathbf{J}_{n}(d^{S}g_{z}), such that the following area formula [AK00b, Theorem 8.2] holds for any Borel function θ:S[0,]\theta:S\to[0,\infty]:

(2) Sθ(p)𝐉n(dSgp)𝑑n(p)=EpSg1(z)θ(p)dn(z).\int_{S}\theta(p)\mathbf{J}_{n}(d^{S}g_{p})d\mathcal{H}^{n}(p)=\int_{E^{\prime}}\sum_{p\in S\cap g^{-1}(z)}\theta(p)d\mathcal{H}^{n}(z).

More concretely, suppose that for {φi:AiE}\{\varphi_{i}:A_{i}\to E\} as (1), for ii and yAiy\in A_{i}, p:=φi(y)Sp:=\varphi_{i}(y)\in S, and suppose that φi\varphi_{i} is ww^{*}-differentiable at yy with a ww^{*}-differential of full rank, and the composition gφi:nEg\circ\varphi_{i}:\mathbb{R}^{n}\to E^{\prime} is ww^{*}-differentiable at yy (all of this holds for almost every yAiy\in A_{i}). Let

Q:=wdyφi(Tyn)Q:=wd_{y}\varphi_{i}(T_{y}\mathbb{R}^{n})

be the tangent nn-plane at pp. Then gg is tangentially differentiable at pp along QQ with tangential differential dSgp:QTg(p)Ed^{S}g_{p}:Q\to T_{g(p)}E^{\prime} satisfying:

dSgp=wdy(gφi)(wdyφi)1.d^{S}g_{p}=wd_{y}(g\circ\varphi_{i})\circ(wd_{y}\varphi_{i})^{-1}.

In our case where E,EE,E^{\prime} are infinite-dimensional Riemannian manifolds, QQ and Tg(p)ET_{g(p)}E^{\prime} are Hilbert linear spaces. The Jacobian 𝐉n(dgp)\mathbf{J}_{n}(dg_{p}) is then equal to the absolute value of the usual Jacobian determinant of the linear map dSgpd^{S}g_{p}. In particular when gg is λ\lambda-Lipschitz then

𝐉n(dgz)λn\mathbf{J}_{n}(dg_{z})\leq\lambda^{n}

as expected. Given a tangent nn-plane QQ of SS at pp as above, we often use the following more classical notations:

dg|Q:=dSgp,dg\big{|}_{Q}:=d^{S}g_{p},
|Jacg|Q|:=𝐉n(dSgp).|\operatorname{Jac}g\big{|}_{Q}|:=\mathbf{J}_{n}(d^{S}g_{p}).

Let SS be a countably n\mathcal{H}^{n}-rectifiable set in EE. Consider a Lipschitz function π:Sk\pi:S\to\mathbb{R}^{k} where knk\leq n. At n\mathcal{H}^{n}-almost every pSp\in S, there is a tangent nn-plane QQ along which the tangential differential dSπpd^{S}\pi_{p} exists. At such pp, there is a nonnegative number called coarea factor and denoted by 𝐂k(dSπp)\mathbf{C}_{k}(d^{S}\pi_{p}) such that the following coarea formula [AK00b, Theorem 9.4] holds for any Borel function θ:S[0,]\theta:S\to[0,\infty]:

(3) Sθ(p)𝐂k(dSπp)𝑑n(p)=k(π1(z)θ(p)𝑑nk(p))𝑑k(z).\int_{S}\theta(p)\mathbf{C}_{k}(d^{S}\pi_{p})d\mathcal{H}^{n}(p)=\int_{\mathbb{R}^{k}}\big{(}\int_{\pi^{-1}(z)}\theta(p)d\mathcal{H}^{n-k}(p)\big{)}d\mathcal{H}^{k}(z).

Besides, for k\mathcal{H}^{k}-almost every zkz\in\mathbb{R}^{k}, the set π1(z)S\pi^{-1}(z)\cap S is countably nk\mathcal{H}^{n-k}-rectifiable. The precise definition of the coarea factor 𝐂k(dπz)\mathbf{C}_{k}(d\pi_{z}) is a bit more involved than the Jacobian [AK00b, Definition 9.1], nevertheless it is similar to what appears in the smooth finite dimensional case and when π\pi is λ\lambda-Lipschitz then

𝐂k(dπz)λk\mathbf{C}_{k}(d\pi_{z})\leq\lambda^{k}

as expected.

1.5. The slicing theorem for integral currents

We move on to the slicing theorem for integral currents, which enables to construct lower dimensional integral currents out of a given integral current and a Lipschitz map. Again, we still assume for simplicity that EE is an infinite-dimensional Riemannian manifold (locally modelled on a Hilbert space). Let SS be a countably n\mathcal{H}^{n}-rectifiable set in EE. Given fdπ1dπn𝒟n(E)fd\pi_{1}\wedge...\wedge d\pi_{n}\in\mathcal{D}^{n}(E), then at n\mathcal{H}^{n}-almost every pSp\in S, an approximate tangent nn-plane QQ exists, the tangential differentials dS(πj)pd^{S}(\pi_{j})_{p} along QQ exist and the nn-covector fdSπ1dSπnfd^{S}\pi_{1}\wedge...\wedge d^{S}\pi_{n} is well-defined. There is a notion of orientations on SS [AK00a, Section 9]. Roughly speaking, an orientation τ\tau on SS endows each approximate tangent nn-plane QQ of SS with an nn-vector 𝐞1𝐞n\mathbf{e}_{1}\wedge...\wedge\mathbf{e}_{n} where (𝐞1,,𝐞n)(\mathbf{e}_{1},...,\mathbf{e}_{n}) is an orthonormal basis of QQ (recall that here EE is a manifold).

Let TT be an nn-dimensional integral current in EE, with multiplicity function θT\theta_{T}. There is an intrinsic description of TT, based on the notion of orientation [AK00a, Theorem 9.1]. In fact there is an orientation τT\tau_{T} on set(T)\operatorname{set}(T) such that for all fdπ1dπn𝒟n(E)fd\pi_{1}\wedge...\wedge d\pi_{n}\in\mathcal{D}^{n}(E),

T(fdπ1dπn)=set(T)f(p)θT(p)dset(T)π1dset(T)πn,τT𝑑n(p).T(fd\pi_{1}\wedge...\wedge d\pi_{n})=\int_{\operatorname{set}(T)}f(p)\theta_{T}(p)\langle d^{\operatorname{set}(T)}\pi_{1}\wedge...\wedge d^{\operatorname{set}(T)}\pi_{n},\tau_{T}\rangle d\mathcal{H}^{n}(p).

Conversely if SS is a countably n\mathcal{H}^{n}-rectifiable set in EE, if we have θ:E\theta:E\to\mathbb{N} such that Sθ𝑑n<\int_{S}\theta d\mathcal{H}^{n}<\infty, and an orientation τ\tau on SS then the current

S,θ,τ\llbracket S,\theta,\tau\rrbracket

defined by the analogue of the formula above is an nn-dimensional integral current.

Now consider a Lipschitz map π:spt(T)k\pi:\operatorname{spt}(T)\to\mathbb{R}^{k} where knk\leq n. Then for each xkx\in\mathbb{R}^{k}, there is an integral current T,π,x\langle T,\pi,x\rangle called sliced current [AK00a, Theorems 5.6 and 5.7], characterized by the fact that for all fdπ1dπnk𝒟nk(E)fd\pi_{1}\wedge...\wedge d\pi_{n-k}\in\mathcal{D}^{n-k}(E) and ψCc(k)\psi\in C_{c}(\mathbb{R}^{k}),

(4) [kT,π,xψ(x)dx](fdπ1dπnk)=T(f.(ψπ)dπdπ1dπnk)\big{[}\int_{\mathbb{R}^{k}}\langle T,\pi,x\rangle\psi(x)dx\big{]}(fd\pi_{1}\wedge...\wedge d\pi_{n-k})=T(f.(\psi\circ\pi)d\pi\wedge d\pi_{1}...\wedge d\pi_{n-k})

where if π=(u1,,uk)\pi=(u_{1},...,u_{k}) and uju_{j} are the coordinate functions, then dπ:=du1dukd\pi:=du_{1}\wedge...\wedge du_{k}. Besides, with the notation T=set(T),θT,τTT=\llbracket\operatorname{set}(T),\theta_{T},\tau_{T}\rrbracket defined above, for almost every xkx\in\mathbb{R}^{k} we have an orientation τx\tau_{x} of set(T)π1(x)\operatorname{set}(T)\cap\pi^{-1}(x) such that

(5) T,π,x=set(T)π1(x),θT,τx,\langle T,\pi,x\rangle=\llbracket\operatorname{set}(T)\cap\pi^{-1}(x),\theta_{T},\tau_{x}\rrbracket,

see [AK00a, Theorem 9.7]. The fact that

spt(T,π,x)spt(T),\operatorname{spt}(\partial\langle T,\pi,x\rangle)\subset\operatorname{spt}(\partial T),

follows from [AK00a, Theorem 5.6 (i)] applied to the boundary T\partial T, or alternatively by iterating [AK00a, Theorem 5.6 (iii), Lemma 5.3].

1.6. Integral current spaces and intrinsic flat topology

We end this section with the definitions of integral current spaces, the intrinsic flat topology of Sormani-Wenger, and Wenger’s compactness theorem [SW11].

Definition 1.3.

An integral current space of dimension nn is a triple

C:=(X,d,T)C:=(X,d,T)

where (X,d)(X,d) is a metric space with completion called X¯\overline{X}, and TT is an nn-dimensional integral current in X¯\overline{X}, such that set(T)=X\operatorname{set}(T)=X. The mass 𝐌(C)\mathbf{M}(C) is by definition 𝐌(T)\mathbf{M}(T).

A simple example of integral current space is given by a complete, oriented Riemannian nn-manifold (M,g)(M,g) with compact boundary and finite volume: the metric space is MM endowed with the geodesic distance distg\operatorname{dist}_{g} induced by gg, and the integral current structure 1M\llbracket 1_{M}\rrbracket is the natural integral current induced by integration on MM. Recall that we allow the metric to assume \infty as value, and so we allow MM to be non-connected. In general, an integral current in a metric space EE determines uniquely an integral current space. The boundary C\partial C is the integral current space induced by T\partial T.

If C:=(X,d,T)C:=(X,d,T), C:=(X,d,T)C^{\prime}:=(X^{\prime},d^{\prime},T^{\prime}) are two integral current spaces such that there is an isometry φ:X¯X¯\varphi:\overline{X}\to\overline{X^{\prime}} with φT=T\varphi_{\sharp}T=T^{\prime} (such a φ\varphi is called a current preserving isometry [SW11, Definition 3.26]), then we say that CC and CC^{\prime} are isomorphic.

In the same way that Gromov-Hausdorff convergence generalizes Hausdorff convergence by allowing arbitrary isometric embeddings in a complete metric space, the notion of intrinsic flat convergence generalizes flat convergence as follows [SW11, Theorem 4.2].

Definition 1.4.

A sequence of integral current spaces {(Xm,dm,Sm)}m0\{(X_{m},d_{m},S_{m})\}_{m\geq 0} converges in the intrinsic flat topology to an integral current space (X,d,S)(X_{\infty},d_{\infty},S_{\infty}) when there is a complete metric space 𝐙\mathbf{Z} and there are isometric embeddings

jm:X¯m𝐙,j:X¯𝐙j_{m}:\overline{X}_{m}\to\mathbf{Z},\quad j_{\infty}:\overline{X}_{\infty}\to\mathbf{Z}

such that (jm)(Sm)(j_{m})_{\sharp}(S_{m}) converges in the flat topology to (j)(S)(j_{\infty})_{\sharp}(S_{\infty}) inside 𝐙\mathbf{Z}.

The complete metric space 𝐙\mathbf{Z} can be taken to be a Banach space if the metrics dmd_{m}, dd_{\infty} only assume finite values, by the standard Kuratowski embedding. Actually the above notion of convergence is induced by a metric called the intrinsic flat distance 𝐝\mathbf{d}_{\mathcal{F}} and defined in [SW11, Definition 1.1]: given two integral current spaces of dimension nn, C:=(X,d,S)C:=(X,d,S) and C:=(X,d,S)C^{\prime}:=(X^{\prime},d^{\prime},S^{\prime}), set

𝐝(C,C):=inf{𝐌(U)+𝐌(V)}\mathbf{d}_{\mathcal{F}}(C,C^{\prime}):=\inf\{\mathbf{M}(U)+\mathbf{M}(V)\}

where the infimum is taken over all complete metric spaces (Z,d)(Z,d) and all integral currents UU, VV in ZZ such that there are isometric embeddings φ:(X¯,d)Z\varphi:(\overline{X},d)\to Z, φ:(X¯,d)Z\varphi^{\prime}:(\overline{X^{\prime}},d^{\prime})\to Z with

φ(S)(φ)(S)=U+V.\varphi_{\sharp}(S)-(\varphi^{\prime})_{\sharp}(S^{\prime})=U+\partial V.

By [SW11, Theorem 3.27], 𝐝(C,C)=0\mathbf{d}_{\mathcal{F}}(C,C^{\prime})=0 if and only if CC and CC^{\prime} are isomorphic.

One of the fundamental properties of integral current spaces is the following compactness theorem [Wen11][SW11, Theorem 4.19]:

Theorem 1.5 ([Wen11, SW11]).

if for some constant cc, {(Xm,dm,Sm)}m0\{(X_{m},d_{m},S_{m})\}_{m\geq 0} is a sequence of nn-dimensional integral current spaces with

𝐌(Sm)+𝐌(Sm)c<,\mathbf{M}(S_{m})+\mathbf{M}(\partial S_{m})\leq c<\infty,
diam(spt(Sm))c,\operatorname{diam}(\operatorname{spt}(S_{m}))\leq c,

then there is a subsequence {(Xmk,dmk,Smk)}k0\{(X_{m_{k}},d_{m_{k}},S_{m_{k}})\}_{k\geq 0} converging in the intrinsic flat topology to an nn-dimensional integral current space (X,d,S)(X_{\infty},d_{\infty},S_{\infty}).

The above compactness result will be necessary when defining spherical Plateau solutions in Subsection 3.1.

1.7. Approximation of integral currents by polyhedral chains

We end this section with a useful approximation theorem for integral currents in spherical manifolds by polyhedral chains. This result can be simply deduced from the analogous result in finite dimensions, which in turn is completely standard.

Let (N,gN)(N,g_{N}) be an infinite-dimensional Riemannian manifold which is locally isometric to the unit sphere of an infinite-dimensional Hilbert space. In our context, a kk-dimensional integral current PP in (N,gN)(N,g_{N}) is called a polyhedral chain of dimension kk if there are smoothly embedded totally geodesic kk-simplices S1,,SmNS_{1},...,S_{m}\subset N endowed with an orientation, and integers aja_{j} so that

P=j=1maj1Sj.P=\sum_{j=1}^{m}a_{j}\llbracket 1_{S_{j}}\rrbracket.
Lemma 1.6.

For any ϵ>0\epsilon>0, and any integral current CC with compact support in (N,gN)(N,g_{N}), there is a polyhedral chain PP such that

  • CC and PP are ϵ\epsilon-close in the flat topology,

  • spt(C)\operatorname{spt}(C) and spt(P)\operatorname{spt}(P) are ϵ\epsilon-close in the Hausdorff topology,

  • |𝐌(C)𝐌(P)|ϵ|\mathbf{M}(C)-\mathbf{M}(P)|\leq\epsilon.

Here is an outline of proof. Consider an open ball BB centered at the origin in the Hilbert space 2()\ell^{2}(\mathbb{N}) where \mathbb{N} is the set of natural numbers, and consider an integral current CC with compact support in BB. For any integer L1L\geq 1, we can use the orthogonal projection onto 2({1,,L})\ell^{2}(\{1,...,L\}) to map CC to an integral current C1C_{1} compactly supported inside the finite dimensional space B2({1,,L})B\cap\ell^{2}(\{1,...,L\}). If LL is chosen large enough, C1C_{1} will be as close to CC as we wish in the sense of Lemma 1.6. Now C1C_{1} is also an integral current in the sense of Federer-Fleming (see the appendix of [AK00a]), so the usual approximation results ([Fed69, Sections 4.1 and 4.2], [DP14]) in finite dimension can be applied and give the desired polyhedral chain approximation P1P_{1} in B2({1,,L})B\cap\ell^{2}(\{1,...,L\}). This construction clearly generalizes to small balls in (N,gN)(N,g_{N}) if NN is separable, which can always be assumed to be true since spt(C)\operatorname{spt}(C) is compact. In the general case where the support of CC is not contained in a small ball of NN, one can argue using a partition of unity and construct via an interpolation map a current C2C_{2} arbitrarily close to CC in the sense of Lemma 1.6, which is locally finite dimensional in the following sense: for any xspt(C2)x\in\operatorname{spt}(C_{2}), there is a ball BxB_{x} containing xx such that spt(C2)Bx\operatorname{spt}(C_{2})\cap B_{x} is contained in a totally geodesic finite dimensional plane of BxB_{x}. The rest is standard as before.

2. Preliminaries on the barycenter map

In [BCG95, BCG96], Besson-Courtois-Gallot proved that the normalized volume entropy on a closed hyperbolic manifold of dimension at least 33 is uniquely achieved at the hyperbolic metric. The proof of this striking result relies on the barycenter map. In this section333I would like to thank Cosmin Manea for corrections and useful discussions about this section., we define a variant of the barycenter map used in [BCG95, BCG96] (see also [Sam99]), which we will need in the proofs of the uniqueness of spherical Plateau solutions. Instead of working with L2L^{2} functions on a boundary at infinity as in [BCG95], we directly work with 2\ell^{2} functions on the underlying group Γ\Gamma. This setup is better adapted to extensions to more general situations (33-manifolds, Plateau Dehn fillings). This section only treats hyperbolic manifolds, even though everything carries out more generally in the locally symmetric rank one case. All the results here are essentially contained in [BCG95].

2.1. Definition of the barycenter map

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, freely and properly on (M~,g0)(\tilde{M},g_{0}). Let S{S^{\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 λΓ:ΓEnd(2(Γ))\lambda_{\Gamma}:\Gamma\to\operatorname{End}(\ell^{2}(\Gamma)), and let S/λΓ(Γ){S^{\infty}}/\lambda_{\Gamma}(\Gamma) be the quotient manifold endowed with the standard round metric (see Subsection 3.1 for more details).

It is well-known that the distance functions on (M~,g0)(\tilde{M},g_{0}) satisfy the following Hessian lower bound: for any fixed wM~w\in\tilde{M}, at any point different from ww we have

Dddistg0(w,.)Idddistg0(w,.)ddistg0(w,.).Dd\operatorname{dist}_{g_{0}}(w,.)\geq\operatorname{Id}-d\operatorname{dist}_{g_{0}}(w,.)\otimes d\operatorname{dist}_{g_{0}}(w,.).

Distance functions are not smooth so in the setting of hyperbolic manifolds, it is more convenient from a technical point of view to work with modified distance functions called ρw\rho_{w}. Fix a smooth strictly convex increasing function

ϰ:[0,)[0,)\varkappa:[0,\infty)\to[0,\infty)

such that limtt1ϰ(t)=1\lim_{t\to\infty}t^{-1}\varkappa(t)=1, ϰ(t)<1\varkappa^{\prime}(t)<1 for all tt, and 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

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

Such a function exists, for example if we set

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

where cc is a positive constant, it is an exercise to check that (6) holds whenever cc is large enough.

Definition 2.1.

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

(7) 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+/λΓ(Γ)M\mathbb{S}^{+}/\lambda_{\Gamma}(\Gamma)\to M

is also denoted by Bar\mathrm{Bar}.

2.2. Notations and a priori Lipschitz bounds

The regularity of the barycenter map

Bar:𝕊+/λΓ(Γ)M\mathrm{Bar}:\mathbb{S}^{+}/\lambda_{\Gamma}(\Gamma)\to M

is not completely immediate. To avoid discussing such issues, we will only consider the barycenter map restricted to the support of polyhedral chains in 𝕊+/λΓ(Γ)\mathbb{S}^{+}/\lambda_{\Gamma}(\Gamma). Recall that by definition (see Subsection 1.7), a kk-dimensional polyhedral chain PP in 𝕊+/λΓ(Γ)\mathbb{S}^{+}/\lambda_{\Gamma}(\Gamma) is a kk-dimensional integral current that can be written as

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

where aja_{j} are integers and

S1,,Sm𝕊+/λΓ(Γ)S/λΓ(Γ)S_{1},...,S_{m}\subset\mathbb{S}^{+}/\lambda_{\Gamma}(\Gamma)\subset{S^{\infty}}/\lambda_{\Gamma}(\Gamma)

are finitely many smoothly embedded totally geodesic kk-simplices endowed with an orientation. In particular the support of a polyhedral chain is by definition a finite union of totally geodesic simplices. Given a polyhedral chain PP as above, each simplex SjS_{j} lifts to a simplex S~j\tilde{S}_{j} in 𝕊+2(Γ)\mathbb{S}^{+}\subset\ell^{2}(\Gamma). We claim that there is a finite set 𝐒jΓ\mathbf{S}_{j}\subset\Gamma depending only on S~j\tilde{S}_{j} such that any element of S~j\tilde{S}_{j} is a function with support in 𝐒j\mathbf{S}_{j}. Indeed, the kk-simplex S~j\tilde{S}_{j} is the convex hull of its extremal points f0,,fk𝕊+f_{0},...,f_{k}\in\mathbb{S}^{+}. If 𝐒j\mathbf{S}_{j} denotes the union of the finite supports of f0,,fkf_{0},...,f_{k}, then any linear combination of those functions has support contained in 𝐒j\mathbf{S}_{j}, and that proves the claim. Now given a polyhedral chain PP in 𝕊+/λΓ(Γ)\mathbb{S}^{+}/\lambda_{\Gamma}(\Gamma), one can check without difficulty that the restriction

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

is continuous (and smooth on each simplex by the discussion below).

Consider f𝕊+f\in\mathbb{S}^{+}. For vTxM~v\in T_{x}\tilde{M}, set

(8) Hf(v,v):=γf2(γ)|dxργ.o(v)|2.H_{f}(v,v):=\sum_{\gamma}f^{2}(\gamma)|d_{x}\rho_{\gamma.o}(v)|^{2}.

The endomorphism HfH_{f} is symmetric, it satisfies

TrHf<1\operatorname{Tr}H_{f}<1

due to the fact that |ργ.o|<1|\nabla\rho_{\gamma.o}|<1, and has eigenvalues

0μ1(f)μn(f)<1.0\leq\mu_{1}(f)\leq...\leq\mu_{n}(f)<1.

In particular IdHf\operatorname{Id}-H_{f} has strictly positive determinant. For vTxM~v\in T_{x}\tilde{M}, set

Kf(v,v):=γf2(γ)Ddxργ.o(v,v).K_{f}(v,v):=\sum_{\gamma}f^{2}(\gamma)Dd_{x}\rho_{\gamma.o}(v,v).

By strict convexity of the modified distance functions,

(9) vTxM~{0},Kf(v,v)>0.\forall v\in T_{x}\tilde{M}\setminus\{0\},\quad K_{f}(v,v)>0.

The barycenter xx of ff is characterized by the equation

(10) γΓf2(γ)dxργ.o(.)=0TxM~.\sum_{\gamma\in\Gamma}f^{2}(\gamma)d_{x}\rho_{\gamma.o}(.)=0\in T_{x}^{*}\tilde{M}.

For 1kn1\leq k\leq n, let ξ\xi be a totally geodesic kk-simplex contained in 𝕊+\mathbb{S}^{+}, embedded in S{S^{\infty}}, passing through f𝕊+f\in\mathbb{S}^{+}. Then the restriction of Bar\mathrm{Bar} to ξ\xi is smoothly differentiable around ff because of the implicit function theorem and (9), (10). Let QQ be the tangent nn-plane of ξ\xi at ff. The differential of Bar\mathrm{Bar} along QQ is denoted by dBar|Q:QTBar(p)M~d\mathrm{Bar}\big{|}_{Q}:Q\to T_{\mathrm{Bar}(p)}\tilde{M}. Sometimes we will drop the subscript |Q|_{Q} when the choice of tangent kk-plane is clear. By differentiating (10) with respect to ff we obtain for all f˙Q\dot{f}\in Q:

γΓ2f(γ)f˙(γ)dxργ.o(.)+γΓf2(γ)Ddxργ.o(dBar|Q(f˙),.)=0.\sum_{\gamma\in\Gamma}2f(\gamma)\dot{f}(\gamma)d_{x}\rho_{\gamma.o}(.)+\sum_{\gamma\in\Gamma}f^{2}(\gamma)Dd_{x}\rho_{\gamma.o}(d\mathrm{Bar}\big{|}_{Q}(\dot{f}),.)=0.

After applying Cauchy-Schwarz, we get the following: for all vTxM~v\in T_{x}\tilde{M} and f˙Q\dot{f}\in Q, with f˙2=1\|\dot{f}\|_{\ell^{2}}=1,

(11) Kf(dBar|Q(f˙),v)2[Hf(v,v)]1/2K_{f}(d\mathrm{Bar}\big{|}_{Q}(\dot{f}),v)\leq 2[H_{f}(v,v)]^{1/2}

and

(12) |JacBar|Q|2n(detHf)1/2detKf2n(detHf)1/2det(IdHf).|\operatorname{Jac}\mathrm{Bar}\big{|}_{Q}|\leq 2^{n}\frac{(\det H_{f})^{1/2}}{\det K_{f}}\leq 2^{n}\frac{(\det H_{f})^{1/2}}{\det(\operatorname{Id}-H_{f})}.

Going from (11) to the first inequality in (12) is an application of the Gram-Schmidt orthonormalization process for matrices, see the proof of [BCG96, Lemma 5.4]. The second inequality in (12) is a consequence of the inequality

(13) KfIdHfK_{f}\geq\operatorname{Id}-H_{f}

which in turn follows from (6).

For any f𝕊+f\in\mathbb{S}^{+}, let μ1(f)μn(f)\mu_{1}(f)\leq...\leq\mu_{n}(f) be the eigenvalues of the endomorphism HfH_{f} defined in (8). The following is an a priori Lipschitz bound corresponding to [BCG95, Lemma 7.5.a].

Lemma 2.2.

Let κ>0\kappa>0 and let α𝕊+\alpha\subset\mathbb{S}^{+} be a connected continuous piecewise geodesic curve. Suppose that for all fαf\in\alpha, μn(f)1κ/2\mu_{n}(f)\leq 1-\kappa/2. Then

lengthg0(Bar(α))K1length(α)\operatorname{length}_{g_{0}}(\mathrm{Bar}(\alpha))\leq K_{1}\operatorname{length}(\alpha)

for a constant K1K_{1} depending only on κ\kappa.

Proof.

Given fαf\in\alpha, let VV be the tangent 11-plane of α\alpha at ff, let f˙V\dot{f}\in V with f˙2=1\|\dot{f}\|_{\ell^{2}}=1, and suppose that dfBar(f˙)0d_{f}\mathrm{Bar}(\dot{f})\neq 0. Set v:=dfBar(f˙)|dfBar(f˙)|TxM~v:=\frac{d_{f}\mathrm{Bar}(\dot{f})}{|d_{f}\mathrm{Bar}(\dot{f})|}\in T_{x}\tilde{M}.

By (11) and (13),

(14) |dfBar(f˙)|2(Hf(v,v))1/21Hf(v,v).|d_{f}\mathrm{Bar}(\dot{f})|\leq 2\frac{(H_{f}(v,v))^{1/2}}{1-H_{f}(v,v)}.

The lemma readily follows from integrating (14) along α\alpha.

Here is another a priori bound corresponding to [BCG95, Lemma 7.5.b].

Lemma 2.3.

Let f,f𝕊+f,f^{\prime}\in\mathbb{S}^{+} and let β\beta be the geodesic segment joining Bar(f)\mathrm{Bar}(f) and Bar(f)\mathrm{Bar}(f^{\prime}). Let PP be the parallel transport from Bar(f)\mathrm{Bar}(f) to Bar(f)\mathrm{Bar}(f^{\prime}) along β\beta. Then

HfPHfK2(lengthg0(β)+ff2)\|H_{f^{\prime}}\circ P-H_{f}\|\leq K_{2}(\operatorname{length}_{g_{0}}(\beta)+\|f-f^{\prime}\|_{\ell^{2}})

for a constant K2K_{2}.

Proof.

Let Y0Y_{0} be a unit norm tangent vector of MM based at Bar(f)\mathrm{Bar}(f) and let Y2Y_{2} be its parallel transport at Bar(f)\mathrm{Bar}(f^{\prime}) along β\beta. Then, writing x:=Bar(f),x:=Bar(f)x:=\mathrm{Bar}(f),x^{\prime}:=\mathrm{Bar}(f^{\prime}),

(15) |Hf(Y2,Y2)Hf(Y0,Y0)|=|γ(f)2(γ)|dxργ.o(Y2)|2γf2(γ)|dxργ.o(Y0)|2||γf2(γ)(|dxργ.o(Y2)|2|dxργ.o(Y0)|2)|+|γ((f)2(γ)f2(γ))|dxργ.o(Y2)|2|.\begin{split}&|H_{f^{\prime}}(Y_{2},Y_{2})-H_{f}(Y_{0},Y_{0})|\\ =&|\sum_{\gamma}(f^{\prime})^{2}(\gamma)|d_{x^{\prime}}\rho_{\gamma.o}(Y_{2})|^{2}-\sum_{\gamma}f^{2}(\gamma)|d_{x}\rho_{\gamma.o}(Y_{0})|^{2}|\\ \leq&|\sum_{\gamma}f^{2}(\gamma)\big{(}|d_{x^{\prime}}\rho_{\gamma.o}(Y_{2})|^{2}-|d_{x}\rho_{\gamma.o}(Y_{0})|^{2}\big{)}|\\ &+|\sum_{\gamma}\big{(}(f^{\prime})^{2}(\gamma)-f^{2}(\gamma)\big{)}|d_{x^{\prime}}\rho_{\gamma.o}(Y_{2})|^{2}|.\end{split}

The Hessian of the smooth modified distance functions ργ.o\rho_{\gamma.o} is uniformly bounded from above. This bound controls uniformly the terms (|dxρw(Y2)|2|dxρw(Y0)|2)\big{(}|d_{x^{\prime}}\rho_{w}(Y_{2})|^{2}-|d_{x}\rho_{w}(Y_{0})|^{2}\big{)} by integration along β\beta, and we conclude the proof with Cauchy-Schwarz.

2.3. The Jacobian bound

We are now ready to state the main estimates for the Jacobian of the barycenter map, which is sharp and is the key technical point in [BCG95, BCG96].

Lemma 2.4.

[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

(16) |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,

(17) |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

(18) 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).
Proof.

The upper bound for the Jacobian is proved by combining (12) and [BCG95, Proposition B.1].

Let us move on to the second part of the lemma. As before, denote by 0μ1(f)μn(f)<10\leq\mu_{1}(f)\leq...\leq\mu_{n}(f)<1 the eigenvalues of HfH_{f}. Let θ>1\theta>1 be such that Tr(θHf)=1\operatorname{Tr}(\theta H_{f})=1. By the key bound in [BCG95, Proposition B.5], there is a universal constant A>0A>0 such that

2n(detHf)1/2det(IdHf)(4n(n1)2)n/2(1Aj=1n(θμj(f)1n)2).2^{n}\frac{(\det H_{f})^{1/2}}{\det\big{(}\operatorname{Id}-H_{f}\big{)}}\leq\big{(}\frac{4n}{(n-1)^{2}}\big{)}^{n/2}\big{(}1-A\sum_{j=1}^{n}(\theta\mu_{j}(f)-\frac{1}{n})^{2}\big{)}.

Hence for any η>0\eta^{\prime}>0 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 for a small η\eta, then

(19) μn(f)1n+η.\mu_{n}(f)\leq\frac{1}{n}+\eta^{\prime}.

and so by (11) we have for all norm 11 tangent vector uQ\vec{u}\in Q:

(20) |dBar|Q(u)|(4n(n1)2)1/2+cη′′|d\mathrm{Bar}\big{|}_{Q}(\vec{u})|\leq\big{(}\frac{4n}{(n-1)^{2}}\big{)}^{1/2}+c^{\prime\prime}_{\eta}

where limη0cη′′=0\lim_{\eta\to 0}c^{\prime\prime}_{\eta}=0. Therefore 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 (20) forces the following to hold: for all norm 11 tangent vector uQ\vec{u}\in Q,

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

where limη0cη=0\lim_{\eta\to 0}c_{\eta}=0.

We want to propagate the estimate (20) on dBar\|d\mathrm{Bar}\| to a whole “neighborhood” of ff inside 𝕊+\mathbb{S}^{+}. Suppose that a continuous piecewise geodesic curve α𝕊+\alpha\subset\mathbb{S}^{+} joins ff to ff^{\prime} and has length less than a number η\eta. Let us check, as an intermediate step, that for κ>0\kappa>0,

(21) if μn(f)1κ\mu_{n}(f)\leq 1-\kappa, then μn(f)1κ+cη\mu_{n}(f^{\prime})\leq 1-\kappa+c^{\prime}_{\eta}

where the constant cηc^{\prime}_{\eta} depends only on η\eta and satisfies limη0cη=0\lim_{\eta\to 0}c^{\prime}_{\eta}=0. Indeed, we can proceed as in [BCG95, Lemma 7.5]. Let K1,K2K_{1},K_{2} as in Lemmas 2.2 and 2.3. Suppose that μn(f)1κ\mu_{n}(f)\leq 1-\kappa, and that η\eta is small so that there is K3K_{3} with K2(K1+1)+1<K3<κ2ηK_{2}(K_{1}+1)+1<K_{3}<\frac{\kappa}{2\eta}. To argue towards a contradiction, assume that (21) is not true and that there is a point f1𝕊+f_{1}\in\mathbb{S}^{+} joined to ff by a continuous piecewise geodesic curve α𝕊+\alpha\subset\mathbb{S}^{+} of length less than η\eta, such that

μn(f1)1κ+K3η.\mu_{n}(f_{1})\geq 1-\kappa+K_{3}\eta.

By truncating the curve, we can also assume that f1f_{1} is the only point on α\alpha which satisfies the above condition, so that μn(f0)1κ/2\mu_{n}(f_{0})\leq 1-\kappa/2 for every f0σf_{0}\in\sigma. We have μn(f1)1κ+K3η\mu_{n}(f_{1})\geq 1-\kappa+K_{3}\eta, but by Lemmas 2.2 and 2.3, we would also have

μn(f1)μn(f)+K2(K1+1)η1κ+K2(K1+1)η<1κ+K3η.\mu_{n}(f_{1})\leq\mu_{n}(f)+K_{2}(K_{1}+1)\eta\leq 1-\kappa+K_{2}(K_{1}+1)\eta<1-\kappa+K_{3}\eta.

This is a contradiction, thus (21) is checked.

We now conclude using (21). Suppose that a continuous piecewise geodesic curve α𝕊+\alpha\subset\mathbb{S}^{+} joins ff to f𝕊+f^{\prime}\in\mathbb{S}^{+} and has length less than a small number η\eta. Let u\vec{u} be a unit tangent vector of α\alpha at a point of α\alpha. By (11), (13), (19) and (21),

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

where cηc_{\eta} is a constant depending only on η>0\eta>0 such that limη0cη=0\lim_{\eta\to 0}c_{\eta}=0. We choose η\eta small enough and we integrate the previous inequality from ff to ff^{\prime} along the points of α\alpha where the tangent vector exists and the differential of Bar\mathrm{Bar} is well-defined. We readily obtain the desired conclusion:

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).

3. The spherical Plateau problem

3.1. Setup and main definitions

Let Γ\Gamma be a countable group. Let 2(Γ)\ell^{2}(\Gamma) be the space of 2\ell^{2} real functions on Γ\Gamma. Set

S:={f:Γ;f2=1}.{S^{\infty}}:=\{f:\Gamma\to\mathbb{R};\quad\|f\|_{\ell^{2}}=1\}.

Note that SS^{\infty} may be finite dimensional, and that SS^{\infty} implicitly depends on Γ\Gamma. The 2\ell^{2}-norm induces a Riemannian metric 𝐠Hil\mathbf{g}_{\mathrm{Hil}} on the unit sphere S{S^{\infty}}. The group Γ\Gamma acts linearly isometrically on 2(Γ)\ell^{2}(\Gamma), and thus on S{S^{\infty}}, by the (left) regular representation

λΓ:ΓEnd(2(Γ))\lambda_{\Gamma}:\Gamma\to\operatorname{End}(\ell^{2}(\Gamma))

defined as follows: for all γΓ,xΓ,fS\gamma\in\Gamma,x\in\Gamma,f\in{S^{\infty}},

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

Denote by

(S/λΓ(Γ),𝐠Hil)({S^{\infty}}/\lambda_{\Gamma}(\Gamma),\mathbf{g}_{\mathrm{Hil}})

the corresponding spherical quotient with the quotient metric. The action of Γ\Gamma on S{S^{\infty}} is not free exactly when Γ\Gamma has torsion elements. When Γ\Gamma is torsion-free and non-trivial, then it is not hard to check that S{S^{\infty}} is an infinite-dimensional contractible sphere, Γ\Gamma acts properly freely on S{S^{\infty}} by the regular representation and the quotient space S/λΓ(Γ){S^{\infty}}/\lambda_{\Gamma}(\Gamma) is in fact a classifying space for Γ\Gamma (namely a K(Γ,1)K(\Gamma,1) space).

Let n0n\geq 0 be an integer. Consider the following homology groups defined using integral currents:

𝒵n(S/λΓ(Γ)):={T;T is an integral n-current with\displaystyle\mathcal{Z}_{n}({S^{\infty}}/\lambda_{\Gamma}(\Gamma)):=\{T;\quad\text{$T$ is an integral $n$-current with}
compact support in S/λΓ(Γ)},\displaystyle\text{compact support in ${S^{\infty}}/\lambda_{\Gamma}(\Gamma)$}\},
n(S/λΓ(Γ)):={D;D is an integral (n+1)-current with\displaystyle\mathcal{B}_{n}({S^{\infty}}/\lambda_{\Gamma}(\Gamma)):=\{\partial D;\quad\text{$D$ is an integral $(n+1)$-current with}
compact support in S/λΓ(Γ)},\displaystyle\text{compact support in ${S^{\infty}}/\lambda_{\Gamma}(\Gamma)$}\},
𝐇n(S/λΓ(Γ)):=𝒵n(S/λΓ(Γ))/n(S/λΓ(Γ)).\mathbf{H}_{n}({S^{\infty}}/\lambda_{\Gamma}(\Gamma)):=\mathcal{Z}_{n}({S^{\infty}}/\lambda_{\Gamma}(\Gamma))/\mathcal{B}_{n}({S^{\infty}}/\lambda_{\Gamma}(\Gamma)).

There is a natural morphism

π^:H(Γ;)𝐇(S/λΓ(Γ))\hat{\pi}:H_{*}(\Gamma;\mathbb{Z})\to\mathbf{H}_{*}({S^{\infty}}/\lambda_{\Gamma}(\Gamma))

where H(Γ;)H_{*}(\Gamma;\mathbb{Z}) are the singular homology groups of the group Γ\Gamma with coefficients in \mathbb{Z}. Given a group homology class hHn(Γ;)h\in H_{n}(\Gamma;\mathbb{Z}), consider the space

𝒞(h)\mathscr{C}(h)

of boundaryless nn-dimensional integral currents with compact supports inside S/λΓ(Γ){S^{\infty}}/\lambda_{\Gamma}(\Gamma) which represent the homology class π^(h)𝐇n(S/λΓ(Γ))\hat{\pi}(h)\in\mathbf{H}_{n}({S^{\infty}}/\lambda_{\Gamma}(\Gamma)); a more careful definition of π^\hat{\pi} and 𝒞(h)\mathscr{C}(h) is given in Subsection 3.2. For simplicity, we will sometimes call these currents “cycles representing hh”. Recall that the notion of mass 𝐌\mathbf{M} for an integral current is reviewed in Section 1. We define the spherical volume of a group homology class as follows:

Definition 3.1.

Let hHn(Γ;)h\in H_{n}(\Gamma;\mathbb{Z}). The spherical volume of hh is defined as

SphereVol(h)=inf{𝐌(C);C𝒞(h)}.\operatorname{SphereVol}(h)=\inf\{\mathbf{M}(C);\quad C\in\mathscr{C}(h)\}.

This is a homological generalization of the spherical volume first introduced by Besson-Courtois-Gallot in the Riemannian setting [BCG91, Section 3,I]. This invariant can be computed in special cases, as will be explained in Sections 4 and 5.

We can now define spherical Plateau solutions:

Definition 3.2.

We call spherical Plateau solution for hh any nn-dimensional integral current space CC_{\infty} which is the limit in the intrinsic flat topology of a sequence {Ci}𝒞(h)\{C_{i}\}\subset\mathscr{C}(h) such that

limi𝐌(Ci)=SphereVol(h).\lim_{i\to\infty}\mathbf{M}(C_{i})=\operatorname{SphereVol}(h).

The spherical Plateau problem consists of studying spherical Plateau solutions and their relation to the original pair (Γ,h)(\Gamma,h).

Note that spherical Plateau solutions CC_{\infty} for a group homology class hh always exist by the compactness theorem of Wenger [Wen11][SW11, Theorem 4.19] which serves as a replacement for the compactness theorem of Federer-Fleming in finite dimensions (it suffices to check that given CiC_{i} as in the above definition, their masses as well as their diameters are uniformly bounded). Any spherical Plateau solution CC_{\infty} is an integral current space without boundary, with uniformly bounded diameter. The mass of CC_{\infty} satisfies

𝐌(C)SphereVol(h)\mathbf{M}(C_{\infty})\leq\operatorname{SphereVol}(h)

by lower semicontinuity of the mass under instrinsic flat convergence [SW11, Theorem 4.6]. However the converse is unclear:

Question 1 (Volume convergence).

Given a group homology class hh and a spherical Plateau solution CC_{\infty} for hh, is it always true that

𝐌(C)=SphereVol(h)?\mathbf{M}(C_{\infty})=\operatorname{SphereVol}(h)?

An nn-dimensional integral current without boundary SS in a complete metric space (E,d)(E,d) is called mass-minimizing if for any (n+1)(n+1)-dimensional integral current DD supported in (E,d)(E,d), 𝐌(S)𝐌(S+D)\mathbf{M}(S)\leq\mathbf{M}(S+\partial D). In striking contrast with the finite dimensional compact case, there is no non-trivial mass-minimizing integral current without boundary inside the manifold S/λΓ(Γ){S^{\infty}}/\lambda_{\Gamma}(\Gamma) when Γ\Gamma is torsion-free. In particular, a spherical Plateau solution CC_{\infty} for hh is in general not isometrically embedded inside S/λΓ(Γ){S^{\infty}}/\lambda_{\Gamma}(\Gamma) as a cycle in 𝒞(h)\mathscr{C}(h). That fact follows from the existence of a distance decreasing flow on the subset of nonnegative functions of S/λΓ(Γ){S^{\infty}}/\lambda_{\Gamma}(\Gamma), as explained in Remark 3.4. This raises the following:

Question 2 (Variational structure).

Does any spherical Plateau solution isometrically embed as a mass-minimizing integral current in a quotient of a Hilbert unit sphere?

Instead of focusing on a fixed spherical Plateau solution, for a given dimension, one can instead look at the set of all spherical volumes of group homology classes, and the space of all spherical Plateau solutions. In Section 6, we will establish the existence of accumulation phenomena for spherical Plateau solutions in all dimensions at least 44.

Remark 3.3.

All the above definitions for the spherical Plateau problem can be extended or modified to study general orthogonal representations different from the regular representation.

3.2. Definition of 𝒞(h)\mathscr{C}(h).

Consider a countable group Γ\Gamma. Let us say more about the natural map

(22) π^:H(Γ;)𝐇(S/λΓ(Γ)).\hat{\pi}:H_{*}(\Gamma;\mathbb{Z})\to\mathbf{H}_{*}({S^{\infty}}/\lambda_{\Gamma}(\Gamma)).

If Γ\Gamma is finite, there is a point in the finite dimensional sphere S{S^{\infty}} which is fixed by the whole group Γ\Gamma so it is natural to define π^\hat{\pi} to be the trivial map sending everything to {0}\{0\}. Assume that Γ\Gamma is infinite. In that case, set

S,:={xS;there is no g1 such that λΓ(g)x=x}.{S^{\infty,*}}:=\{x\in{S^{\infty}};\quad\text{there is no $g\neq 1$ such that $\lambda_{\Gamma}(g)x=x$}\}.

Then one can check that the restriction of the action of Γ\Gamma on S,{S^{\infty,*}} is proper free, S,{S^{\infty,*}} is still weakly contractible (i.e. all the higher homotopy groups are trivial) and S,/λΓ(Γ){S^{\infty,*}}/\lambda_{\Gamma}(\Gamma) is a non-complete, infinite-dimensional, Hilbert Riemannian manifold and a classifying space for Γ\Gamma. Since S,/λΓ(Γ){S^{\infty,*}}/\lambda_{\Gamma}(\Gamma) is an open Riemannian manifold, it is well-known that 𝐇(S,/λΓ(Γ))\mathbf{H}_{*}({S^{\infty,*}}/\lambda_{\Gamma}(\Gamma)) is isomorphic to the singular homology groups H(S,/λΓ(Γ);)H_{*}({S^{\infty,*}}/\lambda_{\Gamma}(\Gamma);\mathbb{Z}) [RS09]. Recall that for any classifying space BΓB\Gamma, by definition

H(Γ;)=H(BΓ;).H_{*}(\Gamma;\mathbb{Z})=H_{*}(B\Gamma;\mathbb{Z}).

Then we set

π^:H(Γ;)=H(S,/λΓ(Γ);)=𝐇(S,/λΓ(Γ))𝐇(S/λΓ(Γ))\hat{\pi}:H_{*}(\Gamma;\mathbb{Z})=H_{*}({S^{\infty,*}}/\lambda_{\Gamma}(\Gamma);\mathbb{Z})=\mathbf{H}_{*}({S^{\infty,*}}/\lambda_{\Gamma}(\Gamma))\to\mathbf{H}_{*}({S^{\infty}}/\lambda_{\Gamma}(\Gamma))

where the last arrow is induced by the inclusion map S,/λΓ(Γ)S/λΓ(Γ){S^{\infty,*}}/\lambda_{\Gamma}(\Gamma)\subset{S^{\infty}}/\lambda_{\Gamma}(\Gamma). When Γ\Gamma is torsion-free infinite, π^\hat{\pi} is an isomorphism because S,/λΓ(Γ)=S/λΓ(Γ){S^{\infty,*}}/\lambda_{\Gamma}(\Gamma)={S^{\infty}}/\lambda_{\Gamma}(\Gamma).

Let hHn(Γ;)h\in H_{n}(\Gamma;\mathbb{Z}). Our definition of the family of cycles 𝒞(h)\mathscr{C}(h) is then the following:

𝒞(h):={\displaystyle\mathscr{C}(h):=\{ integral currents with compact support in S/λΓ(Γ){S^{\infty}}/\lambda_{\Gamma}(\Gamma)
representing the class π^(h)𝐇n(S/λΓ(Γ))}.\displaystyle\text{representing the class $\hat{\pi}(h)\in\mathbf{H}_{n}({S^{\infty}}/\lambda_{\Gamma}(\Gamma))$}\}.

3.3. Basic examples

3.3.1. Manifolds

To any smooth closed oriented nn-manifold MM corresponds a spherical Plateau problem. Indeed, let Γ:=π1(M)\Gamma:=\pi_{1}(M). By elementary topology, there is a canonical homotopy class of continuous maps from MM to a classifying space BΓB\Gamma for Γ\Gamma which induces an isomorphism on fundamental groups. This homotopy class determines a unique homology class

hMHn(Γ;)=Hn(BΓ;)h_{M}\in H_{n}(\Gamma;\mathbb{Z})=H_{n}(B\Gamma;\mathbb{Z})

called the induced class. Often we will make use of the following notation

SphereVol(M):=SphereVol(hM),\operatorname{SphereVol}(M):=\operatorname{SphereVol}(h_{M}),

and the spherical Plateau solutions for hMh_{M} will alternatively be called spherical Plateau solutions for MM.

One could hope that the area-minimization process in the spherical Plateau problem “simplifies” the topology of MM.

Question 3 (Complexity).

How does the topological complexity of a closed oriented manifold MM compares with that of its spherical Plateau solutions?

Question 4 (Stabilization).

Consider a closed oriented manifold MM and its induced homology class hMHn(Γ;)h_{M}\in H_{n}(\Gamma;\mathbb{Z}). How do SphereVol(mhM)\operatorname{SphereVol}(mh_{M}) and its corresponding spherical Plateau solutions behave with respect to the integer mm? What are the properties of limm1mSphereVol(mhM)\lim_{m\to\infty}\frac{1}{m}\operatorname{SphereVol}(mh_{M})?

For related results about multiples of homology classes in finite-dimensional Riemannian manifolds, see [Fed74] [Law75] [Mor84] [Whi84] [Liu23]. For progress on similar questions for the volume entropy, see [BS23] and references therein.

3.3.2. Amenable groups

Suppose Γ\Gamma is finite and let hHn(Γ;)h\in H_{n}(\Gamma;\mathbb{Z}). Then by definition of 𝒞(h)\mathscr{C}(h) (see Subsection 3.2), the zero integral current in S/λΓ(Γ){S^{\infty}}/\lambda_{\Gamma}(\Gamma) is in 𝒞(h)\mathscr{C}(h), so SphereVol(h)=0\operatorname{SphereVol}(h)=0.

If Γ=\Gamma=\mathbb{Z} and hH1(Γ;)h\in H_{1}(\Gamma;\mathbb{Z}) (for instance if M=S1M=S^{1} and h=hMh=h_{M}), then SphereVol(h)=0\operatorname{SphereVol}(h)=0 and the only spherical Plateau solution is the zero integral current space. To see this, one can represent hh by a disjoint union of embedded oriented loops c1,,cLS/λΓ(Γ)c_{1},...,c_{L}\subset{S^{\infty}}/\lambda_{\Gamma}(\Gamma), where each cic_{i} is the projection of some segment in S{S^{\infty}} joining an element ff to γi.f\gamma_{i}.f where γiΓ\gamma_{i}\in\Gamma depends on cic_{i}. Now one can separately move each cic_{i} homotopically to a curve of arbitrarily small length, since for each ii there is fif_{i} such that fiγi.fi2\|f_{i}-\gamma_{i}.f_{i}\|_{\ell^{2}} is arbitrarily small.

In fact, these arguments can be generalized to any amenable groups thanks to the so-called Dixmier condition: a finitely generated group Γ\Gamma is amenable if and only if for any ϵ>0\epsilon>0 and any finite subset SΓS\subset\Gamma, there is uSu\in{S^{\infty}} such that s.uu2ϵ\|s.u-u\|_{\ell^{2}}\leq\epsilon for all sSs\in S.

In particular it is possible to show that if Γ\Gamma is amenable, then SphereVol(h)=0\operatorname{SphereVol}(h)=0 for all hH(Γ;)h\in H_{*}(\Gamma;\mathbb{Z}). With a bit more work, one can work out a vanishing theorem analogous to the vanishing theorem for the simplicial volume due to Gromov [Gro82, Section 3.1]. In that sense the spherical Plateau problem, like many geometric invariants (simplicial volume, volume entropy etc.), is only sensitive to “large” groups.

Question 5 (Vanishing of spherical volume).

What are useful characterizations of group homology classes with vanishing spherical volume?

3.4. Some distance non-increasing maps

In this subsection, we collect some useful distance non-increasing maps.

3.4.1. Group homomorphisms

Let F,GF,G be countable groups with regular representations λF,λG\lambda_{F},\lambda_{G}, and a homomorphism θ:FG\theta:F\to G. The homomorphism θ\theta induces a map sending u2(F)u\in\ell^{2}(F) to Θ(u)2(G)\Theta(u)\in\ell^{2}(G) defined by

(23) Θ(u)(y):={(xθ1(y)|u(x)|2)1/2ifyθ(F)0ifyθ(F).\Theta(u)(y):=\left\{\begin{array}[]{rcl}\big{(}\sum_{x\in\theta^{-1}(y)}|u(x)|^{2}\big{)}^{1/2}&\mbox{if}&y\in\theta(F)\\ 0&\mbox{if}&y\notin\theta(F)\end{array}\right..

By Cauchy-Schwarz, this map is distance non-increasing. One checks directly that this map is also FF-equivariant, which implies that there is an induced map

Θ:S/λF(F)S/λG(G)\Theta:{S^{\infty}}/\lambda_{F}(F)\to{S^{\infty}}/\lambda_{G}(G)

which is still distance non-increasing.

Now let hHn(F;)h\in H_{n}(F;\mathbb{Z}) and let θ(h)Hn(G;)\theta_{*}(h)\in H_{n}(G;\mathbb{Z}) be the class equal to the push-forward of hh by θ\theta. From the above, one can show that:

(24) SphereVol(h)SphereVol(θ(h)).\operatorname{SphereVol}(h)\geq\operatorname{SphereVol}(\theta_{*}(h)).

3.4.2. Absolute value

Let Γ\Gamma be countable, hHn(Γ;)h\in H_{n}(\Gamma;\mathbb{Z}). By a similar argument using Cauchy-Schwarz, note that there is a distance non-increasing equivariant map

𝒜:SS\mathcal{A}:{S^{\infty}}\to{S^{\infty}}
u|u|,u\mapsto|u|,

which descends to a distance non-increasing map

(25) 𝒜:S/λΓ(Γ)S/λΓ(Γ).\mathcal{A}:{S^{\infty}}/\lambda_{\Gamma}(\Gamma)\to{S^{\infty}}/\lambda_{\Gamma}(\Gamma).

3.4.3. Spherical convolution

Suppose here that Γ\Gamma is countably infinite and has no torsion. Let η~:Γ[0,1]\tilde{\eta}:\Gamma\to[0,1] be a function with γη~(γ)=1\sum_{\gamma}\tilde{\eta}(\gamma)=1, and suppose also that η~\tilde{\eta} is strictly positive everywhere. The “spherical convolution” of f0f_{0} by η~\tilde{\eta} is defined as

η~f0(γ):=[γ|f0(γ)|2η~(γ1γ)]1/2.\tilde{\eta}\star f_{0}(\gamma):=\big{[}\sum_{\gamma^{\prime}}|f_{0}(\gamma^{\prime})|^{2}\tilde{\eta}(\gamma^{\prime-1}\gamma)\big{]}^{1/2}.

Whenever f00f_{0}\not\equiv 0, η~f0\tilde{\eta}\star f_{0} has support the whole group Γ\Gamma. By Cauchy-Schwarz,

(26) η~:f0η~f0\star_{\tilde{\eta}}:f_{0}\mapsto\tilde{\eta}\star f_{0}

is a Γ\Gamma-equivariant distance non-increasing map which preserves the 2\ell^{2} norm of functions (essentially this is the discrete version of [BCG95, Remarque 2.7]). Hence η~\star_{\tilde{\eta}} induces a well-defined distance non-increasing map from S/λΓ(Γ){S^{\infty}}/\lambda_{\Gamma}(\Gamma) to itself, homotopic to the identity. Importantly, one checks that η~\star_{\tilde{\eta}} is differentiable on S{S^{\infty}}, and its differential strictly decreases the norm of any tangent vector at any point of S{S^{\infty}}. In particular, η~\star_{\tilde{\eta}} is strictly distance decreasing.

Remark 3.4.

Why is the intrinsic flat topology needed? When defining spherical Plateau solutions, one might hope that they are embedded in the spherical quotient S/λΓ(Γ){S^{\infty}}/\lambda_{\Gamma}(\Gamma), but this is not possible for the following reason. Consider any non-zero integral current C𝒞(h)C\in\mathscr{C}(h) in the spherical quotient S/λΓ(Γ){S^{\infty}}/\lambda_{\Gamma}(\Gamma), where Γ\Gamma is infinite torsion-free and hHn(Γ;)h\in H_{n}(\Gamma;\mathbb{Z}). By applying the spherical convolution η~\star_{\tilde{\eta}} with η~\tilde{\eta} as above, we get a current (η~)(C)(\star_{\tilde{\eta}})_{\sharp}(C) which still belongs to 𝒞(h)\mathscr{C}(h) but with strictly smaller mass by the area formula reviewed in Subsection 1.4. As a consequence, whenever hh is nonzero, no element of 𝒞(h)\mathscr{C}(h) can actually achieve the equality 𝐌(C)=SphereVol(h)\mathbf{M}(C)=\operatorname{SphereVol}(h). More generally, this argument shows that there is no non-trivial mass-minimizing integral current in S/λΓ(Γ){S^{\infty}}/\lambda_{\Gamma}(\Gamma)! This explains why notions such as “intrinsic flat convergence”, “integral current spaces” [SW11] or “ultralimit” are necessary in formulating the spherical Plateau problem.

3.5. Besson-Courtois-Gallot’s definition of the spherical volume

There are several natural ways to define the setup for the spherical Plateau problem. Conjecturally, many of those setups lead to equivalent notions of spherical volume and spherical Plateau solutions, whenever they make sense. In this subsection, we recall the original setup of Besson-Courtois-Gallot in the case of smooth manifolds, which inspired the definition of the spherical Plateau problem.

Let (M,g)(M,g) be a closed, oriented, smooth Riemannian nn-manifold with fundamental group Γ:=π1(M)\Gamma:=\pi_{1}(M), let (M~,g)(\tilde{M},g) be its universal cover. Let DMM~D_{M}\subset\tilde{M} be a Borel fundamental domain and let S2(M~,g)S_{2}(\tilde{M},g) be the unit sphere of L2(M~,g)L^{2}(\tilde{M},g), endowed with the standard metric 𝐠L2(M~,g)\mathbf{g}_{L^{2}(\tilde{M},g)}. There is a natural action λ(M~,g)\lambda_{(\tilde{M},g)} of Γ\Gamma by isometries on S2(M~,g)S_{2}(\tilde{M},g), and changing the metric gg yields Γ\Gamma-equivariantly isometric spaces, see [BCG91, Section 3, I]. Besson-Courtois-Gallot [BCG91, Section 3, I] defined the spherical volume of MM as

SphereVolBCG(M)\displaystyle\operatorname{SphereVol}_{BCG}(M)
:=inf{Vol(DM,ϕ𝐠L2(M~,g)); ϕ:M~S2(M~,g) is a Γ-equivariant immersion}.\displaystyle:=\inf\{\operatorname{Vol}(D_{M},\phi^{*}\mathbf{g}_{L^{2}(\tilde{M},g)});\quad\text{ $\phi:\tilde{M}\to S_{2}(\tilde{M},g)$ is a $\Gamma$-equivariant immersion}\}.

In fact, we could have used S2(Γ)S^{\infty}\subset\ell^{2}(\Gamma), the regular representation λΓ\lambda_{\Gamma}, and Γ\Gamma-equivariant smooth maps, instead of respectively S2(M~,g)L2(M~,g)S_{2}(\tilde{M},g)\subset L^{2}(\tilde{M},g), λ(M~,g)\lambda_{(\tilde{M},g)}, and Γ\Gamma-equivariant immersions, without changing the value of SphereVolBCG(M)\operatorname{SphereVol}_{BCG}(M).

By [BCG91], the simplicial volume [Gro82] of a closed oriented nn-manifold MM is related to the spherical volume of Besson-Courtois-Gallot by

MCnSphereVolBCG(M).\|M\|\leq C_{n}\operatorname{SphereVol}_{BCG}(M).

It is not hard to see that

SphereVolBCG(M)SphereVol(M).\operatorname{SphereVol}_{BCG}(M)\geq\operatorname{SphereVol}(M).

If the fundamental group Γ\Gamma is torsion-free, it is possible to show that equality holds: this non-trivial result is essentially contained in [Bab06, Lemma 3.10][Bru08, Section 2] when the dimension of MM is at least 33.

Question 6 (Equivalence of definitions).

Do we always have

SphereVolBCG(M)=SphereVol(M)?\operatorname{SphereVol}_{BCG}(M)=\operatorname{SphereVol}(M)?

3.6. Intrinsic isomorphism

It is sometimes natural and helpful to consider spherical Plateau solutions up to “intrinsic isomorphism”, and discard some of the non-infinitesimal information. Indeed under such an equivalence relation, uniqueness and rigidity properties emerge naturally, as we will see in Sections 4, 5 and 6. Given a metric space (X,d)(X,d), the intrinsic metric on XX induced by dd is denoted by LdL_{d} [BBI22, Chapter 2, Section 2.3]. By convention, the LdL_{d}-distance between points of different path connected components of (X,d)(X,d) is \infty. Note that the identity map

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

is always 11-Lipschitz.

Consider an integral current space C=(X,d,T)C=(X,d,T) and an oriented complete finite volume Riemannian manifold (N,gN)(N,g_{N}) which is not necessarily connected, which induces the integral current space (N,distgN,1N)(N,\operatorname{dist}_{g_{N}},\llbracket 1_{N}\rrbracket).

Definition 3.5.

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.

One could also try to formulate more general definitions of the type “Two integral current spaces are intrinsically isomorphic if…”. For clarity, we emphasize that “intrinsically isomorphic” does not imply “at intrinsic flat distance 0 from each other”.

4. Spherical Plateau solutions for hyperbolic manifolds

In this section, we review how Besson-Courtois-Gallot compute the spherical volume of hyperbolic manifolds and how it is used in their proof of the volume entropy inequality. Then we outline a proof of our first main theorem, the uniqueness of spherical Plateau solutions for hyperbolic manifolds, up to intrinsic isomorphism. This result in turn leads to an area rigidity property for the regular representation of fundamental groups of hyperbolic manifolds. All those discussions can be adapted to the general rank one locally symmetric case.

4.1. The spherical volume of hyperbolic manifolds

The spherical volume of closed oriented hyperbolic manifolds was computed by Besson-Courtois-Gallot. For completeness, we outline their proof in our setting when the dimension is at least 33.

Theorem 4.1.

[BCG95, BCG96] Let (M,g0)(M,g_{0}) be a closed oriented hyperbolic manifold of dimension at least 33. Then

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

Let us sketch the proof, which is due to Besson-Courtois-Gallot [BCG95, Sections 5 and 6]. Suppose that the dimension of MM is at least 33, let Γ:=π1(M)\Gamma:=\pi_{1}(M) and let hMHn(Γ;)h_{M}\in H_{n}(\Gamma;\mathbb{Z}) be the induced homology class.

Let C𝒞(hM)C\in\mathscr{C}(h_{M}) be a cycle in S/λΓ(Γ){S^{\infty}}/\lambda_{\Gamma}(\Gamma) representing hMh_{M}. It is convenient to consider cycles which are polyhedral chains. Recall from Lemma 1.6 that CC can be nicely approximated by a polyhedral chain P𝒞(hM)P\in\mathscr{C}(h_{M}). Moreover, by a further perturbation if necessary, we get another polyhedral chain P𝒞(hM)P^{\prime}\in\mathscr{C}(h_{M}) so that spt(P)𝕊+/λΓ(Γ)\operatorname{spt}(P^{\prime})\subset\mathbb{S}^{+}/\lambda_{\Gamma}(\Gamma) where 𝕊+\mathbb{S}^{+} is the set of functions with finite support (see Section 2). Afterwards, we will assume without loss of generality that CC is such a polyhedral chain.

Fix a point oM~o\in\tilde{M} in the universal cover of MM, and let 𝐩M\mathbf{p}\in M be its projection in MM. Let

Bar:𝕊+/λΓ(Γ)M\mathrm{Bar}:\mathbb{S}^{+}/\lambda_{\Gamma}(\Gamma)\to M

be the barycenter map whose definition is given in Section 2. Recall that the restriction of Bar\mathrm{Bar} to CC is a Lipschitz map. By Γ\Gamma-equivariance of Bar:𝕊+M~\mathrm{Bar}:\mathbb{S}^{+}\to\tilde{M},

(28) Bar(C)=1M.\mathrm{Bar}_{\sharp}(C)=\llbracket 1_{M}\rrbracket.

For almost every point qspt(C)q\in\operatorname{spt}(C), CC admits a tangent nn-plane at qq. The nn-dimensional Jacobian of Bar{\mathrm{Bar}} along the tangent nn-plane is well-defined and is bounded from above by (4n(n1)2)n/2\big{(}\frac{4n}{(n-1)^{2}}\big{)}^{n/2} thanks to (16) in Lemma 2.4. This implies by (28) and the area formula in Subsection 1.4 that

𝐌(C)((n1)24n)n/2Vol(M,g0).\mathbf{M}(C)\geq\big{(}\frac{(n-1)^{2}}{{4n}}\big{)}^{n/2}\operatorname{Vol}(M,g_{0}).

Since C𝒞(hM)C\in\mathscr{C}(h_{M}) has mass arbitrary close to SphereVol(M)\operatorname{SphereVol}(M), we conclude that

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

The inverse inequality

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

directly follows from a general inequality between the spherical volume and the volume entropy of a closed Riemannian manifold, see [BCG91, Corollary 3.13]. It states that if (M,g)(M,g) is a closed oriented Riemannian nn-manifold, then

(29) SphereVol(M)Vol(M,h(g)24ng)\operatorname{SphereVol}(M)\leq\operatorname{Vol}(M,\frac{h(g)^{2}}{4n}g)

where h(g)h(g) is the volume entropy of the Riemannian metric gg (the definition is recalled in (31)). Note that in [BCG91, Section 3, I], SphereVol(M)\operatorname{SphereVol}(M) is denoted T(M)T(M). The proof of (29) is based on the following maps. Denote by DMD_{M} a Borel fundamental domain in the universal cover 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. For c>h(g)c>h(g), defined the equivariant map

(30) 𝒫c:M~S2(Γ)x{γ1ec2distg(x,.)L2(M~,g)[γ.DMecdistg(x,u)dvolg(u)]1/2}.\displaystyle\begin{split}\mathcal{P}_{c}&:\tilde{M}\to{S^{\infty}}\subset\ell^{2}(\Gamma)\\ &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}\}.\end{split}

Properties of 𝒫c\mathcal{P}_{c} [BCG91, Proof of Lemma 3.1] [Son23, Lemma 3.1] imply that

lim infch(g)𝐌((𝒫c)1M)h(g)n2nnn/2Vol(M,g).\liminf_{c\to h(g)}\mathbf{M}((\mathcal{P}_{c})_{\sharp}\llbracket 1_{M}\rrbracket)\leq\frac{h(g)^{n}}{2^{n}n^{n/2}}\operatorname{Vol}(M,g).

Since by equivariance, (𝒫c)1M(\mathcal{P}_{c})_{\sharp}\llbracket 1_{M}\rrbracket belongs to 𝒞(hM)\mathscr{C}(h_{M}), the above inequality shows (29).

The analogue of Theorem 4.1 for closed oriented surfaces was shown in [BCG91, Proposition 3.9] by methods specific to the 22-dimensional case. Theorem 4.1 also extends directly to rank one (i.e. negatively curved) locally symmetric manifolds [BCG95, BCG96, Rua22]. As for higher rank locally symmetric manifolds, following the logic of Besson-Courtois-Gallot [BCG95] (see Subsection 4.2), a big step towards the entropy rigidity conjecture in higher rank [BCG96, Question (5)][CF02] would be the computation of their spherical volumes:

Question 7 (Spherical volume in higher rank).

What is the spherical volume of a closed oriented locally symmetric manifold of higher rank?

4.2. From spherical volume to volume entropy and back again

Given a closed manifold MM, if gg is a Riemannian metric on MM, let h(g)h(g) denote its volume entropy:

(31) h(g):=limRlogVol(B~g(o,R))Rh(g):=\lim_{R\to\infty}\frac{\log\operatorname{Vol}(\tilde{B}_{g}(o,R))}{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 for hyperbolic manifolds [BCG95, BCG96] states that if (M,g0)(M,g_{0}) is a hyperbolic manifold of dimension n2n\geq 2, and if gg is any other metric on MM, then

h(g)nVol(M,g)h(g0)nVol(M,g0).h(g)^{n}\operatorname{Vol}(M,g)\geq h(g_{0})^{n}\operatorname{Vol}(M,g_{0}).

Equivalently, if gg is normalized to have volume entropy h(g0)h(g_{0}), then

(32) Vol(M,g)Vol(M,g0).\operatorname{Vol}(M,g)\geq\operatorname{Vol}(M,g_{0}).

Moreover when n3n\geq 3, equality holds exactly when gg is isometric to g0g_{0}. A slight extension of this theorem implies the classical Mostow’s rigidity theorem [BCG95].

Nowadays, the proof of Besson-Courtois-Gallot [BCG95, BCG96] in the case n3n\geq 3 is primarily remembered for their barycenter map. Yet, a pivotal insight in the original paper [BCG95] is that determining the minimal volume entropy of hyperbolic manifolds can be reduced to the computation of their spherical volume. This aspect of their work is rooted in the theory of minimal surfaces and calibrations, and should be recalled. We can summarize their strategy as follows. Let MM be as before and let Γ:=π1(M)\Gamma:=\pi_{1}(M). If (M,g)(M,g) is normalized so that h(g)=h(g0)=n1h(g)=h(g_{0})=n-1, then by using properties of the maps 𝒫c\mathcal{P}_{c} defined in (30) for c>h(g0)c>h(g_{0}) [BCG91, Proof of Lemma 3.1] [Son23, Lemma 3.1], the image 𝒫c(M)\mathcal{P}_{c}(M) inside S/λΓ(Γ){S^{\infty}}/\lambda_{\Gamma}(\Gamma) has volume at most cn2nnn/2Vol(M,g)\frac{c^{n}}{2^{n}n^{n/2}}\operatorname{Vol}(M,g). But since 𝒫c(M)\mathcal{P}_{c}(M) determines a cycle in 𝒞(hM)\mathscr{C}(h_{M}), by Theorem 4.1, the volume of 𝒫c(M)\mathcal{P}_{c}(M) is at least the spherical volume (n1)n2nnn/2Vol(M,g0)\frac{(n-1)^{n}}{2^{n}n^{n/2}}\operatorname{Vol}(M,g_{0}). As cc can be taken arbitrarily close to n1n-1, we find that (32) is true. The rigidity part is shown as follows: if equality holds, then as cn1c\to n-1, the composition Bar𝒫c\mathrm{Bar}\circ\mathcal{P}_{c} is almost a Riemannian isometry and converges to a Riemannian isometry, so in the limit we conclude that gg is isometric to g0g_{0}.

Actually, the summary above deviates a bit from the original presentation of Besson-Courtois-Gallot: in [BCG95], the authors use a spherical quotient different from the quotient S/λΓ(Γ){S^{\infty}}/\lambda_{\Gamma}(\Gamma) that we choose to work with, and they use an explicit minimal isometric embedding of (M,(n1)24ng0)(M,\frac{(n-1)^{2}}{4n}g_{0}) into that spherical quotient. Let us describe those geometric objects. Consider the universal cover (M~,g0)(\tilde{M},g_{0}) with the hyperbolic metric. Fix a basepoint 𝐨M~\mathbf{o}\in\tilde{M}, let M~\partial\tilde{M} be the boundary at infinity of M~\tilde{M} with the standard probability measure determined by 𝐨\mathbf{o}, and let S2(M~)S_{2}(\partial\tilde{M}) be the unit sphere in L2(M~)L^{2}(\partial\tilde{M}). For any θM~\theta\in\partial\tilde{M}, the corresponding Busemann function is defined for any xM~x\in\tilde{M} as

Bθ(x):=limt(distg0(y,c(t))t)B_{\theta}(x):=\lim_{t\to\infty}(\operatorname{dist}_{g_{0}}(y,c(t))-t)

where c:[0,)c:[0,\infty) is the half-geodesic starting at 𝐨\mathbf{o}, and converging to θ\theta. The group Γ\Gamma acts naturally on M~\tilde{M} and M~\partial\tilde{M}. There is also a natural proper, free, isometric action ρB\rho_{B} of Γ\Gamma on S2(M~)S_{2}(\partial\tilde{M}) given, for all γΓ\gamma\in\Gamma, fS2(M~)f\in S_{2}(\partial\tilde{M}), by

(33) ρB(γ).f(θ)=f(γ1(θ))en12Bθ(γ(𝐨)),\rho_{B}(\gamma).f(\theta)=f(\gamma^{-1}(\theta))e^{-\frac{n-1}{2}B_{\theta}(\gamma(\mathbf{o}))},

see [BCG95, Lemme 2.2]. This action ρB\rho_{B} is also called a boundary representation of Γ\Gamma. It is possible to show that the spherical quotient S2(M~)/ρB(Γ)S_{2}(\partial\tilde{M})/\rho_{B}(\Gamma) is not isometric to our spherical quotient S/λΓ(Γ){S^{\infty}}/\lambda_{\Gamma}(\Gamma) defined with the regular representation of Γ\Gamma. Nevertheless, these two spaces are closely related, as S2(M~)/ρB(Γ)S_{2}(\partial\tilde{M})/\rho_{B}(\Gamma) is contained in the “ultralimit of S/λΓ(Γ){S^{\infty}}/\lambda_{\Gamma}(\Gamma)”.

Consider the following embedding

(34) 𝒫:(M~,(n1)24ng0)S2(M~)𝒫(x):={θen12Bθ(x)}.\begin{split}\mathscr{P}:(\tilde{M},\frac{(n-1)^{2}}{4n}g_{0})\to S_{2}(\partial\tilde{M})\\ \mathscr{P}(x):=\{\theta\mapsto e^{-\frac{n-1}{2}B_{\theta}(x)}\}.\end{split}

It is not too hard to check that 𝒫\mathscr{P} is an isometric and minimal embedding. As explained in [BCG95, Section 2], this map is equivariant, and so after taking the quotient by Γ\Gamma and rescaling the metric g0g_{0}, we get a map

𝒫:(M,(n1)24ng0)S2(M~)/ρB(Γ).\mathscr{P}:(M,\frac{(n-1)^{2}}{4n}g_{0})\to S_{2}(\partial\tilde{M})/\rho_{B}(\Gamma).

Again, this 𝒫\mathscr{P} is an isometric and minimal embedding. Strikingly, Besson-Courtois-Gallot discovered [BCG95, Proposition 5.7, Section 6] that 𝒫(M)\mathscr{P}(M) is an area-minimizing nn-submanifold of the Hilbert Riemannian manifold S2(M~)/ρB(Γ)S_{2}(\partial\tilde{M})/\rho_{B}(\Gamma), by arguing that 𝒫(M)\mathscr{P}(M) is in fact a calibrated submanifold. When n3n\geq 3, the calibration is essentially constructed as the pull-back of the volume form on (M,g0)(M,g_{0}) by a barycenter map

Bar:S2(M~)/ρB(Γ)(M,g0).\mathrm{Bar}:S_{2}(\partial\tilde{M})/\rho_{B}(\Gamma)\to(M,g_{0}).

Now, here is the relevance of 𝒫(M)\mathscr{P}(M) and our discussion about the spherical Plateau problem. If g=g0g=g_{0} then the image 𝒫c(M)\mathcal{P}_{c}(M) in S/λΓ(Γ){S^{\infty}}/\lambda_{\Gamma}(\Gamma) has volume converging to the spherical volume (n1)n2nnn/2Vol(M,g0)\frac{(n-1)^{n}}{2^{n}n^{n/2}}\operatorname{Vol}(M,g_{0}) as cn1c\to n-1. It can be shown that, as integral currents, 𝒫c(M)\mathcal{P}_{c}(M) converges in the intrinsic flat topology to 𝒫(M)\mathscr{P}(M). With the vocabulary of Section 3, this means that there is an explicit minimizing sequence of cycles in 𝒞(hM)\mathscr{C}(h_{M}) inside S/λΓ(Γ){S^{\infty}}/\lambda_{\Gamma}(\Gamma) converging to a spherical Plateau solution which is given by 𝒫(M)\mathscr{P}(M). Is 𝒫(M)\mathscr{P}(M) the only spherical Plateau solution up to isomorphism when n3n\geq 3 (see Question 8)? We will show in Theorem 4.2 that it is the unique one at least up to instrinsic isomorphism.

4.3. Intrinsic uniqueness and rigidity of spherical Plateau solutions

We now come to our first new result, which states that spherical Plateau solutions are unique for hyperbolic manifolds, up to intrinsic isomorphism (see Definition 3.5):

Theorem 4.2.

Let (M,g0)(M,g_{0}) be 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}).

Outline of proof.

Let Γ:=π1(M)\Gamma:=\pi_{1}(M), and let hMHn(Γ;)h_{M}\in H_{n}(\Gamma;\mathbb{Z}) be the fundamental homology class, where n3n\geq 3 is the dimension of MM. Let Ci𝒞(hM)C_{i}\in\mathscr{C}(h_{M}) be a minimizing sequence, namely

(35) lim𝐌(Ci)=SphereVol(M)=Vol(M,(n1)24ng0).\lim\mathbf{M}(C_{i})=\operatorname{SphereVol}(M)=\operatorname{Vol}(M,\frac{(n-1)^{2}}{4n}g_{0}).

The second equality above is Theorem 27. By Lemma 1.6, as in the proof of Theorem 27, we can assume CiC_{i} to be polyhedral chains, with support in 𝕊+/λΓ(Γ)\mathbb{S}^{+}/\lambda_{\Gamma}(\Gamma), where 𝕊+\mathbb{S}^{+} is the set of functions with finite support.

We assume that CiC_{i} converges to a spherical Plateau solution

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

We use the notation

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

Jacobians, lengths and distances will be computed with respect to gg^{\prime}. Fix oM~o\in\tilde{M} a point in the universal cover. Let

Bar:spt(Ci)𝕊+/λΓ(Γ)M\mathrm{Bar}:\operatorname{spt}(C_{i})\subset\mathbb{S}^{+}/\lambda_{\Gamma}(\Gamma)\to M

be the barycenter map defined in Section 2.

The Jacobian bound (16) on the barycenter map in Lemma 2.4, (28) and (35) imply that, as ii\to\infty, on a larger and larger region of the polyhedral chain spt(Ci)\operatorname{spt}(C_{i}), the Jacobian bound is almost tight. In particular there is a “good region” Ωispt(Ci)\Omega_{i}\subset\operatorname{spt}(C_{i}) such that the mass of CiΩiC_{i}\llcorner\Omega_{i} converges to SphereVol(M)\operatorname{SphereVol}(M), the Jacobian of Bar\mathrm{Bar} converges to 11 on Ωi\Omega_{i} (with respect to the metric gg^{\prime}), and Bar\mathrm{Bar} is injective on Ωi\Omega_{i}.

The barycenter map Bar:spt(Ci)M\mathrm{Bar}:\operatorname{spt}(C_{i})\to M is not uniformly Lipschitz as ii\to\infty. Nevertheless, by Lemma 2.4, it is indeed uniformly Lipschitz in small neighborhoods of Ωi\Omega_{i}. Besides, it can be shown using Lemma 2.4 that the barycenter map is almost 11-Lispchitz in small balls near Ωi\Omega_{i}.

Take DiD_{i} the restriction of CiC_{i} to an r~\tilde{r}-neighborhood of the good region Ωispt(Ci)\Omega_{i}\subset\operatorname{spt}(C_{i}), for some small fixed r~>0\tilde{r}>0. By construction, DiD_{i} still converges in the intrinsic flat topology to CC_{\infty}. The upshot is that the mass of the boundary Di\partial D_{i} converges to 0, the barycenter map restricted to spt(Di)\operatorname{spt}(D_{i}) has a uniform Lipschitz bound and is almost 11-Lispchitz in small balls near Ωi\Omega_{i}. Those properties will be important in the next steps.

By the previous paragraph, one can embed the sequence of integral currents DiD_{i} in a Banach space where they converge in the flat topology to C=(X,d,S)C_{\infty}=(X_{\infty},d_{\infty},S_{\infty}) viewed as an integral current in that Banach space. The barycenter map Bar\mathrm{Bar} is well-defined on each spt(Di)\operatorname{spt}(D_{i}) so by an Arzelá-Ascoli type argument and the uniform Lipschitz control [Sor18, Theorem 6.1], there is a natural Lipschitz “limit barycenter map”

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

where the support sptS\operatorname{spt}S_{\infty} coincides with the completion of (X,d)(X_{\infty},d_{\infty}). Moreover, by slightly extending [BCS23, Lemma 7.3], it can be shown that

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

By Lemma 2.4, when the Jacobian bound for the barycenter map is almost saturated, the differential of Bar\mathrm{Bar} is almost a linear isometry. Since the Jacobian bound for Bar\mathrm{Bar} indeed becomes arbitrarily close to the sharp upper bound on Ωispt(Di)\Omega_{i}\subset\operatorname{spt}(D_{i}) as ii\to\infty, Bar\mathrm{Bar} is almost a Riemannian isometry for large ii. So intuitively, we expect that the differential of the limit barycenter map Bar\mathrm{Bar}_{\infty} is exactly a linear isometry at any point of sptS\operatorname{spt}S_{\infty}, namely that Bar\mathrm{Bar}_{\infty} is a Riemannian isometry. This would essentially finish the proof. That sketch would work if the convergence of CiC_{i} to CC_{\infty} was known to be smooth (at least outside of a small singular set) [BBCG12], or if the limit map was known to be 11-Lipschitz (see for instance for such Lipschitz-volume rigidity results [BCG95, Proposition C.1], [BBCG12, Sections 3, 4, 5], [DP23, Theorem 1.1], [BCS23, Theorem 1.1] and [Züs23, Theorem 1.2]). However in our situation C=(X,d,S)C_{\infty}=(X_{\infty},d_{\infty},S_{\infty}) is a priori just an integral current to which CiC_{i} converges weakly, and the limit map is never 11-Lipschitz in our case (even though it will be shown a posteriori to be 11-Lipschitz with respect to the intrinsic metrics).

In order to show that

Bar:sptS(M,g)\mathrm{Bar}_{\infty}:\operatorname{spt}S_{\infty}\to(M,g^{\prime})

is an isometry for the respective intrinsic metrics, we work with the maps Bar:spt(Di)(M,g)\mathrm{Bar}:\operatorname{spt}(D_{i})\to(M,g^{\prime}) instead of directly with the limit map. We prove a general result [Son23, Proposition 1.4] which roughly says that limits of maps φi\varphi_{i} which are uniformly Lipschitz, almost Riemannian isometries, and almost 11-Lipschitz in small balls, are indeed Riemannian isometries. This general result can be directly applied to the sequence Bar:spt(Di)(M,g)\mathrm{Bar}:\operatorname{spt}(D_{i})\to(M,g^{\prime}). The point of this result is that while the limit map Bar\mathrm{Bar}_{\infty} is constructed using the extrinsic metrics on spt(Di)\operatorname{spt}(D_{i}), the intrinsic information about Bar\mathrm{Bar} (namely the fact that it is almost a Riemannian isometry) still passes to the limit.

Let us make some comments on the proof of [Son23, Proposition 1.4] in our specific case. Since Bar\mathrm{Bar} is almost 11-Lipschitz in small balls near Ωi\Omega_{i}, it is simple to check that Bar\mathrm{Bar}_{\infty} is 11-Lipschitz for the intrinsic metric LdL_{d_{\infty}} on sptS\operatorname{spt}S_{\infty}. Note that we cannot apply Lipschitz-volume rigidity results like [Züs23, Theorem 1.2], because we do not know if a priori there is an integral current TT on the completion of (sptS,L)(\operatorname{spt}S_{\infty},L_{\infty}) such that (Bar)T=1M(\mathrm{Bar}_{\infty})_{\sharp}T=\llbracket 1_{M}\rrbracket, and if volumes for the new metric are preserved. Instead, we argue directly as follows. We want to show that conversely Bar\mathrm{Bar}_{\infty} does not decrease distances for the intrinsic metrics: given x,ysptSx,y\in\operatorname{spt}S_{\infty}, and a geodesic segment σ\sigma between Bar(x)\mathrm{Bar}_{\infty}(x) and Bar(y)\mathrm{Bar}_{\infty}(y) in (M,g)(M,g^{\prime}), we want to lift σ\sigma to a segment of same length in sptS\operatorname{spt}S_{\infty}. Technically, the main tool is the coarea formula reviewed in Subsections 1.4 and Sard’s lemma. As an application, for each ii, we can perturb a bit σ\sigma to a nearby curve σi\sigma_{i} such that the preimage

ϰi:=Bar1(σi)spt(Di)\varkappa_{i}:=\mathrm{Bar}^{-1}(\sigma_{i})\cap\operatorname{spt}(D_{i})

is a rectifiable curve contained inside spt(Di)\operatorname{spt}(D_{i}) of length converging to lengthg(σ)\operatorname{length}_{g^{\prime}}(\sigma) as ii\to\infty. Heuristically, this is because the coarea formula ensures that most of ϰi\varkappa_{i} is contained in the good region Ωi\Omega_{i}, where the differential of Bar\mathrm{Bar} is almost a linear isometry. Moreover ϰi\varkappa_{i} is the union of a segment with two endpoints and some closed curves of small lengths, and the segment component converges as ii\to\infty to a rectifiable segment inside the support sptS\operatorname{spt}S_{\infty} with endpoints xx and yy, and length equal to lengthg(σ)\operatorname{length}_{g^{\prime}}(\sigma).

After proving that Bar:sptS(M,g)\mathrm{Bar}_{\infty}:\operatorname{spt}S_{\infty}\to(M,g^{\prime}) is an isometry for the respective intrinsic metrics, we can conclude the proof using (36) and the inverse map φ:=(Bar)1\varphi:=(\mathrm{Bar}_{\infty})^{-1} from (M,distg)(M,\operatorname{dist}_{g^{\prime}}) to sptS\operatorname{spt}S_{\infty} in Definition 3.5.

From Theorem 4.2, we deduce the following area rigidity result. Let (M,g0)(M,g_{0}) be a closed oriented hyperbolic manifold of dimension n3n\geq 3 and set Γ:=π1(M)\Gamma:=\pi_{1}(M). A fundamental family of representations of Γ\Gamma is the set of orthogonal representations ρ:ΓEnd(H)\rho:\Gamma\to\operatorname{End}(H) which are weakly contained in the regular representation λΓ:ΓEnd(2(Γ))\lambda_{\Gamma}:\Gamma\to\operatorname{End}(\ell^{2}(\Gamma)), see [BdLHV08, Definition F.1.1]. We say444Usually, one considers complex Hilbert spaces and unitary representations but in this paper we will only consider real Hilbert spaces and orthogonal representations. that an orthogonal representation ρ:ΓEnd(H)\rho:\Gamma\to\operatorname{End}(H) is weakly contained in λΓ\lambda_{\Gamma} if for every ξH\xi\in H, every finite subset QQ of Γ\Gamma, and every ϵ>0\epsilon>0, there exist η1,,ηm\eta_{1},...,\eta_{m} in 2(Γ)\ell^{2}(\Gamma) such that for all gQg\in Q,

|ρ(g)ξ,ξj=1mλΓ(g)ηj,ηj|<ϵ.\big{|}\langle\rho(g)\xi,\xi\rangle-\sum_{j=1}^{m}\langle\lambda_{\Gamma}(g)\eta_{j},\eta_{j}\rangle\big{|}<\epsilon.

Now let M~\tilde{M} be the universal cover of MM, on which Γ\Gamma acts properly freely, let g0g_{0} be the lifted hyperbolic metric on M~\tilde{M}, and let DMD_{M} be a Borel fundamental domain in M~\tilde{M}.

Corollary 4.3.

Let SHS_{H} be the unit sphere in a Hilbert space (H,𝐠H)(H,\mathbf{g}_{H}) and let ρ:ΓEnd(H)\rho:\Gamma\to\operatorname{End}(H) be an orthogonal representation weakly contained in λΓ\lambda_{\Gamma}. Consider a smooth Γ\Gamma-equivariant map

f:M~SH.f:\tilde{M}\to S_{H}.

Then

Vol(DM,f𝐠H)SphereVol(M),\operatorname{Vol}(D_{M},f^{*}\mathbf{g}_{H})\geq\operatorname{SphereVol}(M),

with equality if and only if (f(M~),𝐠H)(f(\tilde{M}),\mathbf{g}_{H}) is an embedded nn-plane in SHS_{H}, Riemannian isometric to (M~,(n1)24ng0)(\tilde{M},\frac{(n-1)^{2}}{4n}g_{0}).

Outline of proof.

Consider the infinite direct sum of the regular representation, denoted by

λΓ:ΓEnd(2(Γ)).\bigoplus^{\infty}\lambda_{\Gamma}:\Gamma\to\operatorname{End}(\bigoplus^{\infty}\ell^{2}(\Gamma)).

Let (S,𝐠S)(S^{\prime},\mathbf{g}_{S^{\prime}}) be the unit sphere of 2(Γ)\bigoplus^{\infty}\ell^{2}(\Gamma) with its standard metric. The natural map

𝒜:(a1,a2,)S[i1ai2]1/2S2(Γ)\mathcal{A}^{\prime}:(a_{1},a_{2},...)\in S^{\prime}\mapsto[\sum_{i\geq 1}a_{i}^{2}]^{1/2}\in S^{\infty}\subset\ell^{2}(\Gamma)

is Γ\Gamma-equivariant and 11-Lipschitz by the argument of Subsection 3.4.2.

The fact that ρ\rho is weakly contained in λΓ\lambda_{\Gamma} means that the representation ρ\rho can be arbitrarily well “approximated” by λΓ\bigoplus^{\infty}\lambda_{\Gamma}. Hence for any ϵ>0\epsilon>0, there exists a smooth Γ\Gamma-equivariant map

fϵ:M~Sf_{\epsilon}:\tilde{M}\to S^{\prime}

which is ϵ\epsilon-close to ff in the C1C^{1}-topology, in the sense that

(37) fϵ𝐠Sf𝐠HC0(DM)ϵ.\|f_{\epsilon}^{*}\mathbf{g}_{S^{\prime}}-f^{*}\mathbf{g}_{H}\|_{C^{0}(D_{M})}\leq\epsilon.

Consider the Lipschitz Γ\Gamma-equivariant map

𝒜fϵ:M~S2(Γ)\mathcal{A}^{\prime}\circ f_{\epsilon}:\tilde{M}\to S^{\infty}\subset\ell^{2}(\Gamma)

which after taking the quotient by λΓ(Γ)\lambda_{\Gamma}(\Gamma), gives a map

MS/λΓ(Γ).M\to S^{\infty}/\lambda_{\Gamma}(\Gamma).

The image of this map can be identified with a cycle C𝒞(hM)C\in\mathscr{C}(h_{M}) whose mass satisfies, by (37) and the 11-Lipschitzness of 𝒜\mathcal{A}^{\prime}:

(38) 𝐌(C)Vol(DM,f𝐠H)+cϵ\mathbf{M}(C)\leq\operatorname{Vol}(D_{M},f^{*}\mathbf{g}_{H})+c_{\epsilon}

for some cϵc_{\epsilon} converging to 0 as ϵ0\epsilon\to 0. This establishes the inequality of the statement by definition of SphereVol(M)\operatorname{SphereVol}(M).

As ϵ\epsilon\to\infty, the cycle CC above converges by construction of fϵf_{\epsilon} to the quotient f(M~)/ρ(Γ)f(\tilde{M})/\rho(\Gamma) (viewed as an integral current space). Suppose that equality holds in the statement, namely

Vol(DM,f𝐠H)=SphereVol(M).\operatorname{Vol}(D_{M},f^{*}\mathbf{g}_{H})=\operatorname{SphereVol}(M).

Then by (38), f(M~)/ρ(Γ)f(\tilde{M})/\rho(\Gamma) is a spherical Plateau solution, so by the uniqueness result, Theorem 4.2, we conclude that (f(M~),𝐠H)(f(\tilde{M}),\mathbf{g}_{H}) is an embedded nn-plane in SHS_{H}, Riemannian isometric to (M~,(n1)24ng0)(\tilde{M},\frac{(n-1)^{2}}{4n}g_{0}) as desired. ∎

Let us emphasize that, by Subsection 4.2, Corollary 4.3 implies the entropy inequality and rigidity theorem of Besson-Courtois-Gallot [BCG95]. We may in fact hope for a much stronger “extrinsic” area rigidity in the equality case of Corollary 4.3: ρ\rho should contain as a subrepresentation the boundary representation ρB\rho_{B} defined in (33), and f(M~)f(\tilde{M}) should be equal to the special minimal nn-plane 𝒫(M~)\mathscr{P}(\tilde{M}) defined in (34) up to an isometry of the ambient unit sphere. An analogous strong area rigidity property should hold for hyperbolic surfaces. Those conjectures are closely related to Questions 8 and 9 below.

Details for the proof of Theorem 4.2 appear in [Son23], where for simplicity we use a slightly different but equivalent definition of spherical Plateau solutions. All the arguments above extend to the locally symmetric case of rank one, by [BCG95, BCG96, Rua22], and so we obtain:

Theorem 4.4.

If (M,g0)(M,g_{0}) is a closed oriented locally symmetric manifold of dimension at least 33, with negative curvature between 4-4 and 1-1, then any spherical Plateau solution for MM is intrinsically isomorphic to (M,h(g0)24ng0)(M,\frac{h(g_{0})^{2}}{4n}g_{0}).

The spherical Plateau solutions in Theorem 4.4 are probably unique, i.e. up to isomorphism and not just up to intrinsic isomorphism:

Question 8 (Uniqueness).

If (M,g0)(M,g_{0}) is a closed oriented locally symmetric manifold of rank one of dimension at least 33 , is there a unique spherical Plateau solution for MM?

For closed oriented surfaces Σ\Sigma of genus at least 22, spherical Plateau solutions are non-unique. Indeed each hyperbolic metric on Σ\Sigma, after being rescaled by 18\frac{1}{8}, is intrinsically isomorphic to a spherical Plateau solution. Conjecturally, this is essentially the only source of non-uniqueness. While the barycenter map is not useful for surfaces, the calibration constructed in [BCG95, Section 6] hints at a general classification.

Question 9 (Classification for surfaces).

Can one classify the spherical Plateau solutions for closed oriented surfaces?

5. Spherical Plateau solutions for 3-dimensional manifolds

5.1. The spherical volume of 33-manifolds

By the Geometrization theorem [KL08, MT14], any closed oriented 33-manifold MM is decomposable into a connected sum of irreducible 33-manifolds Y1,,YmY_{1},...,Y_{m}, each of which is divided into a hyperbolic part and a graph manifold (each part could be empty). The disjoint union of the hyperbolic pieces is called the hyperbolic part of MM and denoted by MhypM_{\mathrm{hyp}}. Its complete finite volume hyperbolic metric is called ghypg_{\mathrm{hyp}}. The hyperbolic part (Mhyp,ghyp)(M_{\mathrm{hyp}},g_{\mathrm{hyp}}) is up to isometry canonically determined by MM. We start with the analogue of Theorem 27:

Theorem 5.1.

Let MM be a closed oriented 33-manifold and let (Mhyp,ghyp)(M_{\mathrm{hyp}},g_{\mathrm{hyp}}) be its hyperbolic part with its hyperbolic metric. Then

SphereVol(M)=Vol(Mhyp,13ghyp).\operatorname{SphereVol}(M)=\operatorname{Vol}(M_{\mathrm{hyp}},\frac{1}{3}g_{\mathrm{hyp}}).
Outline of proof.

The proof results from a combination of ideas due to Besson-Courtois-Gallot, Souto and Pieroni [BCG91, BCG95, Sou01, Pie19]. Let MM be a closed oriented 33-manifold. By the Geometrization theorem, it is known that MM is the connected sum of irreducible manifolds Y1,,YmY_{1},...,Y_{m} such that each YjY_{j} is decomposed, after cutting along a disjoint collection of essential tori, into hyperbolic pieces and Seifert pieces (such pieces could be empty). Let Y1,,YkY_{1},...,Y_{k} (k{1,,m}k\in\{1,...,m\}) be the summands with a non-empty union of hyperbolic JSJ components called HjYjH_{j}\subset Y_{j}. The manifold HjH_{j} is not necessarily connected and if non-empty, YjHjY_{j}\setminus H_{j} is a graph manifold with boundary. Let gj,hypg_{j,\mathrm{hyp}} be the complete finite volume metric on HjH_{j}. The disjoint union of HjH_{j} is the hyperbolic part of MM and is denoted by MhypM_{\mathrm{hyp}}.

In [Pie19, Theorem 3.1, Theorem 3.13], for any small δ^>0\hat{\delta}>0, a nonpositively curved metric gj,δ^g_{j,\hat{\delta}} approximating the hyperbolic metric ghypg_{\mathrm{hyp}} is constructed on YjY_{j} for j{1,,k}j\in\{1,...,k\}: there is in particular a region jYj\mathcal{R}_{j}\subset Y_{j} such that gj,δ^g_{j,\hat{\delta}} is a hyperbolic metric on the 1δ^\frac{1}{\hat{\delta}}-neighborhood of j\mathcal{R}_{j} in YjY_{j} with respect to gj,δ^g_{j,\hat{\delta}}, and

(39) Vol(Hj,gj,hyp)Vol(j,gj,δ^)δ^.\operatorname{Vol}(H_{j},g_{j,\mathrm{hyp}})-\operatorname{Vol}(\mathcal{R}_{j},g_{j,\hat{\delta}})\leq\hat{\delta}.

Accordingly, a metric space (X,𝐝δ^)(X,\mathbf{d}_{\hat{\delta}}) obtained by attaching Y1,,YkY_{1},...,Y_{k} at a common point 𝐩\mathbf{p} and collapsing the other summands Yk+1,,YmY_{k+1},...,Y_{m} to 𝐩\mathbf{p} is defined in [Pie19, Section 6]. By a slight abuse of notations, we consider Y1,,YkY_{1},...,Y_{k} as subsets of XX. Set

Γ:=π1(X).\Gamma:=\pi_{1}(X).

Let X~\tilde{X} be the universal cover, endowed with the induced distance denoted by 𝐝~δ^\tilde{\mathbf{d}}_{\hat{\delta}} and let oX~o\in\tilde{X} be a reference point projecting to 𝐩X\mathbf{p}\in X. Except at countably many points of X~\tilde{X} corresponding to the lifts of 𝐩\mathbf{p}, 𝐝~δ^\tilde{\mathbf{d}}_{\hat{\delta}} is in fact given by the path metric of a smooth Riemannian metric, and moreover (X~,𝐝~δ^)(\tilde{X},\tilde{\mathbf{d}}_{\hat{\delta}}) is a CAT(0)\mathrm{CAT}(0) space.

There is a variant of the barycenter map corresponding to Γ\Gamma, (X~,𝐝~δ^)(\tilde{X},\tilde{\mathbf{d}}_{\hat{\delta}}) with properties completely similar to those stated in Section 2:

Bar:𝕊+/λΓ(Γ)(X,𝐝δ^).\mathrm{Bar}:\mathbb{S}^{+}/\lambda_{\Gamma}(\Gamma)\to({X},{\mathbf{d}}_{\hat{\delta}}).

In particular this map enjoys a Jacobian bound of the type

|JacBar|1+error(δ^)|\operatorname{Jac}\mathrm{Bar}|\leq 1+\text{error}(\hat{\delta})

with respect to the metric 13𝐝δ^\frac{1}{3}\mathbf{d}_{\hat{\delta}}, along any totally geodesic 33-simplex SS in 𝕊+/λΓ(Γ)\mathbb{S}^{+}/\lambda_{\Gamma}(\Gamma), and at any point in SS which is sent inside jj{𝐩}\bigcup_{j}\mathcal{R}_{j}\setminus\{\mathbf{p}\}. The error converges to 0 as δ^\hat{\delta} goes to 0.

Let ΓM:=π1(M)\Gamma_{M}:=\pi_{1}(M) and hMH3(ΓM;)h_{M}\in H_{3}(\Gamma_{M};\mathbb{Z}) the induced class. The natural homomorphism θ:ΓMΓ\theta:\Gamma_{M}\to\Gamma induces a 11-Lipschitz map

Θ:S/λΓM(ΓM)S/λΓ(Γ),\Theta:{S^{\infty}}/\lambda_{\Gamma_{M}}(\Gamma_{M})\to{S^{\infty}}/\lambda_{\Gamma}(\Gamma),

see Subsection 3.4.

Let C𝒞(hM)C\in\mathscr{C}(h_{M}). As before, by Lemma 1.6, one can assume that CC is a polyhedral chain and each point of its support lifts to an 2\ell^{2} function with finite support in ΓM\Gamma_{M}. Then Θ(spt(C))𝕊+/λΓ(Γ)\Theta(\operatorname{spt}(C))\subset\mathbb{S}^{+}/\lambda_{\Gamma}(\Gamma). An arbitrarily small perturbation of Θ\Theta, still denoted by Θ\Theta for simplicity, makes Θ(C)\Theta_{\sharp}(C) a polyhedral chain in 𝕊+/λΓ(Γ)\mathbb{S}^{+}/\lambda_{\Gamma}(\Gamma). Set

Bar~:=BarΘ.\tilde{\mathrm{Bar}}:=\mathrm{Bar}\circ\Theta.

By Γ\Gamma-equivariance, Bar~(C)=1M\tilde{\mathrm{Bar}}_{\sharp}(C)=\llbracket 1_{M}\rrbracket. Note that Bar~\tilde{\mathrm{Bar}} depends on CC because of the perturbation of Θ\Theta but this dependence will not play a role later.

As in the proof of Theorem 27, the Jacobian bound for the barycenter map implies by the area formula of Subsection 1.4 that for all ϵ>0\epsilon>0, whenever δ^\hat{\delta} is small enough,

𝐌(C)\displaystyle\mathbf{M}(C) (13)3/2j=1kVol(j,gj,δ^)ϵ\displaystyle\geq(\frac{1}{3})^{3/2}\sum_{j=1}^{k}\operatorname{Vol}(\mathcal{R}_{j},g_{j,\hat{\delta}})-\epsilon
(13)3/2j=1kVol(Hj,gj,hyp)2ϵ.\displaystyle\geq(\frac{1}{3})^{3/2}\sum_{j=1}^{k}\operatorname{Vol}(H_{j},g_{j,\mathrm{hyp}})-2\epsilon.

Since C𝒞(hM)C\in\mathscr{C}(h_{M}) was arbitrary, by sending δ^0\hat{\delta}\to 0 we get

SphereVol(M)j=1kVol(Hj,13gj,hyp))=Vol(Mhyp,13ghyp).\operatorname{SphereVol}(M)\geq\sum_{j=1}^{k}\operatorname{Vol}(H_{j},\frac{1}{3}g_{j,\mathrm{hyp}}))=\operatorname{Vol}(M_{\mathrm{hyp}},\frac{1}{3}g_{\mathrm{hyp}}).

As for the reverse inequality

SphereVol(M)Vol(Mhyp,13ghyp),\operatorname{SphereVol}(M)\leq\operatorname{Vol}(M_{\mathrm{hyp}},\frac{1}{3}g_{\mathrm{hyp}}),

in view of [BCG91, Corollary 3.13] (see (29) in the proof of Theorem 4.1), one just needs to exhibit a metric on MM with volume entropy close to 22 and volume close to that of the hyperbolic part. In [Pie19, Theorem 4.1]555Theorem 4.3 in [Pie19] is not quite correct as stated, but this is not a serious issue. The statement can be replaced by the following: for any ϵ>0\epsilon>0 and any two closed Riemannian manifolds (Y,g)(Y^{\prime},g^{\prime}), (Y′′,g′′)(Y^{\prime\prime},g^{\prime\prime}), there is a metric gg on the connected sum such that the volume entropies satisfy h(g)h(g)+h(g′′)+ϵh(g)\leq h(g^{\prime})+h(g^{\prime\prime})+\epsilon and |Vol(Y,g)Vol(Y,g)Vol(Y′′,g′′)|ϵ|\operatorname{Vol}(Y,g)-\operatorname{Vol}(Y^{\prime},g^{\prime})-\operatorname{Vol}(Y^{\prime\prime},g^{\prime\prime})|\leq\epsilon. the author constructs a metric 𝐠δ^\mathbf{g}_{\hat{\delta}} on MM depending on δ^\hat{\delta}, with the property that, as δ^0\hat{\delta}\to 0,

  • Vol(M,𝐠δ^)\operatorname{Vol}(M,\mathbf{g}_{\hat{\delta}}) converges to j=1kVol(Hj,gj,hyp)=Vol(Mhyp,ghyp),\sum_{j=1}^{k}\operatorname{Vol}(H_{j},g_{j,\mathrm{hyp}})=\operatorname{Vol}(M_{\mathrm{hyp}},g_{\mathrm{hyp}}),

  • the volume entropy h(𝐠δ^)h(\mathbf{g}_{\hat{\delta}}) converges to 22.

This finishes the outline.

5.2. Intrinsic uniqueness of spherical Plateau solutions

In the main theorem of this section, we interpret the hyperbolic part of a closed oriented 33-manifold as its unique spherical Plateau solution, up to intrinsic isomorphism.

Theorem 5.2.

Let MM be a closed oriented 33-manifold, whose hyperbolic part is denoted by (Mhyp,ghyp)(M_{\mathrm{hyp}},g_{\mathrm{hyp}}). Then any spherical Plateau solution CC_{\infty} for MM is intrinsically isomorphic to (Mhyp,13ghyp)(M_{\mathrm{hyp}},\frac{1}{{3}}g_{\mathrm{hyp}}).

Outline of proof.

The proof is similar to that of Theorem 4.2. Let us give a rough idea of how it works.

Let {Ci}𝒞(hM)\{C_{i}\}\subset\mathscr{C}(h_{M}) be a sequence satisfying

(40) limi𝐌(Ci)=SphereVol(M)=Vol(Mhyp,13ghyp),\lim_{i\to\infty}\mathbf{M}({C}_{i})=\operatorname{SphereVol}(M)=\operatorname{Vol}(M_{\mathrm{hyp}},\frac{1}{3}g_{\mathrm{hyp}}),

where the second equality follows from Theorem 5.1. Suppose that Ci{C}_{i} converges in the intrinsic flat topology to a spherical Plateau solution C=(X,d,S)C_{\infty}=(X_{\infty},d_{\infty},S_{\infty}).

Let (Y,𝐝hyp)(Y,\mathbf{d}_{\mathrm{hyp}}) be the space obtained by taking j=1k(Hj,gj,hyp)\bigsqcup_{j=1}^{k}(H_{j},g_{j,{\mathrm{hyp}}}) and identifying kk points {y1,,yk}\{y_{1},...,y_{k}\} which belong respectively to H1,,HkH_{1},...,H_{k}. Denote by 𝐩\mathbf{p} this unique singular point of YY (in general YY is not path connected).

We will use the notations of the proof of Theorem 5.1. Take a sequence {δ^i}\{\hat{\delta}_{i}\} converging to 0 as ii\to\infty. Consider the metric space (X,𝐝δ^i)(X,\mathbf{d}_{\hat{\delta}_{i}}) and its fundamental group Γ:=π1(X).\Gamma:=\pi_{1}(X). Recall that (X,𝐝δ^i)(X,\mathbf{d}_{\hat{\delta}_{i}}) is obtained by attaching summands (Y1,,Yk)(Y_{1},...,Y_{k}) at one point. By abuse of notations, we call that unique singular point of (X,𝐝δ^i)(X,\mathbf{d}_{\hat{\delta}_{i}}) by 𝐩\mathbf{p} too. The pointed sequence (X,𝐝δ^i,𝐩)(X,\mathbf{d}_{\hat{\delta}_{i}},\mathbf{p}) visibly “converges” to (Y,𝐝hyp,𝐩)(Y,\mathbf{d}_{\mathrm{hyp}},\mathbf{p}) as ii\to\infty.

Consider the barycenter maps

Bar~:=Bar Θ:spt(Ci)(X,𝐝δ^i),\tilde{\mathrm{Bar}}:=\mathrm{Bar}\circ \Theta:\operatorname{spt}(C_{i})\to(X,\mathbf{d}_{\hat{\delta}_{i}}),

see notations in the proof of Theorem 5.1. Modulo some technical details, by following the proof of Theorem 4.2 step by step applied to Bar~\tilde{\mathrm{Bar}}, we arrive at the following partial conclusion. Recall that C=(X,d,S)C_{\infty}=(X_{\infty},d_{\infty},S_{\infty}) is the spherical Plateau solution, limit of CiC_{i}. There is a region ZXZ\subset X_{\infty} which is locally path connected, and if LdL_{d_{\infty}} (resp. LhypL_{\mathrm{hyp}}) is the path metric induced by dd_{\infty} (resp. 𝐝hyp\mathbf{d}_{\mathrm{hyp}}),

(Z,Ld) is isometric to (Y{𝐩},13Lhyp)(Z,L_{d_{\infty}})\text{ is isometric to }(Y\setminus\{\mathbf{p}\},\frac{1}{3}L_{\mathrm{hyp}})

via a map

Bar~:ZY{𝐩}\tilde{\mathrm{Bar}}_{\infty}:Z\to Y\setminus\{\mathbf{p}\}

which is a “limit” of the barycenter maps Bari\mathrm{Bar}_{i}. Moreover,

(Bar~)SZ=1Y(\tilde{\mathrm{Bar}}_{\infty})_{\sharp}S_{\infty}\llcorner Z=\llbracket 1_{Y}\rrbracket

where 1Y\llbracket 1_{Y}\rrbracket is the natural current induced by the oriented finite volume Riemannian space YY. We also have 𝐌(SZ)=𝐌(C)\mathbf{M}(S_{\infty}\llcorner Z)=\mathbf{M}(C_{\infty}).

The final step is to show that (X,Ld)(X_{\infty},L_{d_{\infty}}) is in fact isometric to the 13Lhyp\frac{1}{3}L_{\mathrm{hyp}}-completion of (Y{𝐩},13Lhyp)(Y\setminus\{\mathbf{p}\},\frac{1}{3}L_{\mathrm{hyp}}), which is isometric to the disjoint union

j=1k(Hj,13gj,hyp)=(Mhyp,13ghyp).\bigsqcup_{j=1}^{k}(H_{j},\frac{1}{3}g_{j,\mathrm{hyp}})=(M_{\mathrm{hyp}},\frac{1}{3}g_{\mathrm{hyp}}).

Essentially, what we want to rule out is that (X,d)(X_{\infty},d_{\infty}) is isometric to a non-smooth space made of manifolds attached at one point, for instance (Y,13L𝐝hyp)(Y,\frac{1}{3}L_{\mathbf{d}_{\mathrm{hyp}}}). Actually, the apparent issue with the point 𝐩\mathbf{p} is only an artefact of our definition of the barycenter map, and is not related to the geometry of spherical Plateau solutions. By playing with the important property that the set of Riemannian isometries of a finite union of finite volume hyperbolic 33-manifolds is finite, and by choosing the attachment point 𝐩\mathbf{p} generically enough, we conclude that the only possibility is the desired statement: (X,Ld)(X_{\infty},L_{d_{\infty}}) is isometric as a length space to (Mhyp,13ghyp)(M_{\mathrm{hyp}},\frac{1}{3}g_{\mathrm{hyp}}) and the proof is completed.

More details for the proof of Theorem 5.2 appeared in the earlier preprint [Son22].

By Theorems 5.1 and 5.2, the spherical volume behaves nicely under geometric decompositions of 33-manifolds. To what extent this holds for general group homology classes is an interesting problem (see [Gro82] for such properties in the context of the simplicial volume):

Question 10 (Additivity under connected sum).

Given two closed oriented manifolds M1M_{1}, M2M_{2} of same dimension at least 3, if M1M2M_{1}\sharp M_{2} denotes the connected sum, do we have

SphereVol(M1M2)=SphereVol(M1)+SphereVol(M1)?\operatorname{SphereVol}(M_{1}\sharp M_{2})=\operatorname{SphereVol}(M_{1})+\operatorname{SphereVol}(M_{1})?

Does any spherical Plateau solution for M1M2M_{1}\sharp M_{2} decompose into the union of spherical Plateau solutions for M1M_{1} and M2M_{2}?

Another intriguing problem suggested by Theorems 5.1 and 5.2 is the following:

Question 11 (RF and MCF).

What is the interplay between the spherical Plateau problem and the Ricci flow or the mean curvature flow?

6. Plateau Dehn fillings

6.1. Preliminaries on CAT(0)\mathrm{CAT}(0) Dehn fillings

In [FM10] Fujiwara and Manning constructed certain pseudomanifolds out of higher dimensional finite volume hyperbolic manifolds, which generalize 33-dimensional Dehn fillings from a topological and group theoretic point of view. In dimensions larger than 33, the spherical Plateau solutions associated to those pseudomanifolds play the role of hyperbolic Dehn fillings in dimension 33. In this section we are interested in the asymptotic behavior of those “Plateau Dehn fillings”.

First let us review some of the definitions and results in [FM10]. Let n3n\geq 3. Consider a finite volume oriented hyperbolic nn-manifold (M,ghyp)(M,g_{\mathrm{hyp}}) with disjoint toral cusps E1,,EmE_{1},...,E_{m} and no other ends. Here a toral cusp means an end of MM homeomorphic to the product of a torus with (,0](-\infty,0], such that the induced Riemannian metric on E\partial E is flat. Write M¯:=Mj=1mE̊j\bar{M}:=M\setminus\bigcup_{j=1}^{m}\mathring{E}_{j}. Inside each component Ej\partial E_{j} of M¯\partial\bar{M}, choose an embedded totally geodesic torus TjT_{j} of dimension kik_{i} where kj{1,,n1}k_{j}\in\{1,...,n-1\}. TjT_{j} is a leaf of a fibration EjBj\partial E_{j}\to B_{j} with base an (n1kj)(n-1-k_{j})-torus BjB_{j} and leaves kjk_{j}-tori (note that in [FM10], contrarily to here, the ambient dimension is by convention n+1n+1). One can form the topological space M(T1,,Tm)M(T_{1},...,T_{m}) by collapsing each leaf of these fibrations EjBj\partial E_{j}\to B_{j} to points (see [FM10, Definition 2.5]). This space is a pseudomanifold which is smooth outside of the so-called filling cores V1,VmV_{1},...V_{m} where VjV_{j} is an (n1kj)(n-1-k_{j})-torus. M(T1,,Tm)M(T_{1},...,T_{m}) is not a manifold except when for all j{1,,m}j\in\{1,...,m\}, kj=1k_{j}=1. There is a natural map

MM(T1,,Tm)M\to M(T_{1},...,T_{m})

which induces a natural surjection

(41) π1(M)π1(M(T1,,Tm)).\pi_{1}(M)\to\pi_{1}(M(T_{1},...,T_{m})).

In dimension 33 when each kj=1k_{j}=1, by classical works [Thu97, BH96] it is known that when each circle TjT_{j} has length larger than 2π2\pi, M(T1,,Tm)M(T_{1},...,T_{m}) is a hyperbolic manifold: this is the so-called 2π2\pi theorem. The main theorem in [FM10] generalizes the 2π2\pi theorem to higher dimensions. It asserts that M(T1,,Tm)M(T_{1},...,T_{m}) admits a locally CAT(0)\mathrm{CAT}(0) path metric 𝐝\mathbf{d} whenever none of the tori TjT_{j} admit a closed geodesic of length at most 2π2\pi, or equivalently when the injectivity radius of each TjT_{j} with its intrinsic metric is strictly larger than π\pi. Fujiwara and Manning conjectured in [FM11, Conjecture 1.8, Question 1.9] that as injrad(Tj)\operatorname{injrad}(T_{j})\to\infty, the simplicial volume of M(T1,,Tm)M(T_{1},...,T_{m}) should converge to the simplicial volume of MM and approach that limit from below.

A useful property of the locally CAT(0)\mathrm{CAT}(0) metric 𝐝\mathbf{d} constructed in [FM10] is that it can be chosen to approximate the hyperbolic metric on MM on large sets when the injectivity radii of the TjT_{j} are tending to infinity:

Lemma 6.1.

Fix a point pMp\in M. For any R>0R>0, there is iR>0i_{R}>0 such that if

(42) injrad(Tj)>iR for all j{1,,m},\operatorname{injrad}(T_{j})>i_{R}\quad\text{ for all $j\in\{1,...,m\}$},

then dimTj<n1\dim T_{j}<n-1 for all j{1,,m}j\in\{1,...,m\} and one can choose the metric 𝐝\mathbf{d} on M(T1,,Tm)M(T_{1},...,T_{m}) so that there is a closed geodesic ball R\mathcal{B}_{R} of radius RR inside (M(T1,,Tm),𝐝)(M(T_{1},...,T_{m}),\mathbf{d}) isometric to the closed geodesic ball of radius RR centered at pp inside MM.

Proof.

Note that when iRi_{R} is large enough depending also on MM, then (42) necessarily implies that TjEjT_{j}\neq\partial E_{j} i.e. dimTj<n1\dim T_{j}<n-1 for all j{1,,m}j\in\{1,...,m\}. For any ϵ>0\epsilon>0, if iRi_{R} is large enough and satisfies (42), then we can replace the toral cusps EjE_{j} by new toral cusps E~jEj\tilde{E}_{j}\subset E_{j} which are far away inside the ends of MM, so that Vol(Mj=1mE~j,ghyp)Vol(M,ghyp)ϵ\operatorname{Vol}(M\setminus\bigcup_{j=1}^{m}{\tilde{E}}_{j},g_{\mathrm{hyp}})\geq\operatorname{Vol}(M,g_{\mathrm{hyp}})-\epsilon and yet

injrad(T~j)>π,\operatorname{injrad}(\tilde{T}_{j})>\pi,

where T~j\tilde{T}_{j} is the totally geodesic torus in E~j\partial\tilde{E}_{j} corresponding to TjEjT_{j}\subset\partial E_{j}. Then we apply [FM10, Proposition 2.8] to the fillings obtained with T~jE~j\tilde{T}_{j}\subset\partial\tilde{E}_{j} (which is homeomorphic to M(T1,,Tm)M(T_{1},...,T_{m})).

Outside of the filling cores, 𝐝\mathbf{d} is locally induced by a Riemannian metric of strictly negative curvature [FM10, Theorem 2.7]. The nn-dimensional Hausdorff measure on M(T1,,Tm)M(T_{1},...,T_{m}) coincides with the Lebesgue measure outside of the filling cores, which have 0 Hausdorff measure. We denote by M~(T1,,Tm)\tilde{M}(T_{1},...,T_{m}) the universal cover of M(T1,,Tm){M}(T_{1},...,T_{m}), and by Ξ\Xi the singular set in M~(T1,,Tm)\tilde{M}(T_{1},...,T_{m}), namely the union of the lifts of the filling cores in M(T1,,Tm){M}(T_{1},...,T_{m}). Each component of Ξ\Xi is a totally geodesic embedded Euclidean space [FM10, Subsection 4.6]. On M~(T1,,Tm)\tilde{M}(T_{1},...,T_{m}), the lift of 𝐝\mathbf{d} is CAT(0)\mathrm{CAT}(0) and the group

Γ:=π1(M(T1,,Tm))\Gamma:=\pi_{1}(M(T_{1},...,T_{m}))

acts freely properly cocompactly by isometries.

In the sequel we will always assume that minj=1minjrad(Tj)>2π\min_{j=1}^{m}\operatorname{injrad}(T_{j})>2\pi, so that the path metric 𝐝\mathbf{d} constructed in [FM10] exists. An important fact is that one can define a barycenter map associated with (M~(T1,,Tm),𝐝)(\tilde{M}(T_{1},...,T_{m}),\mathbf{d}):

Bar:𝕊+/λΓ(Γ)(M~(T1,,Tm),𝐝)\mathrm{Bar}:\mathbb{S}^{+}/\lambda_{\Gamma}(\Gamma)\to(\tilde{M}(T_{1},...,T_{m}),\mathbf{d})

and it has similar properties to the one reviewed in Section 2 (here the definitions of 𝕊+\mathbb{S}^{+} and Bar\mathrm{Bar} need to be adapted). There is a technical difficulty caused by the fact that the distance functions on (M~(T1,,Tm),𝐝)(\tilde{M}(T_{1},...,T_{m}),\mathbf{d}) are potentially non-smooth on large sets. For instance the fact that Hessian bounds still make sense in a weak sense and the fact that a cycle in S/λΓ(Γ){S^{\infty}}/\lambda_{\Gamma}(\Gamma) can be perturbed to a cycle in 𝕊+/λΓ(Γ)\mathbb{S}^{+}/\lambda_{\Gamma}(\Gamma) require some arguments. We will not discuss those issues but we mention the following useful papers [FM10, KL21].

6.2. Asymptotic rigidity of Plateau Dehn fillings

For n3n\geq 3, let (M,ghyp)(M,g_{\mathrm{hyp}}) be non-compact oriented hyperbolic nn-manifold of finite volume, with only toral cusps E1,,EmE_{1},...,E_{m}. Let TiEiT_{i}\subset\partial E_{i} be an embedded totally geodesic kik_{i}-dimensional torus. We will always suppose that the injectivity radii of TiT_{i} are larger than π\pi. By residual finiteness, any finite volume hyperbolic manifold is finitely covered by such a hyperbolic manifold. Denote by gg^{\prime} the following rescaling of the hyperbolic metric on MM:

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

Let M(T1,,Tm)M(T_{1},...,T_{m}) be a 2π2\pi-filling constructed by Fujiwara-Manning in [FM10] endowed with the metric 𝐝\mathbf{d} satisfying Lemma 6.1, for some positive dimensional tori TiT_{i}, as explained in the previous subsection. Denote by Γ\Gamma (resp. hHn(Γ;)h\in H_{n}(\Gamma;\mathbb{Z})) the fundamental group (resp. the fundamental class) of M(T1,,Tm)M(T_{1},...,T_{m}). We will say that hh is the 2π2\pi-filling homology class corresponding to T1,,TmT_{1},...,T_{m}.

Recall that intrinsic equivalence for integral currents spaces is defined in Definition 3.5. The following theorem establishes the asymptotic rigidity of Plateau Dehn fillings as injrad(Ti)\operatorname{injrad}(T_{i})\to\infty, which in particular implies the spherical volume analogue of the conjecture of Fujiwara-Manning [FM11, Conjecture 1.8, Question 1.9]. This behavior is completely analogous to what happens to hyperbolic 33-dimensional Dehn fillings.

Theorem 6.2.

Let (M,ghyp)(M,g_{\mathrm{hyp}}) be a non-compact finite volume oriented hyperbolic manifold with toral cusps, then the following holds.

  1. (1)

    For any 2π2\pi-filling homology class hh,

    SphereVol(h)<Vol(M,g).\operatorname{SphereVol}(h)<\operatorname{Vol}(M,g^{\prime}).
  2. (2)

    Consider a sequence of families of tori T1p,,TmpT^{p}_{1},...,T^{p}_{m} such that

    limpmini=1minjrad(Tip)=.\lim_{p\to\infty}\min_{i=1}^{m}\operatorname{injrad}(T^{p}_{i})=\infty.

    Let Cp,C_{p,\infty} be any spherical Plateau solution for the 2π2\pi-filling homology class hph_{p} corresponding to T1p,,TmpT^{p}_{1},...,T^{p}_{m}. Then

    limp𝐌(Cp,)=limpSphereVol(hp)=Vol(M,g),\lim_{p\to\infty}\mathbf{M}(C_{p,\infty})=\lim_{p\to\infty}\operatorname{SphereVol}(h_{p})=\operatorname{Vol}(M,g^{\prime}),

    and Cp,C_{p,\infty} subsequentially converges in the intrinsic flat topology to an integral current space which is intrinsically isomorphic to (M,g)(M,g^{\prime}).

Outline of proof.

Let us start with the strict inequality (1). Recall that

Γ:=π1(M(T1,,Tm)).\Gamma:=\pi_{1}(M(T_{1},...,T_{m})).

Let M~\tilde{M} be the universal cover of MM endowed with the hyperbolic metric g:=ghypg:=g_{\mathrm{hyp}}. Set

S2(M~,g):={uL2(M~,g);uL2=1}S_{2}(\tilde{M},g):=\{u\in L^{2}(\tilde{M},g);\quad\|u\|_{L^{2}}=1\}

and denote by λ(M~,g)\lambda_{(\tilde{M},g)} the natural π1(M)\pi_{1}(M)-action on this sphere. The fundamental group π1(M)\pi_{1}(M) naturally surjects onto Γ\Gamma by (41), and as in Subsections 3.4, there is a natural distance non-increasing map

Θ:S2(M~,g)/λ(M~,g)(π1(M))S/λπ1(M)(π1(M))S/λΓ(Γ)\Theta:S_{2}(\tilde{M},g)/\lambda_{(\tilde{M},g)}(\pi_{1}(M))\to{S^{\infty}}/\lambda_{\pi_{1}(M)}(\pi_{1}(M))\to{S^{\infty}}/\lambda_{\Gamma}(\Gamma)

where the first map is obtained by averaging on translates of a fundamental domain.

An explicit sequence of admissible immersions from MM to S2(M~,g)/λ(M~,g)(π1(M))S_{2}(\tilde{M},g)/\lambda_{(\tilde{M},g)}(\pi_{1}(M)) whose image has volume converging to Vol(M,(n1)24ng)\operatorname{Vol}(M,\frac{(n-1)^{2}}{4n}g) is given by immersions defined as follows. Let φ~:[0,)(0,1)\tilde{\varphi}:[0,\infty)\to(0,1) be a nondecreasing smooth function such that the function

ρ~y(.):=φ~(distg(y,.))\tilde{\rho}_{y}(.):=\tilde{\varphi}(\operatorname{dist}_{g}(y,.))

is smooth and coincides with distg(y,.)\operatorname{dist}_{g}(y,.) outside of the 11-neighborhood of yM~y\in\tilde{M}. For each c>n1c>n-1, set

α~c:xα~c,x:=ec2ρ~x(.)L2(M~,g).\tilde{\alpha}_{c}:x\mapsto\tilde{\alpha}_{c,x}:=e^{-\frac{c}{2}\tilde{\rho}_{x}(.)}\in L^{2}(\tilde{M},{g}).

By homogeneity the norm α~c,xL2\|\tilde{\alpha}_{c,x}\|_{L^{2}} does not depend on xM~x\in\tilde{M}. Define the corresponding immersion by

𝒫¯c:(M~,g)S2(M~,g)\overline{\mathcal{P}}_{c}:(\tilde{M},g^{\prime})\to S_{2}(\tilde{M},g)
x𝒫¯c,x=α~c,xα~c,xL2.x\mapsto\overline{\mathcal{P}}_{c,x}=\frac{\tilde{\alpha}_{c,x}}{\|\tilde{\alpha}_{c,x}\|_{L^{2}}}.

This map is π1(M)\pi_{1}(M)-equivariant and descends to a map from MM to S2(M~,g)/λ(M~,g)(π1(M))S_{2}(\tilde{M},g)/\lambda_{(\tilde{M},g)}(\pi_{1}(M)). By well-known properties of hyperbolic spaces and their compactifications [BCG95, Subsection 2.6], as cn1c\to n-1, for any unit vector vTxM~v\in T_{x}\tilde{M},

limcn1dx𝒫¯c,x(v)L22=limcn1c24M~|dxρ~y(v)|2𝒫¯c,x2(y)𝑑volg(y)=(n1)24n.\lim_{c\to n-1}\|d_{x}\overline{\mathcal{P}}_{c,x}(v)\|^{2}_{L^{2}}=\lim_{c\to n-1}\frac{c^{2}}{4}\int_{\tilde{M}}|d_{x}\tilde{\rho}_{y}(v)|^{2}\overline{\mathcal{P}}^{2}_{c,x}(y)dvol_{g}(y)=\frac{(n-1)^{2}}{4n}.

The convergence is uniform on TM~T\tilde{M}. In particular, the pull-back metric converges to

g:=(n1)24ngg^{\prime}:=\frac{(n-1)^{2}}{4n}g

as cn1c\to n-1 and

(43) limcn1Vol(M,(Θ𝒫¯c)𝐠Hil)=Vol(M,g).\lim_{c\to n-1}\operatorname{Vol}(M,(\Theta\circ\overline{\mathcal{P}}_{c})^{*}\mathbf{g}_{\mathrm{Hil}})=\operatorname{Vol}(M,g^{\prime}).

Let θ:π1(M)Γ\theta:\pi_{1}(M)\to\Gamma be the natural surjective homomorphism and fix a fundamental domain DM~D\subset\tilde{M}. Consider the immersion

Θ𝒫¯c:MS/λΓ(Γ).\Theta\circ\overline{\mathcal{P}}_{c}:M\to{S^{\infty}}/\lambda_{\Gamma}(\Gamma).

Unwinding the definitions, we get for x M~x\in \tilde{M}, γΓ\gamma\in\Gamma:

Θ𝒫¯c,x(γ)=[τθ1(γ)τ.Dα~c,x2α~c,xL22𝑑volg]1/2.\Theta\circ\overline{\mathcal{P}}_{c,x}(\gamma)=\big{[}\sum_{\tau\in\theta^{-1}(\gamma)}\int_{\tau.D}\frac{\tilde{\alpha}_{c,x}^{2}}{\|\tilde{\alpha}_{c,x}\|_{L^{2}}^{2}}dvol_{g}\big{]}^{1/2}.

Form this expression, one checks that Θ𝒫¯c\Theta\circ\overline{\mathcal{P}}_{c} has uniformly bounded second derivatives in xx as cn1c\to n-1.

Let hHn(Γ;)h\in H_{n}(\Gamma;\mathbb{Z}) be the 2π2\pi-filling homology class corresponding to T1,,TmT_{1},...,T_{m}. The push-forward current (Θ𝒫¯c)1M(\Theta\circ\overline{\mathcal{P}}_{c})_{\sharp}\llbracket 1_{M}\rrbracket is in general not an element of 𝒞(h)\mathscr{C}(h) since MM is noncompact. Nevertheless given any η>0\eta>0, we can consider toral cusps E~1,,E~j\tilde{E}_{1},...,\tilde{E}_{j} respectively far enough inside E1,,EjE_{1},...,E_{j} and “close” the image tori Θ𝒫¯c(E~j)\Theta\circ\overline{\mathcal{P}}_{c}(\partial\tilde{E}_{j}) without adding much volume (by the cone construction in S{S^{\infty}}), to obtain a new admissible Lipschitz map

Ψ:M(T1,,Tm)S/λΓ(Γ)\Psi:M(T_{1},...,T_{m})\to{S^{\infty}}/\lambda_{\Gamma}(\Gamma)

such that Ψ(1M(T1,,Tm))𝒞(h)\Psi_{\sharp}(\llbracket 1_{M(T_{1},...,T_{m})}\rrbracket)\in\mathscr{C}(h) and

(44) 𝐌(Ψ(1M(T1,,Tm)))Vol(M,(Θ𝒫¯c)𝐠Hil)+η.\mathbf{M}(\Psi_{\sharp}(\llbracket 1_{M(T_{1},...,T_{m})}\rrbracket))\leq\operatorname{Vol}(M,(\Theta\circ\overline{\mathcal{P}}_{c})^{*}\mathbf{g}_{\mathrm{Hil}})+\eta.

Taking η0\eta\to 0, (43) and the above inequality already imply

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

Suppose towards a contradiction that SphereVol(h)=Vol(M,g)\operatorname{SphereVol}(h)=\operatorname{Vol}(M,g^{\prime}). Then (44) implies that if 𝒟c,x\mathcal{D}_{c,x} is the differential of Θ𝒫¯c:(M,g)S/λΓ(Γ)\Theta\circ\overline{\mathcal{P}}_{c}:(M,g^{\prime})\to{S^{\infty}}/\lambda_{\Gamma}(\Gamma) at xMx\in M, we must have for any unit norm tangent vector vTxMv\in T_{x}M

(45) limcn1𝒟c,x(v)=1\lim_{c\to n-1}\|\mathcal{D}_{c,x}(v)\|=1

and the convergence is uniform on compact sets in MM since the second derivatives of Θ𝒫¯c\Theta\circ\overline{\mathcal{P}}_{c} are uniformly bounded.

For any small enough l>0l>0, there is a smooth closed curve a:[0,l](M,g)a:[0,l]\to(M,g^{\prime}) of length ll parametrized by arclength with a(0)=a(1)a(0)=a(1), such that

  • a:[0,l]Ma:[0,l]\to M to represents an element in the non-empty kernel ker(θ)\ker(\theta) of θ:π1(M)Γ,\theta:\pi_{1}(M)\to\Gamma,

  • at any point of the closed curve aa, the geodesic curvature is equal to 11.

From what we said above, the image of

Θ𝒫¯ca:[0,1]S/λΓ(Γ)\Theta\circ\overline{\mathcal{P}}_{c}\circ a:[0,1]\to{S^{\infty}}/\lambda_{\Gamma}(\Gamma)

is a loop contained in S/λΓ(Γ){S^{\infty}}/\lambda_{\Gamma}(\Gamma) which is homotopically trivial (hence it lifts isometrically to the sphere S{S^{\infty}}) and has length at most 2l2l for all cc close enough to n1n-1. Moreover the norm of the second derivative of Θ𝒫¯ca\Theta\circ\overline{\mathcal{P}}_{c}\circ a is uniformly bounded from above. By (45), the norm of the differential of Θ𝒫¯ca\Theta\circ\overline{\mathcal{P}}_{c}\circ a converges to 11 uniformly as cn1c\to n-1. But if ll is small enough, it is impossible to have an almost arclength parametrized closed curve from [0,l][0,l] to a Hilbert unit sphere with second derivatives bounded independently of ll. This contradiction finishes the proof of the strict inequality.

As for (2), our proof uses the barycenter map in the same fashion as the proofs of Theorems 4.1, 5.1, 4.2 and 5.2. Let (M(T1p,,Tmp),𝐝p)(M(T^{p}_{1},...,T^{p}_{m}),\mathbf{d}_{p}) be a sequence of 2π2\pi-fillings where

(46) limpmink=1minjrad(Tkp)=.\lim_{p\to\infty}\min_{k=1}^{m}\operatorname{injrad}(T^{p}_{k})=\infty.

For each pp, let

Γp:=π1(M(T1p,,Tmp)),\Gamma_{p}:=\pi_{1}(M(T^{p}_{1},...,T^{p}_{m})),

let hpHn(Γp;)h_{p}\in H_{n}(\Gamma_{p};\mathbb{Z}) be the corresponding homology classes and let Cp,C_{p,\infty} be a spherical Plateau solution for hph_{p}. By Wenger’s compactness theorem [SW11, 4.19], one can assume that Cp,C_{p,\infty} converges in the intrinsic flat topology to some limit integral current space

W=(XW,dW,SW).W=(X_{W},d_{W},S_{W}).

As a spherical Plateau solution, Cp,C_{p,\infty} is the intrinsic flat limit of a minimizing sequence {Cp,i}i0𝒞(hp)\{C_{p,i}\}_{i\geq 0}\subset\mathscr{C}(h_{p}).

From now on, we will consider

𝐝p:=(n1)24n𝐝p.\mathbf{d}^{\prime}_{p}:=\frac{(n-1)^{2}}{4n}\mathbf{d}_{p}.

In the sequel, Jacobians, lengths and distances will be computed with respect to 𝐝p\mathbf{d}^{\prime}_{p} on M(T1p,,Tmp)M(T^{p}_{1},...,T^{p}_{m}). As we briefly explained earlier, there is a well-defined barycenter map

Bar:𝕊+/λΓp(Γp)M(T1p,,Tmp)\mathrm{Bar}:\mathbb{S}^{+}/\lambda_{\Gamma_{p}}(\Gamma_{p})\to M(T^{p}_{1},...,T^{p}_{m})

associated with Γp\Gamma_{p}, (M~(T1p,,Tmp),𝐝p)(\tilde{M}(T^{p}_{1},...,T^{p}_{m}),\mathbf{d}^{\prime}_{p}) (here 𝕊+\mathbb{S}^{+} depends on Γp\Gamma_{p}.). As before, using Lemma 1.6, all the Cp,iC_{p,i} can be assumed to be polyhedral chains without loss of generality. For each pp and each ii, one can show that

(Barp)(Cp,i)=1M(T1p,,Tmp)({\mathrm{Bar}}_{p})_{\sharp}(C_{p,i})=\llbracket 1_{M(T^{p}_{1},...,T^{p}_{m})}\rrbracket

where 1M(T1p,,Tmp)\llbracket 1_{M(T^{p}_{1},...,T^{p}_{m})}\rrbracket is the integral current of M(T1p,,Tmp)M(T^{p}_{1},...,T^{p}_{m}) representing the fundamental class [M(T1p,,Tmp)]Hn(M(T1p,,Tmp);)[M(T^{p}_{1},...,T^{p}_{m})]\in H_{n}(M(T^{p}_{1},...,T^{p}_{m});\mathbb{Z}).

Given ϵ>0\epsilon>0, let RR be such that if

(47) mink=1minjrad(Tkp)>iR,\min_{k=1}^{m}\operatorname{injrad}(T^{p}_{k})>i_{R},

then for the ball RM(T1p,,Tmp)\mathcal{B}_{R}\subset M(T^{p}_{1},...,T^{p}_{m}) as in Lemma 6.1,

(48) Vol(R/2,g)>Vol(M,g)ϵ/2\operatorname{Vol}(\mathcal{B}_{R/2},g^{\prime})>\operatorname{Vol}(M,g^{\prime})-\epsilon/2

where g:=(n1)24ngg^{\prime}:=\frac{(n-1)^{2}}{4n}g. As in Theorems 4.1 and 5.1, the Jacobian bound for the barycenter maps and the area formula imply that for all RR large,

𝐌(Cp,i)>Vol(M,g)ϵ.\mathbf{M}(C_{p,i})>\operatorname{Vol}(M,g^{\prime})-\epsilon.

Combined with the upper bound shown in (1), this already proves that

(49) limpSphereVol(hp)=Voln(M,g).\lim_{p\to\infty}\operatorname{SphereVol}(h_{p})=\operatorname{Vol}_{n}(M,g^{\prime}).

It remains to explain why the intrinsic flat limit W=(XW,dW,SW)W=(X_{W},d_{W},S_{W}) of {Cp,}\{C_{p,\infty}\} is intrinsically isomorphic to (M,g)(M,g^{\prime}). The proof is based on arguments similar to those of Theorem 4.2 and Theorem 5.2. By Lemma 6.1, the pseudomanifolds (M(T1p,,Tmp),𝐝p)(M(T^{p}_{1},...,T^{p}_{m}),\mathbf{d}^{\prime}_{p}) contain larger and larger geodesic balls R\mathcal{B}_{R} isometrically contained in (M,g)(M,g^{\prime}) and

M=R>0R.M=\bigcup_{R>0}\mathcal{B}_{R}.

Repeating the arguments of Theorem 4.2, we conclude that there is a region ZXWZ\subset X_{W} which is path connected, and if LdWL_{d_{W}} (resp. LgL_{g^{\prime}}) denotes the path metric induced by dWd_{W} (resp. gg^{\prime}),

(Z,LdW) is isometric to (M,g)(Z,L_{d_{W}})\text{ is isometric to }(M,g^{\prime})

via a “limit barycenter map”

Bar:ZM{\mathrm{Bar}}_{\infty}:Z\to M

such that (Bar)SWZ=1M({\mathrm{Bar}}_{\infty})_{\sharp}S_{W}\llcorner Z=\llbracket 1_{M}\rrbracket. Moreover by lower semicontinuity of the mass, 𝐌(W)Vol(M,g)\mathbf{M}(W)\leq\operatorname{Vol}(M,g^{\prime}) so by the above isometry Bar{\mathrm{Bar}}_{\infty}, necessarily XW=ZX_{W}=Z. This finishes the proof.

More details for the proof of Theorem 6.2 appeared in the earlier preprint [Son22].

Theorem 6.2 provides many examples of sequences of spherical Plateau solutions {Cp,}\{C_{p,\infty}\} which “accumulate” towards a limit. Note that this accumulation occurs from below in the sense that the mass of each Cp,C_{p,\infty} is strictly less than the mass of the limit. The set of simplicial volumes [Gro82] of closed oriented manifolds is well-ordered in dimension 22 and 33, but not in higher dimensions, see [HL21]. There is also a conjecture (due to Harold Rosenberg?) stating that the set of areas of closed minimal surfaces in the round 33-sphere is well-ordered. By Theorem 4.1, Theorem 5.1 and [BCG91, Proposition 3.9], the set of spherical volumes of closed oriented manifolds is well-ordered in dimensions 22 and 33. All these elements suggest the following question:

Question 12 (Well-ordering).

Is the set of spherical volumes of closed oriented manifolds well-ordered in any dimension?

In dimensions at least 44, it is possible that the Plateau Dehn fillings we constructed have supports which are non-smooth, but are smooth on large domains, due to the ϵ\epsilon-regularity theorem in [ADLS18]. In fact, no example of singular spherical Plateau solution is known:

Question 13 (Singularity).

Is there a closed oriented manifold with a singular spherical Plateau solution?

The situation treated in Theorem 6.2 is very special due to the existence of a model hyperbolic manifold and the availability of barycenter maps. Considering that there are group theoretic versions of Dehn fillings developed in [Osi07, GM08], it would be desirable to investigate Plateau Dehn fillings beyond the case of cusped hyperbolic manifolds.

Question 14 (Convergence phenomenon).

Are there other instances where “convergence” of a sequence of pairs (Γi,hi)(\Gamma_{i},h_{i}) implies convergence of the corresponding spherical volumes and spherical Plateau solutions?

Theorem 6.2 involves a non-compact hyperbolic manifold MM, whose fundamental group π1(M)\pi_{1}(M) surjects onto the fundamental groups of Dehn fillings M(T1,,Tm)M(T_{1},...,T_{m}). Note that here, MM does not determine a nontrivial group homology class due to its non-compactness, but we can nevertheless interpret the spherical Plateau problems for the Dehn fillings M(T1,,Tm)M(T_{1},...,T_{m}) as spherical Plateau problems for MM with respect to orthogonal representations obtained by composing of the fundamental group surjections and the regular representations. Those remarks point to the following problem:

Question 15 (Generalization).

Can the spherical Plateau problem be extended to non-compact manifolds and general orthogonal representations?

References

  • [ADLS18] Luigi Ambrosio, Camillo De Lellis, and Thomas Schmidt. Partial regularity for mass-minimizing currents in Hilbert spaces. J. Reine Angew. Math., 734:99–144, 2018.
  • [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.
  • [Alm00] Frederick J. Almgren, Jr. Almgren’s big regularity paper, volume 1 of World Scientific Monograph Series in Mathematics. World Scientific Publishing Co., Inc., River Edge, NJ, 2000. QQ-valued functions minimizing Dirichlet’s integral and the regularity of area-minimizing rectifiable currents up to codimension 2, With a preface by Jean E. Taylor and Vladimir Scheffer.
  • [Amb16] Luigi Ambrosio. The regularity theory of area-minimizing integral currents [after Almgren–De Lellis–Spadaro]. Astérisque, (380, Séminaire Bourbaki. Vol. 2014/2015):Exp. No. 1093, 139–169, 2016.
  • [And06] Michael T Anderson. Dehn filling and einstein metrics in higher dimensions. Journal of Differential Geometry, 73(2):219–261, 2006.
  • [AS13] Luigi Ambrosio and Thomas Schmidt. Compactness results for normal currents and the Plateau problem in dual Banach spaces. Proc. Lond. Math. Soc. (3), 106(5):1121–1142, 2013.
  • [Bab06] Ivan K. Babenko. Topologie des systoles unidimensionnelles. Enseign. Math. (2), 52(1-2):109–142, 2006.
  • [Bam12] Richard H Bamler. Construction of einstein metrics by generalized dehn filling. Journal of the European Mathematical Society, 14(3):887–909, 2012.
  • [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.
  • [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.
  • [BdLHV08] Bachir Bekka, Pierre de La Harpe, and Alain Valette. Kazhdan’s property (t). Kazhdan’s property (T), 11, 2008.
  • [BH96] Steven A. Bleiler and Craig D. Hodgson. Spherical space forms and Dehn filling. Topology, 35(3):809–833, 1996.
  • [Bru08] Michael Brunnbauer. Homological invariance for asymptotic invariants and systolic inequalities. Geom. Funct. Anal., 18(4):1087–1117, 2008.
  • [BS23] Ivan Babenko and Stéphane Sabourau. Volume entropy semi-norm and systolic volume semi-norm. Journal of the European Mathematical Society, 2023.
  • [CF02] Christopher Connell and Benson Farb. Some recent applications of the barycenter method in geometry. arXiv preprint math/0204093, 2002.
  • [CWN23] Tamunonye Cheetham-West and Alexander Nolte. Characterizing candidates for cannon’s conjecture from geometric measure theory. Bulletin of the London Mathematical Society, 2023.
  • [DL16] Camillo De Lellis. The regularity of minimal surfaces in higher codimension. In Current developments in mathematics 2014, pages 153–229. Int. Press, Somerville, MA, 2016.
  • [DLS14] Camillo De Lellis and Emanuele Spadaro. Regularity of area minimizing currents I: gradient LpL^{p} estimates. Geom. Funct. Anal., 24(6):1831–1884, 2014.
  • [DP14] Thierry De Pauw. Approximation by polyhedral GG chains in Banach spaces. Z. Anal. Anwend., 33(3):311–334, 2014.
  • [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.
  • [Fed69] Herbert Federer. Geometric measure theory. Die Grundlehren der mathematischen Wissenschaften, Band 153. Springer-Verlag New York Inc., New York, 1969.
  • [Fed74] Herbert Federer. Real flat chains, cochains and variational problems. Indiana University Mathematics Journal, 24(4):351–407, 1974.
  • [FF60] Herbert Federer and Wendell H. Fleming. Normal and integral currents. Ann. of Math. (2), 72(3):458–520, 1960.
  • [FM10] Koji Fujiwara and Jason Fox Manning. CAT(0){\rm CAT(0)} and CAT(1){\rm CAT}(-1) fillings of hyperbolic manifolds. J. Differential Geom., 85(2):229–269, 2010.
  • [FM11] Koji Fujiwara and Jason Fox Manning. Simplicial volume and fillings of hyperbolic manifolds. Algebr. Geom. Topol., 11(4):2237–2264, 2011.
  • [GM08] Daniel Groves and Jason Fox Manning. Dehn filling in relatively hyperbolic groups. Israel J. Math., 168:317–429, 2008.
  • [Gro81] Michael Gromov. Hyperbolic manifolds according to thurston and jørgensen. In Séminaire Bourbaki vol. 1979/80 Exposés 543–560, pages 40–53. Springer, 1981.
  • [Gro82] Michael Gromov. Volume and bounded cohomology. Inst. Hautes Études Sci. Publ. Math., (56):5–99 (1983), 1982.
  • [HL21] Nicolaus Heuer and Clara Löh. The spectrum of simplicial volume. Invent. Math., 223(1):103–148, 2021.
  • [KL08] Bruce Kleiner and John Lott. Notes on Perelman’s papers. Geom. Topol., 12(5):2587–2855, 2008.
  • [KL21] Vitali Kapovitch and Alexander Lytchak. Remarks on manifolds with two-sided curvature bounds. Anal. Geom. Metr. Spaces, 9(1):53–64, 2021.
  • [Lan11] Urs Lang. Local currents in metric spaces. J. Geom. Anal., 21(3):683–742, 2011.
  • [Law75] H Blaine Lawson. The stable homology of a flat torus. Mathematica Scandinavica, 36(1):49–73, 1975.
  • [Liu23] Zhenhua Liu. Homologically area-minimizing surfaces that cannot be calibrated. arXiv preprint arXiv:2310.19860, 2023.
  • [Man23] Cosmin Manea. The spherical Plateau problem for group homology and Cannon’s conjecture. Master’s Thesis, https://doi.org/10.3929/ethz-b-000606500, 2023.
  • [Mor84] Frank Morgan. Area-minimizing currents bounded by higher multiples of curves. Rendiconti del Circolo Matematico di Palermo, 33:37–46, 1984.
  • [MT14] John Morgan and Gang Tian. The geometrization conjecture, volume 5 of Clay Mathematics Monographs. American Mathematical Society, Providence, RI; Clay Mathematics Institute, Cambridge, MA, 2014.
  • [Osi07] Denis V. Osin. Peripheral fillings of relatively hyperbolic groups. Invent. Math., 167(2):295–326, 2007.
  • [Pie19] Erika Pieroni. Minimal entropy of 33-manifolds. arXiv:1902.09190 [math.DG], PhD Thesis, 2019.
  • [RS09] Christian Riedweg and Daniel Schäppi. Singular (lipschitz) homology and homology of integral currents. arXiv preprint arXiv:0902.3831, 2009.
  • [Rua22] Yuping Ruan. The cayley hyperbolic space and volume entropy rigidity. arXiv preprint arXiv:2203.14418, 2022.
  • [Sam99] Andrea Sambusetti. Minimal entropy and simplicial volume. Manuscripta Math., 99(4):541–560, 1999.
  • [Son22] Antoine Song. The spherical Plateau problem for group homology. unpublished preprint arXiv:2202.10636v1 [math.DG], 2022.
  • [Son23] Antoine Song. Entropy and stability of hyperbolic manifolds. 2023. arXiv:2302.07422 [math.DG].
  • [Sor18] Christina Sormani. Intrinsic flat Arzela–Ascoli theorems. Communications in Analysis and Geometry, 26(6):1317–1373, 2018.
  • [Sou01] Juan Souto. Minimal volume and minimal entropy. PhD Thesis, 2001.
  • [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.
  • [Thu97] William P. Thurston. Three-dimensional geometry and topology. Vol. 1, volume 35 of Princeton Mathematical Series. Princeton University Press, Princeton, NJ, 1997. Edited by Silvio Levy.
  • [Wan72] Hsien Chung Wang. Topics on totally discontinuous groups. In Symmetric spaces (Short Courses, Washington Univ., St. Louis, Mo., 1969–1970), pages 459–487. Pure and Appl. Math., Vol. 8. 1972.
  • [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.
  • [Wen14] Stefan Wenger. Plateau’s problem for integral currents in locally non-compact metric spaces. Adv. Calc. Var., 7(2):227–240, 2014.
  • [Whi84] Brian White. The least area bounded by multiples of a curve. Proceedings of the American Mathematical Society, 90(2):230–232, 1984.
  • [Züs23] Roger Züst. Lipschitz rigidigy of Lipschitz manifolds among integral current spaces. arxiv preprint arXiv:2302.07587v1, 2023.