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

Stability and Finiteness of Wasserstein Spaces

Mohammad Alattar Department of Mathematical Sciences, Durham University, United Kingdom mohammad.al-attar@durham.ac.uk
Abstract.

Under Gromov–Hausdorff convergence, and equivariant Gromov–Hausdorff convergence, we prove stability results of Wasserstein spaces over certain classes of singular and non-singular spaces. For example, we obtain an analogue of Perelman’s stability theorem on Wasserstein spaces.

Key words and phrases:
equivariant Gromov–Hausdorff convergence, Gromov–Hausdorff convergence, Wasserstein spaces, stability
2020 Mathematics Subject Classification:
53C23, 53C20, 51K10

1. Introduction

Finiteness theorems for finite dimensional spaces with curvature bounds can be traced back to 1967, with Weinstein [32] proving that given any even natural number nn, and any δ>0\delta>0, there are only finitely many homotopy types of nn-dimensional, δ\delta-pinched manifolds. Shortly after, in 1970, Cheeger obtained several finiteness theorems in his celebrated work [10]. For instance, he proved that given any natural number nn, and real numbers D,v,k>0D,v,k>0, there are only finitely many diffeomorphism types of compact nn-dimensional manifolds XX admitting a Riemannian metric such that diameter(X)(X) D\leq D, vol(X)(X) v\geq v and |sec(X)|k2|\sec(X)|\leq k^{2}, where sec(X)\sec(X) denotes the sectional curvature. These results served as departure points for further seminal and important developments in Riemannian geometry (see for instance [3, 21, 22, 20, 29]).

In 1981, Gromov [19] introduced a powerful notion of convergence, now known as Gromov–Hausdorff convergence, that has been useful in establishing (among other things) finiteness theorems through the notion of “stability”. For instance, using Gromov–Hausdorff convergence, Grove, Petersen and Wu [22], in 1990, relaxed the curvature condition in Cheeger’s finiteness theorem and showed that for any n3n\neq 3, real number kk, and any positive numbers DD and vv, the class consisting of compact nn-dimensional Riemannian manifolds XX with diameter(X)D(X)\leq D, sec(X)k\sec(X)\geq k, and vol(X)v(X)\geq v contains finitely many homeomorphism types.

In 1991, Perelman [29, 25], using Gromov–Hausdorff convergence, showed that given a non-collapsing convergent sequence of compact Alexandrov spaces, with a uniform lower curvature bound, uniform finite dimension and uniform upper diameter bound, the tail of the sequence stabilizes topologically. That is, all spaces in the tail of the sequence are mutually homeomorphic. As a corollary, one obtains homeomorphism finiteness (in all dimensions) for a more general class of spaces that need not be manifolds.

In all cases of finiteness and stability, attention has been restricted to finite dimensional spaces. Thus, the question of stability in infinite dimensions remains open. In this paper, we focus our attention on Wasserstein spaces, which are infinite-dimensional, and prove stability and finiteness results for these spaces in different settings. Our proofs are not technical, and these results are, to the best of our knowledge, the first stability and finiteness results for infinite-dimensional spaces.

Let 𝒜(n,K,D)\mathcal{A}^{\leq}(n,K,D) denote the class of compact Alexandrov spaces with dimension at most nn, curvature bounded below by KK, and diameter bounded above by DD. We denote by 𝒜(n,K,D)\mathcal{A}(n,K,D) all Alexandrov spaces in 𝒜(n,K,D)\mathcal{A}^{\leq}(n,K,D) with dimension nn. Given v>0v>0, we will denote by 𝒜(n,K,D,v)\mathcal{A}(n,K,D,v) the subset of 𝒜(n,K,D)\mathcal{A}(n,K,D) where the Alexandrov spaces under consideration have volume bounded below by vv. Furthermore, given a sequence of compact metric spaces {Xi}i\{X_{i}\}_{i\in\mathbb{N}} and a compact metric space XX, we will denote by XiGHXX_{i}\rightarrow_{GH}X the Gromov–Hausdorff convergence of XiX_{i} to XX. We will denote by 2(X)\mathbb{P}_{2}(X) the Wasserstein space over XX equipped with the 22-Wasserstein metric (see Definition 2.5).

Theorem A.

Assume {Xi}i\{X_{i}\}_{i\in\mathbb{N}} and XX are in 𝒜(n,K,D)\mathcal{A}^{\leq}(n,K,D). Then the following assertions are equivalent:

  1. (1)

    2(Xi)GH2(X)\mathbb{P}_{2}(X_{i})\rightarrow_{GH}\mathbb{P}_{2}(X).

  2. (2)

    XiGHXX_{i}\rightarrow_{GH}X.

  3. (3)

    (2(Xi),Xi)GH(2(X),X)(\mathbb{P}_{2}(X_{i}),X_{i})\rightarrow_{GH}(\mathbb{P}_{2}(X),X).

We point out a few remarks concerning the above theorem. First, the interesting implication is (1)(2)(1)\implies(2). The implication (2)(1)(2)\implies(1) is well known (see Corollary 4.3 in [28]) and holds for general compact metric spaces. Second, the notion of convergence in (3)(3) is known as Gromov–Hausdorff convergence of metric pairs. It was introduced in [9], and later studied extensively in [1]. This notion of convergence generalizes Gromov–Hausdorff convergence and takes into account pairs of the form (X,A)(X,A), where XX is a metric space, and AA is a closed non-empty subset of XX. In our theorem, we identify the base space XX with the set of Dirac deltas in the Wasserstein space 2(X)\mathbb{P}_{2}(X). Thus, (2)(3)(2)\implies(3) becomes trivial. However, the new idea for this implication is that we give an equivalent characterization of convergence of metric pairs using the notion of ϵ\epsilon-isometries (see Proposition 2.1). This equivalent characterization will be useful when we introduce the notion of relative equivariant Gromov–Hausdorff convergence (see Definition 2.2) which takes into account a triple (X,A,G)(X,A,G), where XX is a compact metric space, AA is a closed, non-empty GG-invariant subset of XX and GG is a closed subgroup of Isom(X)\operatorname{Isom}(X), the group of isometries of XX.

A consequence of the above theorem, is the following analogue of Perelman’s stability theorem for Wasserstein spaces.

Corollary B.

Let {Xi}i\{X_{i}\}_{i\in\mathbb{N}} and XX be spaces in 𝒜(n,K,D)\mathcal{A}(n,K,D). If 2(Xi)GH2(X)\mathbb{P}_{2}(X_{i})\rightarrow_{GH}\mathbb{P}_{2}(X) then for all sufficiently large ii, 2(Xi)\mathbb{P}_{2}(X_{i}) and 2(X)\mathbb{P}_{2}(X) are homeomorphic, and the homeomorphisms can be taken to be Gromov–Hausdorff approximations.

Note that in the preceding corollary we do not impose any other structure on the Wasserstein spaces other than that coming from the base spaces.

We further have the following finiteness result. For any nn\in\mathbb{N}, K,D>0K\in\mathbb{R},D>0 and v>0v>0, let W2(n,K,D,v)W_{2}(n,K,D,v) denote the class consisting of Wasserstein spaces with base spaces in 𝒜(n,K,D,v)\mathcal{A}(n,K,D,v) and equipped with the 22-Wasserstein metric.

Corollary C.

There are only finitely many topological types in W2(n,K,D,v)W_{2}(n,K,D,v).

We note that the technique in this paper can be applied whenever we have a stability result on the base spaces and essentially reduces to the following result.

Theorem D.

Let 𝔛\mathfrak{X} denote a precompact class (with respect to the Gromov–Hausdorff topology) of compact non-branching metric measure spaces having good transport behavior (GTB)(GTB) and their reference measures are qualitivatively non-degenerate. Assume that if {Xi}i=1\{X_{i}\}_{i=1}^{\infty} is a sequence in 𝔛\mathfrak{X}, and XiGHXX_{i}\rightarrow_{GH}X, then XX is a non-branching compact metric space admitting a measure that is qualitatively non-degenerate and has good transport behavior (we will take such a measure as a reference measure on XX). Then, the following assertions are equivalent:

  1. (1)

    2(Xi)GH2(X)\mathbb{P}_{2}(X_{i})\rightarrow_{GH}\mathbb{P}_{2}(X).

  2. (2)

    XiGHXX_{i}\rightarrow_{GH}X.

The notion of good transport behavior was first introduced in the paper by Galaz-García, Kell, Mondino, and Sosa [18], and further studied by Kell [26]. The class of metric measure spaces satisfying such condition is large. For instance, according to [18], strong CD(K,N)\mathrm{CD}^{*}(K,N) and essentially non-branching MCP(K,N)\mathrm{MCP}(K,N) spaces admit good transport behavior. The notion of a measure being qualitatively non-degenerate was introduced by Cavalletti and Huesmann [7] and later studied by Kell [26]. It is technical to state, and we refer the reader to [26, 7, 30] for more details, but we note a measure being qualitatively non-degenerate yields desirable information. For example, a qualitatively non-degenerate measure is doubling (Lemma 2.8 in [30]).

We note that the classes 𝒜(n,K,D)\mathcal{A}^{\leq}(n,K,D), 𝒜(n,K,D)\mathcal{A}(n,K,D) satisfy the hypotheses of Theorem D. However, the properties in Theorem D are not closed under Gromov–Hausdorff convergence in general. This is because a compact metric measure space endowed with a qualitatively non-degenerate measure is doubling and thus must have finite dimension.

Using almost verbatim the proofs of Theorems D and A, and noting that RCD(K,N)\mathrm{RCD}(K,N) spaces are non-branching [14], we obtain the following corollary.

Corollary E.

Assume each (Xi,di,𝔪i)(X_{i},d_{i},\mathfrak{m}_{i}) is an compact RCD(K,N)\mathrm{RCD}(K,N) space with uniform upper diameter bound. Assume (X,d)(X,d) is an NN-dimensional compact Riemannian manifold. If 2(Xi)GH2(X)\mathbb{P}_{2}(X_{i})\rightarrow_{GH}\mathbb{P}_{2}(X) then for all large ii, there exists a homeomorphism Gi:2(Xi)2(X)G_{i}\colon\mathbb{P}_{2}(X_{i})\rightarrow\mathbb{P}_{2}(X) that is Lipschitz and its inverse is Hölder.

A consequence of our next main theorem is an equivariant stability result for Wasserstein spaces. We first discuss notation. Given a closed subgroup GG of Isom(X)\operatorname{Isom}(X), where XX is a compact metric space, we denote by G#G^{\#} the induced subgroup of Isom(2(X))\operatorname{Isom}(\mathbb{P}_{2}(X)) given by the push-forward of maps. Given a sequence of pairs {(Xi,Gi)}i\{(X_{i},G_{i})\}_{i\in\mathbb{N}}, where each XiX_{i} is a compact metric space and GiG_{i} is a closed subgroup of Isom(X)\operatorname{Isom}(X), we will denote, by (Xi,Gi)eGH(X,G)(X_{i},G_{i})\rightarrow_{eGH}(X,G) the equivariant Gromov–Hausdorff convergence of (Xi,Gi)(X_{i},G_{i}) to (X,G)(X,G). This convergence, originally introduced by Fukaya (see [16, 17, 15]), has been useful in establishing interesting results in both the singular and non-singular setting (see for instance [34, 12, 5, 23, 33, 13, 8]). Furthermore, it has been useful in obtaining finiteness and stability results in the singular setting. For instance, using equivariant Gromov–Hausdorff convergence, Zamora extended Anderson finiteness to the RCD(K,N)\mathrm{RCD}^{*}(K,N) setting [33].

Theorem F.

Let {Xi}i\{X_{i}\}_{i\in\mathbb{N}} and XX be closed Riemannian manifolds with uniform lower sectional curvature bound K>0K>0, uniform upper diameter bound and uniform upper dimension bound. Assume HiH_{i} and HH are closed subgroups of Isom(2(Xi))\operatorname{Isom}(\mathbb{P}_{2}(X_{i})) and Isom(2(X))\operatorname{Isom}(\mathbb{P}_{2}(X)) respectively. Then the following assertions are equivalent:

  1. (1)

    (2(Xi),Hi)eGH(2(X),H)(\mathbb{P}_{2}(X_{i}),H_{i})\rightarrow_{eGH}(\mathbb{P}_{2}(X),H).

  2. (2)

    There exists unique closed subgroups GiG_{i} of Isom(X)\operatorname{Isom}(X) and GG of Isom(X)\operatorname{Isom}(X) such that Gi#=HiG_{i}^{\#}=H_{i} and G#=HG^{\#}=H and such that (Xi,Gi)eGH(X,G)(X_{i},G_{i})\rightarrow_{eGH}(X,G).

  3. (3)

    (2(Xi),Xi,Hi)eGH(2(X),X,H)(\mathbb{P}_{2}(X_{i}),X_{i},H_{i})\rightarrow_{eGH}(\mathbb{P}_{2}(X),X,H).

We emphasize that we identify the base space with the (closed) subset of Dirac deltas in the Wasserstein space. In the presence of a uniform lower curvature bound, we have the following corollary, an equivariant stability result on Wasserstein spaces.

Corollary G.

Let {Xi}i\{X_{i}\}_{i\in\mathbb{N}} and XX be closed Riemannian manifolds with uniform lower sectional curvature bound K>0K>0, uniform upper diameter bound and uniform dimension. Assume HiH_{i} and HH are closed subgroups of Isom(2(Xi))\operatorname{Isom}(\mathbb{P}_{2}(X_{i})) and Isom(2(X))\operatorname{Isom}(\mathbb{P}_{2}(X)) respectively of the same dimension. If (2(Xi),Hi)eGH(2(X),H)(\mathbb{P}_{2}(X_{i}),H_{i})\rightarrow_{eGH}(\mathbb{P}_{2}(X),H), then for all large ii, there exists a homeomorphism fi:2(Xi)2(X)f_{i}\colon\mathbb{P}_{2}(X_{i})\rightarrow\mathbb{P}_{2}(X) that is compatible with some group isomorphism χi:HiH\chi_{i}\colon H_{i}\rightarrow H.

Our article is organized as follows. In section 2, we discuss the preliminaries that will be useful in this paper. Moreover, we will give a new characterization of convergence of metric pairs. In section 3 we give proofs of the theorems and corollaries.

Acknowledgements.

It is a pleasure to thank Fernando Galaz–García, Martin Kerin, Kohei Suzuki, Jaime Santos-Rodríguez, and Mauricio Che for helpful conversations and comments during the Durham Metric Geometry Reading Seminar. I would also like to thank Sergio Zamora and Yanpeng Zhi for several interesting conversations and Aseel alnajjar for several comments on the first draft of this manuscript. Part of this work was completed at the Erwin Schrödinger International Institute for Mathematics and Physics (ESI), where the author received financial support to participate in the workshop Synthetic Curvature Bounds for Non-Smooth Spaces: Beyond Finite Dimension. The author would like to thank ESI and the conference organizers for the excellent atmosphere and working conditions.

2. Preliminaries

In this section, we collect preliminary definitions and results we will use in the proof of our main results.

2.1. Convergence of Metric Pairs and Equivariant Convergence of Metric Pairs

In this section, we give a new equivalent characterization of the notion of convergence of metric pairs which will allow us to introduce, canonically, the notion of relative equivariant Gromov–Hausdorff convergence. The equivalent characterization is demonstrated in the following proposition. We refer the reader to the paper by Ahumada Gómez and Che [1] for results concerning the (non-equivariant) convergence of metric pairs.

Proposition 2.1.

Let {Xi}i\{X_{i}\}_{i\in\mathbb{N}} be a sequence of compact metric spaces. Let {Ai}i\{A_{i}\}_{i\in\mathbb{N}} be a collection of closed non-empty subspaces AiXiA_{i}\subseteq X_{i}. Let XX be a compact metric space, and AXA\subseteq X a closed non-empty subset. Then the following assertions are equivalent:

  1. (1)

    (Xi,Ai)GH(X,A)(X_{i},A_{i})\rightarrow_{GH}(X,A).

  2. (2)

    There exists a sequence of positive numbers {ϵi}i\{\epsilon_{i}\}_{i\in\mathbb{N}} such that ϵi0\epsilon_{i}\rightarrow 0 and ϵi\epsilon_{i}-Gromov–Hausdorff approximations fi:XiXf_{i}\colon X_{i}\rightarrow X, gi:AiAg_{i}\colon A_{i}\rightarrow A such that fif_{i} and gig_{i} are ϵi\epsilon_{i}-close, i.e., dX(fi(xi),gi(xi))ϵid_{X}(f_{i}(x_{i}),g_{i}(x_{i}))\leq\epsilon_{i} for all xiAix_{i}\in A_{i}.

Proof.

Assume (Xi,Ai)GH(X,A)(X_{i},A_{i})\rightarrow_{GH}(X,A). Then there exists a sequence of numbers {ϵi}i\{\epsilon_{i}\}_{i\in\mathbb{N}} that tends to 0 and ϵi\epsilon_{i}-Gromov–Hausdorff approximations fi:XiXf_{i}\colon X_{i}\rightarrow X such that dH(fi(Ai),A)<ϵid_{H}(f_{i}(A_{i}),A)<\epsilon_{i}. Here, dHd_{H} denotes the Hausdorff distance. We define gi:AiAg_{i}\colon A_{i}\rightarrow A as follows. Since dH(fi(Ai),A)<ϵid_{H}(f_{i}(A_{i}),A)<\epsilon_{i}, for each aiAia_{i}\in A_{i} choose gi(ai)Ag_{i}(a_{i})\in A such that dX(fi(ai),gi(ai))<ϵid_{X}(f_{i}(a_{i}),g_{i}(a_{i}))<\epsilon_{i}. We will show that the map gi:AiAg_{i}\colon A_{i}\rightarrow A is a 3ϵi3\epsilon_{i}-Gromov–Hausdorff approximation. Indeed, for each aAa\in A, there exists an aiAia_{i}\in A_{i} such that dX(fi(ai),a)ϵd_{X}(f_{i}(a_{i}),a)\leq\epsilon. Therefore, by the triangle inequality, we get dX(gi(ai),a)3ϵid_{X}(g_{i}(a_{i}),a)\leq 3\epsilon_{i}. Now we show that gig_{i} is almost distance preserving up to an error 3ϵi3\epsilon_{i}. We only show one side of the inequality, for the other is similar. Let ai,aiAia_{i},a_{i}^{\prime}\in A_{i}. Then

dX(gn(ai),gn(ai))dX(gi(ai),fi(ai))+dX(fi(ai),fi(ai))+dX(fi(ai),gi(ai)).d_{X}(g_{n}(a_{i}),g_{n}(a_{i}^{\prime}))\leq d_{X}(g_{i}(a_{i}),f_{i}(a_{i}))+d_{X}(f_{i}(a_{i}),f_{i}(a_{i}^{\prime}))+d_{X}(f_{i}(a_{i}^{\prime}),g_{i}(a_{i}^{\prime})).

Now note that dX(gi(ai),fi(ai))ϵid_{X}(g_{i}(a_{i}),f_{i}(a_{i}))\leq\epsilon_{i} and dX(fi(ai),gi(ai))ϵid_{X}(f_{i}(a_{i}^{\prime}),g_{i}(a_{i}^{\prime}))\leq\epsilon_{i}. Thus, as fif_{i} is an ϵi\epsilon_{i}-Gromov–Hausdorff approximation, we get

dX(gi(ai),gi(ai))3ϵi+dXi(ai,ai).d_{X}(g_{i}(a_{i}),g_{i}(a_{i}^{\prime}))\leq 3\epsilon_{i}+d_{X_{i}}(a_{i},a_{i}^{\prime}).

Now we show the reverse direction. Define ξi:XiX\xi_{i}\colon X_{i}\rightarrow X by

ξi(xi)={fi(xi)xiXi\Aigi(xi)xiAi.\xi_{i}(x_{i})=\begin{cases}f_{i}(x_{i})&x_{i}\in X_{i}\backslash A_{i}\\ g_{i}(x_{i})&x_{i}\in A_{i}.\\ \end{cases}

Since gi:AiAg_{i}\colon A_{i}\rightarrow A is an ϵi\epsilon_{i}-Gromov–Hausdorff approximation, it follows that dH(gi(Ai),A)ϵid_{H}(g_{i}(A_{i}),A)\leq\epsilon_{i}. It remains to verify that ξi\xi_{i} is a Gromov–Hausdorff approximation. It is clear that ξi\xi_{i} is ϵi\epsilon_{i}-surjective. Therefore, it remains to prove that ξi\xi_{i} is almost distance preserving up to an error that tends to 0. Once again, we only verify one direction of the inequality, since the other is similar. Let xi,xiXix_{i},x_{i}^{\prime}\in X_{i}. It is clear that it suffices to assume that xiXi\Aix_{i}\in X_{i}\backslash A_{i} and that xiAix_{i}^{\prime}\in A_{i}. Then

dX(ξi(xi),ξi(xi))=dX(fi(xi),gi(xi))dX(fi(xi),fi(xi))+dX(fi(xi),gi(xi)).d_{X}(\xi_{i}(x_{i}),\xi_{i}(x_{i}^{\prime}))=d_{X}(f_{i}(x_{i}),g_{i}(x_{i}^{\prime}))\leq d_{X}(f_{i}(x_{i}),f_{i}(x_{i}^{\prime}))+d_{X}(f_{i}(x_{i}^{\prime}),g_{i}(x_{i}^{\prime})).

Since fif_{i} is an ϵi\epsilon_{i}-Gromov–Hausdorff approximation, dX(fi(xi),fi(xi))ϵi+dXi(xi,xi)d_{X}(f_{i}(x_{i}),f_{i}(x_{i}^{\prime}))\leq\epsilon_{i}+d_{X_{i}}(x_{i},x_{i}^{\prime}). Since fif_{i} and gig_{i} are ϵi\epsilon_{i}-close, dX(fi(xi),gi(xi))ϵid_{X}(f_{i}(x_{i}^{\prime}),g_{i}(x_{i}^{\prime}))\leq\epsilon_{i}. ∎

The proposition above suggests the following definition of relative equivariant Gromov–Hausdorff convergence.

Definition 2.2.

Let XX and YY be compact metric spaces. Suppose GXG_{X} and GYG_{Y} are closed subgroups of Isom(X)\operatorname{Isom}(X) and Isom(Y)\operatorname{Isom}(Y) respectively. Assume AXA_{X} is a closed non-empty GXG_{X}-invariant subset of XX, and AYA_{Y} is a closed non-empty GYG_{Y}-invariant subset of YY. For ϵ>0\epsilon>0, we define an ϵ\epsilon-equivariant Gromov–Hausdorff approximation between (X,AX,GX)(X,A_{X},G_{X}) and (Y,AY,GY)(Y,A_{Y},G_{Y}) to be a quadruple of maps

(f:XY,f:AXAY,θ:GXGY,ψ:GYGX)(f\colon X\rightarrow Y,f^{\prime}\colon A_{X}\rightarrow A_{Y},\theta\colon G_{X}\rightarrow G_{Y},\psi\colon G_{Y}\rightarrow G_{X})

subject to the following conditions:

  1. (1)

    The triple (f,θ,ψ)(f,\theta,\psi) is an ϵ\epsilon-equivariant Gromov–Hausdorff approximation (in the usual sense).

  2. (2)

    ff^{\prime} is an ϵ\epsilon-Gromov–Hausdorff approximation.

  3. (3)

    The maps ff and ff^{\prime} are ϵ\epsilon-close.

Remark 2.3.

We use the terms “equivariant Gromov–Hausdorff approximations” and “equivariant approximations” interchangeably.

Definition 2.4.

Given a sequence of triples {(Xi,Ai,Gi)}i\{(X_{i},A_{i},G_{i})\}_{i\in\mathbb{N}}, where XiX_{i} is a compact metric space, GiG_{i} is a closed subgoup of Isom(Xi)\operatorname{Isom}(X_{i}) and AiA_{i} is a closed non-empty GiG_{i}-invariant subset of XiX_{i}, we say that (Xi,Ai,Gi)(X_{i},A_{i},G_{i}) Gromov–Hausdorff equivariantly converges to (X,A,G)(X,A,G), denoted (Xi,Ai,Gi)eGH(X,A,G)(X_{i},A_{i},G_{i})\rightarrow_{eGH}(X,A,G), where XX is a compact metric space, GG is a closed subgroup of Isom(X)(X) and AA is a closed subset of XX that is GG-invariant, if there exists a sequence of positive real numbers {ϵi}i\{\epsilon_{i}\}_{i\in\mathbb{N}} such that ϵi0\epsilon_{i}\rightarrow 0 and ϵi\epsilon_{i}-equivariant Gromov–Hausdorff approximations (fi,fi,θi,ψi):(Xi,Ai,Gi)(X,A,G)(f_{i},f^{\prime}_{i},\theta_{i},\psi_{i})\colon(X_{i},A_{i},G_{i})\rightarrow(X,A,G).

2.2. Further definitions and auxiliary results

For the convenience of the reader, we include the definitions of Wasserstein spaces, good transport behavior and isometric rigidity.

Definition 2.5.

Let (X,d)(X,d) be a Polish metric space and let p[1,)p\in[1,\infty). Given any two probability measure μ\mu and ν\nu on XX, we define the LpL^{p}-Wasserstein metric between μ\mu and ν\nu, denoted by WpW_{p}, to be

Wpp(μ,ν)=infπd(x,y)p𝑑π(x,y),W_{p}^{p}(\mu,\nu)=\inf_{\pi}\int d(x,y)^{p}d\pi(x,y),

where the infimum is taken over all admissible measures π\pi having marginals μ\mu and ν\nu. The LpL^{p}-Wasserstein metric is a metric on the space of probability measures with finite pp-moments, which is denoted by p(X)\mathbb{P}_{p}(X).

Definition 2.6.

A metric measure space (X,d,𝔪)(X,d,\mathfrak{m}) is said to have good transport behavior if, for any two probability measures μ,ν\mu,\nu in 2(X)\mathbb{P}_{2}(X), where μ𝔪\mu\ll\mathfrak{m}, any optimal transport map between them is induced by a map.

We motivate the concept of isometric rigidity by the following example.

Example 2.7.

If XX is a compact metric space and f:XXf:X\rightarrow X is an isometry, then ff induces an isometry f#:2(X)2(X)f_{\#}:\mathbb{P}_{2}(X)\rightarrow\mathbb{P}_{2}(X).

Thus, it is natural to formulate the following definition.

Definition 2.8.

Let (X,d)(X,d) be a metric space. Then, XX is said to be isometrically rigid if given any isometry θ:2(X)2(X)\theta:\mathbb{P}_{2}(X)\rightarrow\mathbb{P}_{2}(X), there exists an isometry f:XXf:X\rightarrow X such that θ=f#\theta=f_{\#}. In particular, every isometry of 2(X)\mathbb{P}_{2}(X) is induced by an isometry of XX; the base space.

The concept of isometric rigidity was first introduced by Kloeckner in the context of euclidean space in [27], where he showed, among other things, that, 2()\mathbb{P}_{2}(\mathbb{R}) is not isometrically rigid. Later, Bertrand and Kloeckner [4] studied the concept of isometric rigidity and showed that if XX is a negatively curved geodesically complete Hadamard space, then 2(X)\mathbb{P}_{2}(X) is isometrically rigid. In 2021, Santos Rodriguez showed [30] that if XX is a positively curved closed Riemannian manifold, then XX is isometrically rigid and that, if XX is a CROSS, then, in fact, the isometry groups of p(X)\mathbb{P}_{p}(X) and XX are the same for all p>1p>1. In the same paper, Santos Rodriguez showed the following result which we will use frequently in this paper (Corollary 3.8, [30]).

Proposition 2.9.

Let (X,dX,𝔪X)(X,d_{X},\mathfrak{m}_{X}) and (Y,dY,𝔪Y)(Y,d_{Y},\mathfrak{m}_{Y}) be two compact non-branching metric measure spaces equipped with qualitatively non-degenerate measures and such that they have good transport behavior. Suppose that there exists an isometry Φ:2(X)2(Y)\Phi:\mathbb{P}_{2}(X)\rightarrow\mathbb{P}_{2}(Y). Then (X,dX)(X,d_{X}) and (Y,dY)(Y,d_{Y}) are isometric.

In fact, Santos Rodriguez in [30] showed the above proposition for Wasserstein spaces with finite pp-moments, where p(1,)p\in(1,\infty).

3. Proofs

In what follows, we assume that the maps are sufficiently regular. Indeed, this is possible because if XX is a compact metric space then Isom(X)\operatorname{Isom}(X) is compact with the compact-open topology. Moreover, the compact open topology on Isom(X)\operatorname{Isom}(X) can be metrized with the uniform metric. One can always replace ϵ\epsilon-isometries between compact metric spaces by almost-isometries that are measurable. In particular, we have the following simple lemma.

Lemma 3.1.

Assume XX and YY are compact metric spaces and GXG_{X} and GYG_{Y} are closed subgroups of Isom(X)\operatorname{Isom}(X) and Isom(Y)\operatorname{Isom}(Y) respectively. Equip GXG_{X} and GYG_{Y} with the uniform metrics dGXd_{G_{X}} and dGYd_{G_{Y}} respectively. If (f:XY,θ:GXGY,ψ:GYGX)(f\colon X\rightarrow Y,\theta\colon G_{X}\rightarrow G_{Y},\psi\colon G_{Y}\rightarrow G_{X}) is an ϵ\epsilon-equivariant approximation, then, ff, θ\theta and ψ\psi can be chosen to be measurable.

Proof.

By Corollary 3.4 in [2], it follows that θ\theta is an 5ϵ5\epsilon-approximation. Furthermore, observe that ψ\psi is an almost inverse to θ\theta. In particular, for λGY\lambda\in G_{Y}, and gGXg\in G_{X}, one has dGY(θ(ψ(λ)),λ)4ϵd_{G_{Y}}(\theta(\psi(\lambda)),\lambda)\leq 4\epsilon and dGX(ψ(θ(g)),g)4ϵd_{G_{X}}(\psi(\theta(g)),g)\leq 4\epsilon. Hence, as shown in Lemma 4.1 in [11], one can obtain measurable approximations f1:XYf_{1}\colon X\rightarrow Y, θ1:GXGY\theta_{1}\colon G_{X}\rightarrow G_{Y} and ψ1:GYGX\psi_{1}\colon G_{Y}\rightarrow G_{X} that remain close to ff, θ\theta and ψ\psi respectively. Hence the result follows. ∎

Proof of Theorem D

The implication (2)(1)(2)\implies(1) is well known (see Corollary 4.3 in [28]). Let us show (1)(2)(1)\implies(2). Assume for the sake of obtaining a contradiction that 2(Xi)GH2(X)\mathbb{P}_{2}(X_{i})\rightarrow_{GH}\mathbb{P}_{2}(X) holds but XiGHXX_{i}\rightarrow_{GH}X does not hold. Hence, there exists a δ>0\delta>0 such that (up to a subsequence), dGH(Xi,X)δ>0d_{GH}(X_{i},X)\geq\delta>0 for all ii. Since the class 𝔛\mathfrak{X} is pre-compact, and by our assumptions, it follows that up to a further subsequence, XiGHXX_{i}\rightarrow_{GH}X^{\prime}, where XX^{\prime} is a compact non-branching metric space. Hence, 2(Xi)GH2(X)\mathbb{P}_{2}(X_{i})\rightarrow_{GH}\mathbb{P}_{2}(X^{\prime}). By assumption, 2(X)\mathbb{P}_{2}(X) and 2(X)\mathbb{P}_{2}(X^{\prime}) are isometric. Now, by Proposition 2.9, XX and XX^{\prime} are therefore isometric, which is a contradiction. ∎

Proof of Theorem A

The proof is similar to the proof of Theorem D. However, we give details for the convenience of the reader. The implications (2)(1)(2)\implies(1) and (3)(1)(3)\implies(1) are clear. The implication (2)(3)(2)\implies(3) follows at once from the following observation. Let fi:XiXf_{i}\colon X_{i}\rightarrow X be ϵi\epsilon_{i}-approximations by measurable maps. Then each fif_{i} induces an ϵi~\tilde{\epsilon_{i}}-approximation (fi)#:2(Xi)2(X)(f_{i})_{\#}\colon\mathbb{P}_{2}(X_{i})\rightarrow\mathbb{P}_{2}(X), with ϵi~0\tilde{\epsilon_{i}}\rightarrow 0, and such that, on Dirac deltas, (fi)#(f_{i})_{\#} is a Hausdorff approximation. Hence, one may use Proposition 2.1 to establish this implication. We shall now show (1)(2)(1)\implies(2).

We must prove that XiGHXX_{i}\rightarrow_{GH}X. Assume otherwise. Hence, there exists δ>0\delta>0 such that, up to a subsequence, dGH(Xi,X)δ>0d_{GH}(X_{i},X)\geq\delta>0 for all ii. Here, dGHd_{GH} denotes the Gromov–Hausdorff distance. Since 𝒜(n,K,D)\mathcal{A}^{\leq}(n,K,D) is, with respect to dGHd_{GH}, compact [6], there exists a compact Alexandrov space XX^{\prime} such that, up to a further subsequence, XiGHXX_{i}\rightarrow_{GH}X^{\prime}. Hence, 2(Xi)GH2(X)\mathbb{P}_{2}(X_{i})\rightarrow_{GH}\mathbb{P}_{2}(X^{\prime}). By assumption, 2(Xi)GH2(X)\mathbb{P}_{2}(X_{i})\rightarrow_{GH}\mathbb{P}_{2}(X). Therefore, 2(X)\mathbb{P}_{2}(X^{\prime}) and 2(X)\mathbb{P}_{2}(X) are isometric. Alexandrov spaces have good transport behavior, and on them, the Hausdorff measure is qualitatively non-degenerate, therefore XX and XX^{\prime} are isometric. Hence we have XiGHXX_{i}\rightarrow_{GH}X. ∎

Proof of Corollary B

The result follows from Perelman’s stability Theorem. Note that if f:XYf\colon X\rightarrow Y is a map, then f#f_{\#} is continuous if and only if ff is continuous (see Remark 4.14 in [31]). ∎

Proof of Corollary C

We proceed by contradiction. Assume we have a sequence (2(Xi))i(\mathbb{P}_{2}(X_{i}))_{i\in\mathbb{N}} of topologically pairwise inequivalent Wasserstein spaces in W2(n,K,D,v)W_{2}(n,K,D,v). Then, since 𝒜(n,K,D,v)\mathcal{A}(n,K,D,v) is precompact and we have volume bounded below by v>0v>0, there exists a compact nn-dimensional Alexandrov space XX such that (up to a subsequence), XiGHXX_{i}\rightarrow_{GH}X. Hence, 2(Xi)GH2(X)\mathbb{P}_{2}(X_{i})\rightarrow_{GH}\mathbb{P}_{2}(X). Therefore, by Corollary B, for all large ii, 2(Xi)\mathbb{P}_{2}(X_{i}) and 2(X)\mathbb{P}_{2}(X) are homeomorphic, which is a contradiction.

Proof of Corollary E

This follows in an analogous manner as in the proof of Corollary B. However, one uses the result in [24] instead of Perelman’s stability theorem. ∎

Now we will prove Theorem F. First, we need two lemmas.

Lemma 3.2.

Let XX and YY be compact metric spaces and GXG_{X} and GYG_{Y} are closed subgroups of Isom(X)\operatorname{Isom}(X) and Isom(Y)\operatorname{Isom}(Y) respectively. An ϵ\epsilon-equivariant approximation, (f,θ,ψ)(f,\theta,\psi) between (X,GX)(X,G_{X}) and (Y,GY)(Y,G_{Y}) induces an ϵ~\tilde{\epsilon}-equivariant approximation (f#,θ#,ψ#)(f_{\#},\theta_{\#},\psi_{\#}) between (2(X),GX#)(\mathbb{P}_{2}(X),G_{X}^{\#}) and (2(Y),GY#)(\mathbb{P}_{2}(Y),G_{Y}^{\#}), where ϵ~\tilde{\epsilon} tends to 0 as ϵ\epsilon tends to 0.

Proof.

Define θ#\theta_{\#} and ψ#\psi_{\#} by the rules θ#(g#)=(θ(g))#\theta_{\#}(g_{\#})=(\theta(g))_{\#} and ψ#(λ#)=(ψ(λ))#\psi_{\#}(\lambda_{\#})=(\psi(\lambda))_{\#}. By Corollary 4.3 in [28], the map f#f_{\#} is an ϵ~\tilde{\epsilon}-approximation, where

ϵ~=6ϵ+ϵ(2diam(Y)+ϵ).\tilde{\epsilon}=6\epsilon+\sqrt{\epsilon(2\operatorname{diam}(Y)+\epsilon)}.

Let μ2(X)\mu\in\mathbb{P}_{2}(X). Then,

W22(θ#(g#)(f#μ),f#(g#μ))d2(θ(g)(f(x)),f(gx))𝑑μϵ2.W_{2}^{2}(\theta_{\#}(g_{\#})(f_{\#}\mu),f_{\#}(g_{\#}\mu))\leq\int d^{2}(\theta(g)(f(x)),f(gx))d\mu\leq\epsilon^{2}.

The last inequality is similar. ∎

As a corollary, we have the following result.

Corollary 3.3.

Let XiX_{i} and XX be compact metric spaces. If (Xi,Gi)eGH(X,G)(X_{i},G_{i})\rightarrow_{eGH}(X,G), then (2(Xi),Gi#)eGH(2(X),G#)(\mathbb{P}_{2}(X_{i}),G_{i}^{\#})\rightarrow_{eGH}(\mathbb{P}_{2}(X),G^{\#}).

We will also need the following lemma, an equivariant analogue of Proposition 2.9.

Lemma 3.4.

Let XX and YY be positively curved closed Riemannian manifolds and let GXG_{X} and GYG_{Y} be closed subgroups of Isom(X)\operatorname{Isom}(X) and Isom(Y)\operatorname{Isom}(Y), respectively. If (2(X),GX#)(\mathbb{P}_{2}(X),G_{X}^{\#}) and (2(Y),GY#)(\mathbb{P}_{2}(Y),G_{Y}^{\#}) are equivariantly isometric then so are (X,GX)(X,G_{X}) and (Y,GY)(Y,G_{Y}).

Proof.

Let Φ:2(X)2(Y)\Phi\colon\mathbb{P}_{2}(X)\rightarrow\mathbb{P}_{2}(Y) be an isometry and let Θ:GX#GY#\Theta:G_{X}^{\#}\rightarrow G_{Y}^{\#} be an isomorphism such that for any gIsom(X)g\in\operatorname{Isom}(X), one has Φg#=Θ(g#)Φ\Phi\circ g_{\#}=\Theta(g_{\#})\circ\Phi. Then, from Corollary 3.8 in [30], it follows that Φ\Phi sends Dirac deltas onto Dirac deltas. Thus, there exists an isometry F:XYF\colon X\rightarrow Y such that F#F_{\#} and Φ\Phi agree on Dirac deltas. Moreover, there exists an isomorphism θ:GXGY\theta\colon G_{X}\rightarrow G_{Y} such that for gGXg\in G_{X}, θ(g)#=Θ(g#)\theta(g)_{\#}=\Theta(g_{\#}). Thus it follows that Fg=θ(g)FF\circ g=\theta(g)\circ F. ∎

We are now ready to prove theorem F.

Proof of Theorem F

The implication (2)(1)(2)\implies(1) follows from Lemma 3.2. The implication (2)(3)(2)\implies(3) is trivial. Now we shall verify the implication (1)(2).(1)\implies(2). Since XiX_{i} and XX are isometrically rigid, there exists unique closed subgroups GiG_{i} of Isom(Xi)\operatorname{Isom}(X_{i}) and GG of Isom(X)\operatorname{Isom}(X) such that Gi#=HiG_{i}^{\#}=H_{i} and G#=HG^{\#}=H. It now remains to verify that (Xi,Gi)eGH(X,G)(X_{i},G_{i})\rightarrow_{eGH}(X,G). As usual, assume otherwise. So, there exists a δ>0\delta>0 such that, up to a subsequence, deGH((Xi,Gi),(X,G))δd_{eGH}((X_{i},G_{i}),(X,G))\geq\delta for all ii. Here, deGHd_{eGH} denotes the equivariant Gromov–Hausdorff distance. Since {Xi}i\{X_{i}\}_{i} forms a family of compact Alexandrov spaces with curvature uniformly bounded below, dimension uniformly bounded above, and uniform upper diameter bound, it follows that there exists a compact Alexandrov space XX^{\prime}, such that, up to a subsequence, XiGHXX_{i}\rightarrow_{GH}X^{\prime}. Up to a further subsequence, there exists a closed subgroup GG^{\prime} of Isom(X)\operatorname{Isom}(X^{\prime}) such that (Xi,Gi)eGH(X,G)(X_{i},G_{i})\rightarrow_{eGH}(X^{\prime},G^{\prime}). Hence, by Lemma 3.2, it follows that (2(Xi),Hi)eGH(2(X),(G)#)(\mathbb{P}_{2}(X_{i}),H_{i})\rightarrow_{eGH}(\mathbb{P}_{2}(X^{\prime}),(G^{\prime})^{\#}). Therefore, (2(X),(G)#)(\mathbb{P}_{2}(X^{\prime}),(G^{\prime})^{\#}) and (2(X),G#)(\mathbb{P}_{2}(X),G^{\#}) are equivariantly isometric. The previous lemma shows that (X,G)(X^{\prime},G^{\prime}) and (X,G)(X,G) are equivariantly isometric, which is a contradiction. ∎

Proof of Corollary G

Theorem F ensures there exists closed subgroups GiG_{i} and GG such that Gi#=HiG_{i}^{\#}=H_{i}, G#=HG^{\#}=H and such that (Xi,Gi)eGH(X,G)(X_{i},G_{i})\rightarrow_{eGH}(X,G). Note that the isometric rigidity of the XiX_{i} and XX ensures that GiG_{i} and GG have the same dimension as HiH_{i} and HH. Hence the result follows from Theorem A in [2]. ∎

References

  • [1] Andrés Ahumada Gómez and Mauricio Che “Gromov–Hausdorff convergence of metric pairs and metric tuples” In Differential Geom. Appl. 94, 2024, pp. Paper No. 102135 DOI: 10.1016/j.difgeo.2024.102135
  • [2] Mohammad Alattar “Stability and equivariant Gromov–Hausdorff convergence” In Bulletin of the London Mathematical Society 47, 2024 DOI: https://doi.org/10.1112/blms.13073
  • [3] Michael T. Anderson and Jeff Cheeger “Diffeomorphism finiteness for manifolds with Ricci curvature and Ln/2L^{n/2}-norm of curvature bounded” In Geom. Funct. Anal. 1.3, 1991, pp. 231–252 DOI: 10.1007/BF01896203
  • [4] Jérôme Bertrand and Benoît R. Kloeckner “A geometric study of Wasserstein spaces: isometric rigidity in negative curvature” In Int. Math. Res. Not. IMRN, 2016, pp. 1368–1386 DOI: 10.1093/imrn/rnv177
  • [5] Elia Brue, Aaron Naber and Daniele Semola “Stability of Tori under Lower Sectional Curvature”, 2023 arXiv:2307.03824 [math.DG]
  • [6] Yuri. Burago, Misha Gromov and G. Perelman “A. D. Aleksandrov spaces with curvatures bounded below” In Uspekhi Mat. Nauk 47.2(284), 1992, pp. 3–51\bibrangessep222 DOI: 10.1070/RM1992v047n02ABEH000877
  • [7] Fabio Cavalletti and Martin Huesmann “Existence and uniqueness of optimal transport maps” In Ann. Inst. H. Poincaré C Anal. Non Linéaire 32.6, 2015, pp. 1367–1377 DOI: 10.1016/j.anihpc.2014.09.006
  • [8] Nicola Cavallucci “A GH-compactification of CAT(0)(0)-groups via totally disconnected, unimodular actions”, 2023 arXiv:2307.05640 [math.MG]
  • [9] Mauricio Che, Fernando Galaz-García, Luis Guijarro and Ingrid Membrillo Solis “Metric Geometry of Spaces of Persistence Diagrams”, 2022 arXiv:2109.14697 [math.MG]
  • [10] Jeff Cheeger “Finiteness theorems for Riemannian manifolds” In Amer. J. Math. 92, 1970, pp. 61–74 DOI: 10.2307/2373498
  • [11] Nhan-Phu Chung “Gromov-Hausdorff distances for dynamical systems” In Discrete Contin. Dyn. Syst. 40.11, 2020, pp. 6179–6200 DOI: 10.3934/dcds.2020275
  • [12] Diego Corro “Collapsing regular Riemannian foliations with flat leaves”, 2024 arXiv:2403.11602 [math.DG]
  • [13] Diego Corro, Jaime Santos-Rodríguez and Jesús Núñez-Zimbrón “Cohomogeneity one RCD-spaces”, 2024 arXiv:2405.09448 [math.MG]
  • [14] Qin Deng “Hölder continuity of tangent cones in RCD(K,N) spaces and applications to non-branching”, 2020 arXiv:2009.07956 [math.DG]
  • [15] Kenji Fukaya “Hausdorff convergence of Riemannian manifolds and its applications” In Recent topics in differential and analytic geometry 18, Adv. Stud. Pure Math. Academic Press, Boston, MA, 1990, pp. 143–238 DOI: 10.2969/aspm/01810143
  • [16] Kenji Fukaya “Theory of convergence for Riemannian orbifolds” In Japan. J. Math. (N.S.) 12.1, 1986, pp. 121–160 DOI: 10.4099/math1924.12.121
  • [17] Kenji Fukaya and Takao Yamaguchi “The fundamental groups of almost non-negatively curved manifolds” In Ann. of Math. (2) 136.2, 1992, pp. 253–333 DOI: 10.2307/2946606
  • [18] Fernando Galaz-García, Martin Kell, Andrea Mondino and Gerardo Sosa “On quotients of spaces with Ricci curvature bounded below” In J. Funct. Anal. 275.6, 2018, pp. 1368–1446 DOI: 10.1016/j.jfa.2018.06.002
  • [19] Mikhael Gromov “Groups of polynomial growth and expanding maps” In Inst. Hautes Études Sci. Publ. Math., 1981, pp. 53–73 URL: http://www.numdam.org/item?id=PMIHES_1981__53__53_0
  • [20] Karsten Grove “Finiteness theorems in Riemannian geometry” In Explorations in complex and Riemannian geometry 332, Contemp. Math. Amer. Math. Soc., Providence, RI, 2003, pp. 101–120 DOI: 10.1090/conm/332/05931
  • [21] Karsten Grove and Peter Petersen “Bounding homotopy types by geometry” In Ann. of Math. (2) 128.1, 1988, pp. 195–206 DOI: 10.2307/1971439
  • [22] Karsten Grove, Peter Petersen and Jyh Yang Wu “Geometric finiteness theorems via controlled topology” In Invent. Math. 99.1, 1990, pp. 205–213 DOI: 10.1007/BF01234417
  • [23] John Harvey “Convergence of isometries, with semicontinuity of symmetry of Alexandrov spaces” In Proceedings of the American Mathematical Society 144.8, 2016, pp. 3507–3515
  • [24] Shouhei Honda and Yuanlin Peng “A note on the topological stability theorem from RCD spaces to Riemannian manifolds” In Manuscripta Math. 172.3-4, 2023, pp. 971–1007 DOI: 10.1007/s00229-022-01418-7
  • [25] Vitali Kapovitch “Perelman’s stability theorem” In Surveys in differential geometry. Vol. XI 11, Surv. Differ. Geom. Int. Press, Somerville, MA, 2007, pp. 103–136 DOI: 10.4310/SDG.2006.v11.n1.a5
  • [26] Martin Kell “Transport maps, non-branching sets of geodesics and measure rigidity” In Adv. Math. 320, 2017, pp. 520–573 DOI: 10.1016/j.aim.2017.09.003
  • [27] Benoît Kloeckner “A geometric study of Wasserstein spaces: Euclidean spaces” In Ann. Sc. Norm. Super. Pisa Cl. Sci. (5) 9.2, 2010, pp. 297–323
  • [28] John Lott and Cédric Villani “Ricci curvature for metric-measure spaces via optimal transport” In Ann. of Math. (2) 169.3, 2009, pp. 903–991 DOI: 10.4007/annals.2009.169.903
  • [29] Grigori Perelman “Alexandrov spaces with curvatures bounded from below II” In preprint, 1991
  • [30] Jaime Santos-Rodríguez “On isometries of compact LpL^{p}-Wasserstein spaces” In Adv. Math. 409, 2022, pp. Paper No. 108632\bibrangessep21 DOI: 10.1016/j.aim.2022.108632
  • [31] Cédric Villani “Optimal transport: old and new” Springer, 2009
  • [32] Alan Weinstein “On the homotopy type of positively-pinched manifolds” In Arch. Math. (Basel) 18, 1967, pp. 523–524 DOI: 10.1007/BF01899493
  • [33] Sergio Zamora “Anderson finiteness for RCD spaces” In MPIM Preprint Series (4), 2023 URL: https://archive.mpim-bonn.mpg.de/id/eprint/5009/1/mpim-preprint_2023-4.pdf
  • [34] Sergio Zamora “First Betti number and collapse”, 2022 arXiv:2209.12628 [math.DG]