Skip to main content
Springer logoLink to Springer
. 2022 Oct 30;387(3-4):1435–1480. doi: 10.1007/s00208-022-02493-7

Moduli spaces of compact RCD(0,N)-structures

Andrea Mondino 1,, Dimitri Navarro 1
PMCID: PMC10587325  PMID: 37869581

Abstract

The goal of the paper is to set the foundations and prove some topological results about moduli spaces of non-smooth metric measure structures with non-negative Ricci curvature in a synthetic sense (via optimal transport) on a compact topological space; more precisely, we study moduli spaces of RCD(0,N)-structures. First, we relate the convergence of RCD(0,N)-structures on a space to the associated lifts’ equivariant convergence on the universal cover. Then we construct the Albanese and soul maps, which reflect how structures on the universal cover split, and we prove their continuity. Finally, we construct examples of moduli spaces of RCD(0,N)-structures that have non-trivial rational homotopy groups.

Introduction

One of Riemannian geometry’s most fundamental problems is studying metrics satisfying a particular curvature constraint on a fixed smooth manifold. Three thoroughly studied types of curvature are the sectional, the Ricci, and the scalar curvature; common curvature constraints are lower (resp. upper) bounds on the corresponding curvature. Also, when a smooth manifold admits a metric of the desired type, it is interesting to describe such metrics’ space. A way to tackle this problem is to study the topological properties of the associated moduli space, i.e., the quotient of the space of all metrics satisfying the curvature condition by isometry equivalence. In the last decades, moduli spaces of metrics with positive scalar curvature (resp. negative sectional curvature) have been studied intensively (see [38] for a comprehensive introduction). Yet, there are not as many results on moduli spaces of non-negatively Ricci curved metrics. In 2017, Tuschmann and Wiemeler published the first result on these moduli spaces’ homotopy groups (see Theorem 1.1 in [37]).

Until recently, most of the results on moduli spaces in metric geometry focused on smooth metrics. Nevertheless, Belegradek [7] recently tackled the case of non-negatively curved Alexandrov spaces, studying the moduli space of non-negatively curved length metrics on the 2-sphere. Whereas Alexandrov introduced curvature lower bounds in the setting of length spaces (generalizing sectional curvature lower bounds), RCD spaces generalize lower bounds on the Ricci curvature to the setting of metric measure spaces. Indeed, roughly, RCD(0,N) spaces should be thought as possibly non smooth spaces with dimension bounded above by N[1,) and non-negative Ricci curvature, in a synthetic sense. The first goal of the present paper is to set the foundations for studying moduli spaces of RCD(0,N)-structures. The question we will be studying in this paper is the following:

Question

Let N[1,), and let X be a compact topological space that admits an RCD(0,N)-structure. What can be said about the topology of the moduli space of RCD(0,N)-structures on X?

We will start by recalling in Sect. 1.1 what an RCD(0,N) space is. Then, in Sect. 1.2, we will introduce RCD(0,N)-structures on a fixed topological space, together with the associated moduli space. In Sect. 1.3, we will introduce the notions of lift and push-forward of an RCD(0,N)-structure. Afterwards, in Sect. 1.4, we will present concisely the Albanese map and the soul map associated to a compact topological space that admits an RCD(0,N)-structure. Finally, in Sect. 1.5, we will present our main results.

RCD(0,N)-spaces

The story of RCD spaces has its roots in Gromov’s precompactness Theorem (see Corollary 11.1.13 of [30]). The result states that sequences of compact Riemannian manifolds with a lower bound on the Ricci curvature, and an upper bound on both the dimension and the diameter are precompact in the Gromov–Hausdorff topology (GH topology for short). Since then, there has been much work to understand the properties of limits of such sequences, called Ricci limit spaces. In the early ’00s, Cheeger and Colding published in [14, 15] and [16] an extensive study of the aforementioned spaces. One important observation (already noticed by Fukaya) is that, to retain good stability properties at the limit space, it is fundamental to keep track of the Riemannian measures’ behaviour associated to the approximating sequence. Since then, it is common to use the measured Gromov–Hausdorff topology (mGH topology for short), endowing Riemannian manifolds with their normalized volume measure.

A related (but slightly different) approach is to introduce a new definition of Ricci curvature lower bounds and dimension upper bound, at the more general level of possibly non-smooth metric measure spaces, which generalizes the classical notions and is stable when passing to the limit in the mGH topology. The definition of RCD spaces is an example of such a definition. Therefore, any result proven for RCD spaces (with the tools of metric measure theory) would hold a fortiori for Ricci limit spaces.

Throughout the paper, we will use the following definition of metric measure spaces.

Definition 1.1

Let (X,d,m) be a triple where (X,d) is a metric space and m is a measure on X. We say that (X,d,m) is a metric measure space (m.m.s. for short) when (X,d) is a complete separable metric space and m is a non-negative boundedly finite Radon measure on (X,d).

CD-spaces were introduced independently by Lott and Villani in [25], and by Sturm in [35] and [36]. For simplicity, we only give the definition of CD(0,N)-spaces (following Definition 1.3 in [36]). An extensive study of CD-spaces is given in [39, Chapters 29 and 30].

Denote by P2(X,d,m) the space of probability measures that are absolutely continuous w.r.t. m and with finite variance, and let W2 be the quadratic Kantorovitch-Wasserstein transportation distance. Let also SN(μm) be the Renyi entropy of μP2(X,d,m) with respect to m, i.e. SN(ρmm)=-NXρ1-1Nm.

Definition 1.2

Given N[1,), a CD(0,N)-space is a m.m.s. (X,d,m) such that m(X)>0 and satisfying the following property: for every pair μ0,μ1P2(X,d,m), there exists a W2-geodesic (μt)t[0,1]P2(X,d,m) from μ0 to μ1 such that the function [0,1]tSN(μtm) is convex.

The class of CD spaces includes (some) non-Riemannian Finsler structures which, from the work of Cheeger-Colding [1416], cannot appear as Ricci limits. In order to single out the “Riemannian” CD structures, it is convenient to add the assumption that the Sobolev space W1,2 is a Hilbert space [3, 22] (see also [2]): indeed for a Finsler manifold, W1,2 is a Banach space and W1,2 is a Hilbert space if and only if the Finsler structure is actually Riemannian. Such a condition is known as “infinitesimal Hilbertianity”.

Definition 1.3

Given N[1,), a m.m.s. (X,d,m) is an RCD(0,N) space if it is an infinitesimally Hilbertian CD(0,N)-space.

An equivalent way to characterise the RCD(0,N) condition is via the validity of the Bochner inequality [4, 5, 19] (see also [12] for the globalization of general Ricci lower bounds RCD(K,N)).

Moduli spaces of RCD(0,N)-structures on compact topological spaces

In this section, we will start by defining RCD(0,N)-structures on a fixed topological space; then, we will introduce the associated moduli space, together with its topology.

Definition 1.4

Given a topological space X and N[1,), an RCD(0,N)-structure on X is a metric measure structure (X,d,m) such that: d metrizes the topology of X, Spt(m)=X, and (X,d,m) is an RCD(0,N) space.

It is common to identify m.m.s. that are isomorphic. There are two distinct notions of isomorphisms for m.m.s. (see the discussion in Chapter 27, Section “adding the measure” in [39]). However, both notions coincide when restricted to the set of RCD(0,N)-structures on a topological space (since we imposed measures to have full support in that case). We will adopt the following definition of isomorphism between metric measure spaces.

Definition 1.5

Two m.m.s. (X1,d1,m1) and (X2,d2,m2) are isomorphic when there is a bijective isometry ϕ:(X1,d1)(X2,d2) such that ϕm1=m2.

Thanks to [1, Theorem 2.3] and [23, Remark 3.29], the following result holds.

Theorem 1.1

The compact Gromov–Hausdorff–Prokhorov distance dGHPc (see [1, Section 2.2]) is a complete separable metric on the set X of isomorphism classes of compact metric measure spaces. Moreover, dGHPc metrizes the mGH topology (see for instance [39, Definition 27.30]).

We are now in position to introduce the moduli space M0,N(X) of RCD(0,N)-structures on a compact topological space X; this will be the main object of study in the paper.

Notation 1.1

Let N[1,), and let X be a compact topological space that admits an RCD(0,N)-structure. We introduce the following spaces:

  • (i)

    RCD(0,N)X is the set of isomorphism classes of compact RCD(0,N) spaces with full support, endowed with the mGH-topology (seen as a subspace of X),

  • (ii)

    R0,N(X) is the set of all RCD(0,N)-structures on X,

  • (iii)

    M0,N(X) is the quotient of R0,N(X) by isomorphisms, endowed with the mGH-topology (seen as a subspace of RCD(0,N)).

We call M0,N(X) the moduli space of RCD(0,N)-structures on X.

Lift and push-forward

To our aims, a fundamental result is the existence of a universal cover for an RCD space (see [27, Theorem 1.1]).

Theorem 1.2

Let N[1,), and let X be a compact topological space that admits an RCD(0,N)-structure. Then X admits a universal cover p:X~X. We denote π¯1(X) the associated group of deck transformations, also called the revised fundamental group of X.

Remark 1.1

In Theorem 1.2, the universal cover must be understood in a general sense (as defined in [34, Chap. 2, Sect. 5]). To be precise, it is not known whether it is simply connected or not. In particular, the revised fundamental group π¯1(X) is a quotient of the fundamental group π1(X) of X, and may a priori not be isomorphic to π1(X).

As in the case of a Riemannian manifold, it is possible to lift an RCD(0,N)-structure on a compact topological space to its universal cover, and, conversely, to push-forward an equivariant RCD(0,N)-structure on the universal cover back to the base space. Indeed, let us fix a real number N[1,), a compact topological space X that admits an RCD(0,N)-structure, and denote p:X~X the universal cover of X. We will see in Sect. 2.1 that:

  • given an RCD(0,N)-structure (X,d,m) on X, there exists a unique RCD(0,N)-structure (X~,d~,m~) on X~ (called the lift of (X,d,m)) such that p:(X~,d~,m~)(X,d,m) is a local isomorphism, and π¯1(X) acts by isomorphism on (X~,d~,m~) (see Corollary 2.1);

  • given an RCD(0,N)-structure (X~,d~,m~) on X~ such that π¯1(X) acts by isomorphism on (X~,d~,m~), there exists a unique RCD(0,N)-structure (X,d,m) on X (called the push-forward of (X~,d~,m~)) such that p:(X~,d~,m~)(X,d,m) is a local isomorphism (see Proposition 2.3).

Theorem A, which will be introduced in Sect. 1.5, relates the convergence of RCD(0,N)-structures on X to the convergence of the associated lifts.

Soul map and Albanese map

In this section, we fix a real number N[1,), a compact topological space X that admits an RCD(0,N)-structure, and denote p:X~X the universal cover of X.

A special property enjoyed by RCD(0,N) spaces is the existence of splittings. More precisely, given an RCD(0,N)-structure (X,d,m) on X, and denoting (X~,d~,m~) the associated lift (see Sect. 1.3), there exists (thanks to Theorem 1.3 in [27]) an isomorphism:

ϕ:(X~,d~,m~)(X¯,d¯,m¯)×Rk,

where kN[0,N] is called the degree of ϕ, Rk is endowed with the Euclidean distance and the Lebesgue measure, and (X¯,d¯,m¯) is a compact RCD(0,N-k)-space such that π¯1(X¯)=0; the space X¯ is called the soul of ϕ (see Theorem 1.2 for the definition of π¯1). Such a map ϕ is called a splitting of (X~,d~,m~), and induces an isomorphism:

ϕ:Iso(X~,d~,m~)Iso(X¯,d¯,m¯)×Iso(Rk).

Moreover, the revised fundamental group π¯1(X) acts by isomorphism onto (X~,d~,m~). Therefore, applying ϕ and projecting onto Iso(Rk), we get a group homomorphism ρRϕ:π¯1(X)Iso(Rk). In Sect. 2.2, we will prove the following properties:

  • the degree k does not depend either on the chosen splitting ϕ, or the chosen RCD(0,N)-structure (X,d,m) on X (see Corollary 2.2);

  • the image Γϕ:=Im(ρRϕ) is a crystallographic subgroup of Iso(Rk); moreover, up to conjugating with an affine transformation, Γϕ does not depend either on the chosen splitting ϕ, or the chosen RCD(0,N)-structure (X,d,m) on X (see Proposition 2.7).

Therefore, it is possible to introduce k(X):=k (called the splitting degree of X), and the set Γ(X) of crystallographic subgroups of Iso(Rk) that are conjugated to Γϕ by an affine transformation (called the crystallographic class of X), both being topological invariants of X.

Thanks to the above discussion, to any RCD(0,N)-structure (X,d,m) on X with lift (X~,d~,m~), and to any splitting ϕ of (X~,d~,m~), we can associate:

  • the soul (X¯,d¯,m¯) of ϕ which is a compact RCD(0,N-k(X))-space;

  • the compact k(X)-dimensional flat orbifold (Rk(X)/Γϕ,dΓϕ) with orbifold fundamental group equal to ΓϕΓ(X) (see the discussion preceding Definition 2.3).

We will denote A(X) the set of compact flat k(X)-dimensional orbifolds whose orbifold fundamental group belong to Γ(X) (called the Albanese class of X).

Remark 1.2

Let (MN,g) be a compact N-dimensional Riemannian manifold with non-negative Ricci curvature, and such that π1(M)=Zk. Observe that, in that case, (M,dg,mg) is an RCD(0,N) space, where dg is the geodesic distance and mg is the Riemannian measure. It is possible to show that the orbifold obtained following the discussion above is nothing but the usual Albanese variety of (Mg) (up to isometry).

In Sect. 2.4.2, we will see that, up to isomorphism, the orbifold (resp. the soul) does not depend either on the choice of the splitting map ϕ, or on the isomorphism class of (X,d,m) (see Lemma 2.1). Therefore, we will be able to introduce:

  • the Albanese map A:M0,N(X)Mflat(A(X)) associated to X, where Mflat(A(X)) is the quotient of A(X) by isometry equivalence (endowed with the GH topology);

  • the soul map S:M0,N(X)RCD(0,N-k(X)) associated to X;

such that for any RCD(0,N)-structure (X,d,m) on X with lift (X~,d~,m~) and any splitting ϕ of (X~,d~,m~) with soul (X¯,d¯,m¯), we have A([X,d,m])=[Rk(X)/Γϕ,dΓϕ] and S([X,d,m])=[X¯,d¯,m¯] (where brackets denote the equivalence class in the appropriate moduli space).

Theorem B, which will be presented in Sect. 1.5, states the continuity of the Albanese and soul maps.

Main results

Our first result relates the convergence of RCD(0,N)-structures on a compact topological space to the convergence of the associated lifts.

Theorem A

Let N[1,), let X be a compact topological space that admits an RCD(0,N)-structure, and denote p:X~X the universal cover of X. Assume that for every nN{}:

  • Xn=(X,dn,mn,n) is a pointed RCD(0,N)-structure on X;

  • X~n=(X~,d~n,m~n,~n) is the associated pointed lift, where ~n is any point in p-1(n).

Then {Xn}nN converges to X in the pmGH topology if and only if {X~n}nN converges to X~ in the equivariant pmGH topology (introduced in Sect. 2.3.2).

Remark 1.3

Note that, since X is compact, it is also possible to formulate Theorem A as follows (forgetting about the reference points in the base space):

Assume that for every nN{}, Xn=(X,dn,mn) is an RCD(0,N)-structure on X with lift X~n=(X~,d~n,m~n). Then {Xn}nN converges to X in the mGH topology if and only if for every nN{}, there exist ~nX~ such that {(X~n,~n)}nN converges to (X~,~) in the equivariant pmGH topology.

As we will observe at the end of Sect. 2.4.1, Theorem A implies the following corollary, which is particularly useful when computing the homeomorphism type of specific examples of moduli spaces (see for example the case of RP2 in [28]).

Corollary A

Let N[1,), let X be a compact topological space that admits an RCD(0,N)-structure, and denote p:X~X the universal cover of X.

Then the lift map:

p:M0,Np(X)M0,Np,eq(X~),

is a homeomorphism (introduced in Sect. 2.4.1), where M0,Np,eq(X~) and M0,Np(X) are respectively the moduli space of equivariant pointed RCD(0,N)-structures on X~ and the moduli space of pointed RCD(0,N)-structures on X (introduced in Sect. 2.3).

Observe that it is more straightforward to obtain Corollary A by using Theorem A than its equivalent version given in Remark 1.3.

Our next result states the continuity of the Albanese map and the soul map associated to a compact topological space that admits an RCD(0,N)-structure (with N[1,)). On the first hand, this result is essential when computing the homeomorphism type of specific examples of moduli spaces (see for example the case of the Möbius band M2 and the finite cylinder S1×[0,1] in [28]). On the other hand, the continuity of the Albanese map will be crucial in the proof of Theorem C.

Theorem B

Let N[1,), and let X be a compact topological space that admits an RCD(0,N)-structure. Then, the Albanese map A:M0,N(X)Mflat(A(X)) and the soul map S:M0,N(X)RCD(0,N-k(X)) are continuous, where Mflat(A(X)) and RCD(0,N-k(X)) are respectively endowed with the GH and mGH topology.

Let us recall that if X is a compact topological space that admits an RCD(0,2)-structure, then the moduli space M0,2(X) is contractible (see Theorem 1.1 in [28]). Theorem C should be put in contrast with that result since it shows that the topology of moduli spaces of RCD(0,N)-structures is not always as trivial. Moreover, Theorem C can also be seen as a non-smooth analogue of Theorem 1.1 in [37].

Theorem C

Let N[1,) and let X be a compact topological space that admits an RCD(0,N)-structure such that π¯1(X)=0 (see Theorem 1.2 for the definition of π¯1(X)). In addition, let Y be either S1×K2 (where K2 is the Klein bottle) or a torus of dimension k4 such that k8,9,10. Then, the moduli space M0,N+dim(Y)(X×Y) has non-trivial higher rational homotopy groups.

Thanks to Theorem C, we immediately obtain the following corollary, which can be seen as a non-smooth analogue of Corollary 1.2 in [37].

Corollary B

For every N3 (resp. N4 / N5) there exists a compact topological space X such that M0,N(X) is not simply connected (resp. has non-trivial third rational homotopy group / non-trivial fifth rational homotopy group).

In Sect. 2, we will introduce in full details the main objects and constructions of the paper. In Sect. 3, we will prove the main results.

Preliminaries

Throughout this section:

  • N[1,) is a real number,

  • X is a compact topological space that admits an RCD(0,N)-structure,

  • p:X~X denotes the universal cover of X (whose existence is given by Theorem 1.2).

In Sect. 2.1, we will introduce the notions of lift (resp. push-forward) of an RCD(0,N)-structure on X (resp. on X~).

In Sect. 2.2, we will present splittings and use them to construct topological invariants associated to X (splitting degree, crystallographic class and Albanese class).

In Sect. 2.3, we will define the moduli space of pointed RCD(0,N)-structures on X; then, we will introduce the moduli space of equivariant pointed RCD(0,N)-structures on the universal cover X~.

In Sect. 2.4, we will define the lift and push-forward map (which are important to get Corollary A), and the Albanese and soul maps.

Covering space theory of RCD(0,N)-spaces

In this section, we will start by introducing δ-covers associated to an RCD(0,N)-structure on X. Then, we will explain how to lift an RCD(0,N)-structure on X to the associated δ-cover. Afterwards, we will explain how the universal cover of X is related to δ-covers. Subsequently, we will explain how to lift an RCD(0,N)-structure on X to its universal cover X~, and, conversely, how to push-forward an equivariant RCD(0,N)-structure on X~ onto X. Finally, we will introduce the Dirichlet domain associated to an RCD(0,N)-structure on X.

Before introducing δ-covers, we recall the following result (Chapter 2, Sections 4 and 5 of [34]) that associates a regular covering pU:XUX to any open cover U of X.

Proposition 2.1

Given an open cover U of X, there exists a unique regular covering pU:XUX (up to equivalence) such that:

yXU,pUπ1(XU,y)=π1(U,pU(y)),

where π1(U,pU(y)) is composed of homotopy classes of loops of the form ω-1αω, where α is a loop contained in some UU and ω is a path from pU(y) to α(0). Moreover, every connected open set UU is evenly covered bu pU.

The notion of δ-cover was introduced first by Sormani and Wei to prove the existence of a universal cover for Ricci limit spaces (see Theorem 1.1 in [33]). Later, it has also been used by Mondino and Wei in [27] to prove Theorem 1.2. These covering spaces will be very important in the proof of Theorem A.

Definition 2.1

Given δ>0 and (X,d,m) an RCD(0,N)-structure on X, the δ-cover associated to (X,d,m) is the regular covering pdδ:XdδX associated to the open cover U(δ,d) consisting of balls of radius δ for the distance d (see Proposition 2.1). We write G(δ,d) the associated group of deck transformations.

In the following result, we introduce the lift of an RCD(0,N)-structure on X to a δ-cover.

Proposition 2.2

Given δ>0 and (X,d,m) an RCD(0,N)-structure on X, there exists a unique RCD(0,N)-structure (Xdδ,dδ,mδ) on Xdδ such that pdδ:(Xdδ,dδ,mδ)(X,d,m) is a local isomorphism. Moreover, we have the following properties:

  • (i)

    for every x~,y~Xdδ, we have dδ(x~,y~)=inf{Ld(pdδγ~)}, where the infimum is taken over all continuous path γ~:[0,1]Xdδ from x~ to y~ and Ld is the length structure induced by d,

  • (ii)

    for every Borel subset E~Xdδ such that pdδ|E~ is an isometry, we have mδ(E~)=m(E),

  • (iii)

    the group of deck transformations G(δ,d) is a subgroup of   Isom.m.s.(Xdδ,dδ,mδ),

  • (iv)

    for every x~Xdδ and every rδ, the restriction of pdδ to Bdδ(x~,r) is a homeomorphism onto Bd(pdδ(x~),r),

  • (v)

    for every x~Xdδ and every rδ/2, the restriction of pdδ to (Bdδ(x~,r),dδ,mδ) is an isomorphism onto (Bd(x,r),d,m).

Proof

First of all, there is obviously at most one RCD(0,N)-structure on Xdδ such that pdδ is a local isomorphism.

Then, thanks to Lemma 2.18 of [27], (Xdδ,dδ,mδ) is an RCD(0,N) space (where dδ and mδ are defined as in (i) and (ii)). Moreover, it is readily checked that dδ metrizes the topology of Xdδ, that Spt(mδ)=Xdδ, and that G(δ,d) acts by isomorphism. Therefore, (Xdδ,dδ,mδ) is an RCD(0,N)-structure on Xdδ satisfying point (i) to (iii).

Finally, thanks to Proposition 15 of [31], and by definition of mδ, point (iv) and (v) are satisfied.

Now, we put the universal cover of X in relation with δ-covers (see Theorem 2.7 in [27] for a proof).

Theorem 2.1

Let (X,d,m) be an RCD(0,N)-structure on X, and let δ(X,d) be the supremum of all δ>0 such that every ball of radius δ in (X,d) is evenly covered by p. Then δ(X,d)>0, and for every δ<δ(X,d), p and pdδ are equivalent, and every equivalence map is an isomorphism between (X~,d~,m~) and (Xdδ,dδ,mδ).

Thanks to Proposition 2.2 and Theorem 2.1, we can introduce the lift of an RCD(0,N)-structure on X to the universal cover X~.

Corollary 2.1

Let (X,d,m) be an RCD(0,N)-structure on X. There is a unique RCD(0,N)-structure (X~,d~,m~) on X~ (called the lift of (X,d,m)) such that p:(X~,d~,m~)(X,d,m) is a local isomorphism. Moreover, the revised fundamental group π¯1(X) acts by isomorphism on (X~,d~,m~).

The following proposition is a sort of converse to Corollary 2.1; it introduces the push-forward of an equivariant RCD(0,N)-structure on X~ (cf. [27, Lemma 2.18] and [26, Lemma 2.24]).

Proposition 2.3

Let (X~,d~,m~) be an RCD(0,N)-structure on X~ such that π¯1(X) acts by isomorphisms on (X~,d~,m~). There is a unique RCD(0,N)-structure (X,d,m) on X (called the push-forward of (X~,d~,m~)) such that p:(X~,d~,m~)(X,d,m) is a local isomorphism. It satisfies the following properties:

  • (i)

    for every x,yX, we have d(x,y)=inf{d~(x~,y~)}, where the infimum is taken over all x~p-1(x) and y~p-1(y),

  • (ii)

    for every open set UX that is evenly covered by p, we have m(U)=m~(U~), where U~ is any open set in X~ such that p:U~U is a homeomorphism.

Proof

First of all, there is obviously at most one RCD(0,N)-structure on X such that p is a local isomorphism.

Then, let us define d and m as in points (i) and (ii). Observe that since X~ is locally compact, and since π¯1(X) acts by isometries, m is well defined, and the infimum in the definition of d is achieved. It is then readily checked that d is a distance on X, and that m defines a measure on X (using the fact that the Borel σ-algebra of X is generated by evenly covered open sets).

Let us now show that p is a local isomorphism. Let x~X~ and define x:=p(x~). There exists an open neighborhood U~ of x~ such that p:U~U:=p(U~) is a homeomorphism. Moreover, there exists r>0 such that Bd~(x~,r)U~. Let us show that, for every 0<rr, p is a homeomorphism from Bd~(x~,r) onto Bd(x,r). First, notice that p is distance decreasing; in particular, we have p(Bd~(x~,r))Bd(x,r). Now, let yBd(x,r). Since the infimum in the definition of d is achieved, there exists y~p-1(y) such that d~(x~,y~)=d(x,y)<r. Hence, p(Bd~(x~,r))=Bd(x,r). Since Bd~(x~,r) is a subset of U~, p is injective on Bd~(x~,r). Hence, p is a bijective map from Bd~(x~,r) onto Bd(x,r). However, p is an open map, hence it is a homeomorphism from Bd~(x~,r) onto Bd(x,r).

Now, let y~,z~Bd~(x~,r/3). Looking for a contradiction, let us suppose that d(y,z)<d~(y~,z~), where y:=p(y~) and z:=p(z~). In that case, there exists z~p-1(z) such that d~(y~,z~)=d(y,z)<d~(y~,z~)d~(x~,y~)+d~(x~,z~)<2r/3<r. However, p is a homeomorphism from Bd~(x~,r) onto Bd(x,r), so we should have z~=z~, which is the contradiction we were looking for. Hence, p is an isometry from Bd~(x~,r/3) onto Bd(x,r/3). Moreover, by definition of m, this implies that p is an isomorphism of metric measure space from Bd~(x~,r/3) onto Bd(x,r/3).

Then, it is easy to check that d metrizes the topology of X and that Spt(m)=X. To conclude, we just need to show that (X,d,m) is an RCD(0,N) space. To this aim, first of all observe that RCD(0,N) is equivalent to RCD(0,N) (by the explicit form of the distortion coefficients), which in turn is equivalent to CDe(0,N) plus infinitesimally Hilbertianity.

Observe that, since p is a local isometry, it preserves the length of curves; therefore, (X,d) is a compact geodesic space. Moreover, since X is compact and m~ is boundedly finite, m is necessarily a finite measure on X. Summarising: (X,d,m) is a compact geodesic space endowed with a finite measure, and it is locally isomorphic to an RCD(0,N) space in the sense that for every point xX there exists a closed metric ball B¯d(x,r). centred at x isomorphic to a closed metric ball B¯d~(x~,r). inside the RCD(0,N) space X~.

Notice that, by triangle inequality, if y,zB¯d(x,r/4). then any geodesic joining them is contained in B¯d(x,r). Recall also that, given two absolutely continuous probability measures with compact support in an RCD space, there exists a unique W2-geodesic joining them [24, Theorem 1.1]. It follows that, given two absolutely continuous probability measures with compact support contained in B¯d(x,r/4) (which in turm is isomorphic to B¯d(x,r/4)X~, and X~ satisfies RCD(0,N)), there exists a unique W2-geodesic joining them, its support is contained in B¯d(x,r), and it satisfies the convexity property of the CDe(0,N) condition.

In particular, (X,d,m) satisfies the strong CDloce(0,N) condition in the sense of [19] and it is locally infinitesimally Hilbertian. Then, using [19, Theorem 3.25], we obtain that (X,d,m) satisfies RCD(0,N).

Remark 2.1

Observe that if (X,d,m) is an RCD(0,N)-structure on X, then the push-forward of the lift of (X,d,m) is equal to (X,d,m), thanks to Proposition 2.3 and Corollary 2.1. The same is true in the other direction; if (X~,d~,m~) is an RCD(0,N)-structure on X~ such that π¯1(X) acts by isomorphisms, then the lift of the push-forward of (X~,d~,m~) is equal to (X~,d~,m~).

We conclude this section with the following results that introduces the Dirichlet domain associated to an RCD(0,N)-structure on X.

Proposition 2.4

Let (X,d,m) be an RCD(0,N)-structure on X and let x~X~. We define the Dirichlet domain with center x~ associated to (X,d,m) by:

F(x~):=ηπ¯1(X)ϕη-1(R0)

where ϕη(y~):=d~(y~,ηx~)-d~(y~,x~), for y~X~. The Dirichlet domain satisfies the following two properties:

  • (i)

    for every y~X~, there exists ηπ¯1(X) such that ηy~F(x~),

  • (ii)

    for every y~F(x~), we have d~(x~,y~)=d(x,y), where x:=p(x~) and y:=p(y~).

In particular, F(x~)Bd~(x~,D), where D:=Diam(X,d).

Proof

We start with the proof of (i). Let y~X~ and define R:=d~(x~,y~). Then, p-1(x)B¯d~(y~,R) is a compact, discrete, non empty set; hence, it contains finitely many points. In particular, there exists ηπ¯1(X) such that ηx~B¯d~(y~,R), and such that:

z~p-1(x)B¯d~(y~,R),d~(y~,ηx~)d~(y~,z~). 1

Now, assume that μπ¯1(X). If Rd~(y~,μx~), we have d~(y~,ηx~)d~(y~,μx~) since d~(y~,ηx~)R. If d~(y~,μx~)<R, then, thanks to equation 1, we also have d~(y~,ηx~)d~(y~,μx~). Thus, for every μπ¯1(X), we get d~(y~,ηx~)d~(y~,μx~). Hence, for every μπ¯1(X), we have ϕμ(η-1y~)=d~(η-1y~,μx~)-d~(η-1y~,x~)=d~(y~,ημx~)-d~(y~,ηx~)0. In conclusion, η-1y~F(x~).

Now we prove (ii). Assume that y~F(x~). We define y:=p(y~) and we assume that β:[0,1]X is a minimizing geodesic from x to y. Let β~ be the lift of β starting at x~ and let ηπ¯1(X) such that β~(1)=ηy~. Looking for a contradiction, let us suppose that d(x,y)<d~(x~,y~). Then, observe that d~(x~,ηy~)L(β)=d(x,y); in particular, d~(x~,ηy~)=d(x,y), since p contracts distances. Hence, we have ϕη-1(y~)=d~(y~,η-1x~)-d~(x~,y~)=d~(ηy~,x~)-d~(x~,y~)=d(x,y)-d~(x~,y~)<0. In particular, y~F(x~), which is the contradiction we were looking for. Thus, d(x,y)d~(x~,y~), and, since p contracts distances, we have d(x,y)=d~(x~,y~). This concludes the proof.

Splittings and topological invariants

In this section, we will introduce the notion of splitting associated to an RCD(0,N)-structure on X. To any splitting ϕ, we will associate a degree k and a Euclidean homomorphism ρRϕ:π¯1(X)Iso(Rk), and we will investigate the properties of Γ(ϕ)=Im(ρRϕ). We will prove that the degree k and the affine conjugacy class of Γ(ϕ) do not depend either on the chosen splitting ϕ, or the chosen RCD(0,N)-structure on X. This will lead us to introduce the splitting degree k(X) and the crystallographic class Γ(X) of X, which are topological invariants of X. Finally, we will introduce the Albanese class A(X) of X, which consists of orbifolds whose fundamental group belong to Γ(X).

First of all, let us introduce the definition of splittings.

Definition 2.2

Let (X,d,m) be an RCD(0,N)-structure on X, and denote (X~,d~,m~) its lift. A splitting of (X~,d~,m~) is an isomorphism ϕ:(X~,d~,m~)(X¯,d¯,m¯)×Rk, where Rk is endowed with the Euclidean distance and Lebesgue measure, kN[0,N] is called the degree of ϕ, and (X¯,d¯,m¯) is a compact RCD(0,N-k)-space with trivial revised fundamental group called the soul of ϕ.

Thanks to Theorem 1.3 in [27], which in turn built on top of the Splitting Theorem for RCD(0,N) spaces [21], we have the following existence result.

Theorem 2.2

For every RCD(0,N)-structure (X,d,m) on X, the lift (X~,d~,m~) admits a splitting. Moreover, for every splitting ϕ of (X~,d~,m~), the group of isomorphisms of (X¯,d¯,m¯)×Rk splits, i.e., we have:

Isom.m.s.((X¯,d¯,m¯)×Rk)=Isom.m.s.(X¯,d¯,m¯)×Iso(Rk),

where (X¯,d¯,m¯) is the soul of ϕ and k is the degree of ϕ.

Thanks to Theorem 2.2, Theorem 2.1, and Proposition 2.2, we can introduce the following notations.

Notation 2.1

Let (X,d,m) be an RCD(0,N)-structure on X and let ϕ be a splitting of its lift (X~,d~,m~) with degree k and soul (X¯,d¯,m¯). We write:

  • (i)

    pSϕ (resp. pRϕ) the projection of Isom.m.s.((X¯,d¯,m¯)×Rk) onto Isom.m.s.(X¯,d¯,m¯) (resp. Iso(Rk)),

  • (ii)

    ι the inclusion of π¯1(X) into Isom.m.s.(X~,d~,m~),

  • (iii)

    ϕ the isomorphism from Isom.m.s.(X~,d~,m~) onto Isom.m.s.((X¯,d¯,m¯)×Rk) defined by ϕ(η):=ϕηϕ-1 for every ηIsom.m.s.(X~,d~,m~).

We call ρSϕ:=pSϕϕι (resp. ρRϕ:=pRϕϕι) the soul homomorphism associated to ϕ (resp. the Euclidean homomorphism associated to ϕ) and we write K(ϕ):=Ker(ρRϕ) and Γ(ϕ):=Im(ρRϕ).

The next result shows that the kernel and the image of the Euclidean homomorphism associated to a splitting enjoy particular group structures.

Proposition 2.5

Let (X,d,m) be an RCD(0,N)-structure on X and let ϕ be a splitting of (X~,d~,m~) with degree k and soul (X¯,d¯,m¯). Then, K(ϕ) is a finite normal subgroup of π¯1(X) and Γ(ϕ) is a crystallographic subgroup of Iso(Rk) (i.e. it acts cocompactly and discretly on Rk).

Proof

First, let us show that K(ϕ) is finite. Observe that every element ηK(ϕ) satisfies η(X¯×{0})(X¯×{0}). However, X¯×{0} is a compact subset of X¯×Rk and π¯1(X) acts properly on X¯×Rk; thus, K(ϕ) is finite.

Now, let us show that Γ(ϕ) acts cocompactly on Rk. Thanks to the first isomorphism theorem for topological spaces, there is a continuous map μ such that the following diagram is commutative: graphic file with name 208_2022_2493_Figa_HTML.jpg where qi (i{1,2}) are the quotient maps. Moreover, μ is surjective since q2pRk is surjective. Finally, X is homeomorphic to (X¯×Rk)/π¯1(X); in particular, (X¯×Rk)/π¯1(X) is compact and Rk/Γ(ϕ) is compact, being the image of a compact topological space by a continuous surjective map. In conclusion, Γ(ϕ) acts cocompactly on Rk.

Let us prove that Γ(ϕ) acts discretely on Rk (i.e. its orbits are discrete subsets of Rk). First, observe that it is sufficient to prove that Γ(ϕ) acts properly on Rk. To prove this, let K be a compact subset of Rk and let us show that there are only finitely many elements gΓ(ϕ) such that gKK. By definition of Γ(ϕ), we have:

{gΓ(ϕ),gKK}=ρRϕ({ηπ¯1(X),η(X¯×K)(X¯×K)}).

However, π¯1(X) acts properly on X¯×Rk and X¯×K is compact; hence:

{ηπ¯1(X),η(X¯×K)(X¯×K)}

is finite. Thus, {gΓ(ϕ),gKK} is finite, being the image of a finite set.

The following corollary of Proposition 2.5 defines the splitting degree of X (cf. [26, Proposition 2.25]).

Corollary 2.2

(Splitting degree k(X)) The revised fundamental group π¯1(X) is a finitely generated group which has polynomial growth of order k(X)N[0,N]. Moreover, given any RCD(0,N)-structure (X,d,m) on X with lift (X~,d~,m~), the degree of any splitting ϕ of (X~,d~,m~) is equal to k(X). We call k(X) the splitting degree of X.

Proof

Thanks to Proposition 2.5, Γ(ϕ) is a crystallographic subgroup of Iso(Rk), where k[0,N]N is the degree of ϕ. We need to prove that π¯1(X) has polynomial growth of order k.

By Bieberbach’s first Theorem (see Theorem 3.1 in [13]), Γ(ϕ) admits a normal subgroup Γ(ϕ)Rk such that Γ(ϕ)Rk is isomorphic to Zk and Γ(ϕ)Rk has finite index in Γ(ϕ). In particular, Γ(ϕ)Rk is finitely generated, has polynomial growth of order k, and is a normal subgroup of Γ(ϕ) with finite index; thus, Γ(ϕ) is also finitely generated and has polynomial growth of order k. Now, π¯1(X)/K(ϕ) is isomorphic to Γ(ϕ); hence it is finitely generated with polynomial growth of order k. However, K(ϕ) is finite and is a normal subgroup of π¯1(X); thus, π¯1(X) is also finitely generated and has polynomial growth of order k.

The revised fundamental group satisfies the following additional group property (which will be crucial in the proof of Theorem A).

Proposition 2.6

The revised fundamental group π¯1(X) is a Hopfian group, i.e., every surjective group homomorphism from π¯1(X) onto itself is an isomorphism.

Proof

First of all, let us recall some results from group theory:

  • (i)

    Noetherian groups (every subgroup is finitely generated) are Hopfian groups.

  • (ii)

    If H is a normal subgroup of G such that both H and G/H are Noetherian, then G is Noetherian.

  • (iii)

    Finite groups are Noetherian.

  • (iv)

    Finitely generated abelian groups are Noetherian.

Let us fix an RCD(0,N)-structure (X,d,m) on X and let ϕ be a splitting of its lift (X~,d~,m~). By Proposition 2.5 and Corollary 2.2, Γ(ϕ) is a crystallographic subgroup of Iso(Rk(X)). Hence, by Bieberbach’s 1st Theorem (Theorem 3.1 in [13]), Γ(ϕ)Rk(X) is isomorphic to Zk(X). In particular, Γ(ϕ)Rk(X) is Noetherian thanks to (iv). Moreover, Γ(ϕ)Rk is normal in Γ(ϕ) and the quotient Γ(ϕ)/Γ(ϕ)Rk is finite. Thanks to (iii), Γ(ϕ)/Γ(ϕ)Rk is Noetherian, and, using (ii), Γ(ϕ) is Noetherian. In addition, K(ϕ) is finite by Proposition 2.5, so it is Noetherian by (iii). Finally, π¯1(X)/K(ϕ) is isomorphic to Γ(ϕ), so it is Noetherian. In conclusion, thanks to (ii), π¯1(X) is Noetherian; hence, it is Hopfian using (i).

Given kN, two crystallographic subgroups of Iso(Rk) are called equivalent if they are conjugated by an affine transformation. The set Crys(k) of equivalence classes of crystallographic subgroups of Iso(Rk) is a finite set thanks to Bieberbach’s third Theorem (see Theorem 7.1 in [13]). The following result defines the crystallographic class of X.

Proposition 2.7

(Crystallographic class Γ(X)) For i{1,2}, let (X,di,mi) be an RCD(0,N)-structure on X, and let ϕi be a splitting of its lift (X~,d~i,m~i). Then Γ(ϕ1) and Γ(ϕ2) are equivalent as crystallographic subgroups of Iso(Rk(X)). We denote by Γ(X) the common equivalence class and call it the crystallographic class of X.

Proof

By Bieberbach’s second Theorem (see Theorem 4.1 of [13]), two crystallographic subgroups of Iso(Rk) are conjugated by an affine transformation if and only if they are isomorphic (we let k:=k(X)). We need to show that Γ(ϕ1) and Γ(ϕ2) are isomorphic. Observe that, for i{1,2}, we have the following exact sequence of groups: graphic file with name 208_2022_2493_Figb_HTML.jpg where ι is just the inclusion, Γ(ϕi) is a crystallographic subgroup of Iso(Rk), and K(ϕi) is finite. By Remark 2.5 of [40], K(ϕi)=ι(K(ϕi)) is uniquely characterized as the maximal finite normal subgroup of π¯1(X). In particular, we necessarily have K(ϕ1)=K(ϕ2). In conclusion, Γ(ϕ1)π¯1(X)/K(ϕ1)=π¯1(X)/K(ϕ2)Γ(ϕ2); thus, Γ(ϕ1) is isomorphic to Γ(ϕ2).

Given kN, and Γ a crystallographic subgroup of Iso(Rk), the quotient space Rk/Γ has the structure of a compact flat orbifold of dimension k, whose orbifold metric dΓ satisfies:

dΓ([x],[y])=inf{|x-y|}, 2

where x,yRk, [x] and [y] are their equivalence classes in Rk/Γ, and the infimum is taken over all x[x] and y[y]. Moreover, equivalent crystallographic groups give rise to orbifolds that are affinely equivalent.

Conversely, given a compact flat orbifold (X,d) of dimension k, the orbifold fundamental group π1orb(X) acts by isometries on the orbifold universal cover, which is Rk (the action being discrete and cocompact). Hence, one can associate a crystallographic group to (X,d). Finally, two affinely equivalent flat orbifolds of dimension k have isomorphic orbifold fundamental groups. Hence, by Bieberbach’s second Theorem (see Theorem 4.1 in [13]), they give rise to equivalent crystallographic groups (see the introduction of Section 2.1 in [8] for more details and some references).

Therefore, there is a one-to-one correspondence between equivalence classes of crystallographic subgroups of Iso(Rk) and affine equivalence classes of compact flat orbifolds of dimension k. This leads us to the definition of the Albanese class of X.

Definition 2.3

(Albanese class A(X)) We write A(X) the set of the affine equivalence classes of compact flat orbifolds determined by Γ(X), and call it the Albanese class of X. More explicitly, A(X) is the set of all flat orbifolds (Rk(X)/Γ,dΓ), where ΓΓ(X), and dΓ is defined in equation (2).

Moduli spaces and their topology

In Sect. 2.3.1, we will introduce the moduli space of pointed RCD(0,N)-structures on X. Then, in Sect. 2.3.2, we will introduce the moduli space of equivariant pointed RCD(0,N)-structures on X~. In particular, (based on the equivariant distance introduced by Fukaya and Yamaguchi in [20]), we will introduce the equivariant pmGH topology.

Moduli space of pointed RCD(0,N)-structures

Throughout the paper, we will use the following definition of pointed metric measure spaces.

Definition 2.4

A pointed metric measure space (p.m.m.s. for short) is a 4-tuple (X,d,m,), where (X,d,m) is a m.m.s. and X.

As for m.m.s., there are two distinct notions of isomorphisms between two pointed metric measure spaces. In this paper, we decided to use the following definition (which emphasizes the whole space’s metric structure, not only the metric structure of the measure’s support).

Definition 2.5

Two p.m.m.s. (X1,d1,m1,1) and (X2,d2,m2,2) are isomorphic when there is an isomorphism of metric measure spaces ϕ:(X1,d1,m1)(X2,d2,m2) such that ϕ(1)=2.

Thanks to Theorem 2.7 of [1], and Remark 3.29 of [23], we have the following result.

Theorem 2.3

The Gromov–Hausdorff–Prokhorov distance dGHP (see Section 2.3 of [1]) is a complete separable metric on the set Xp of isomorphism classes of pointed metric measure spaces that are locally compact and geodesic. Moreover, dGHP metrizes the pointed measured Gromov–Hausdorff topology (introduced in Definition 27.30 of [39]).

As we will see in Remark 2.2 below, it is possible to realize the pmGH convergence using maps with small distortion.

Notation 2.2

Let f:(X,dX)(Y,dY) be a map between metric spaces, we denote:

Dis(f):=sup{|dY(f(x),f(y))-dX(x,y)|},

where the supremum is taken over all couples x,yX.

Remark 2.2

Assume that {Xn=(Xn,dn,mn,n)}nN{} is a familly of locally compact geodesic p.m.m.s. such that XnX in the pmGH topology. Then, Theorem 2.3 implies that there exists a sequence {fn,gn,ϵn} where fn:(Xn,n)(X,) and gn:(X,)(Xn,n) are pointed Borel maps, ϵn0, and the following properties are satisfied:

  • (i)

    for every xXn, dn(n,x)ϵn-1dn(gn(fn(x)),x)ϵn, and, for every xX, d(,x)ϵn-1d(fn(gn(x)),x)ϵn,

  • (ii)

    max{Dis(fn|Bn(ϵn-1)),Dis(gn|B(ϵn-1))}ϵn (see Notation 2.2),

  • (iii)

    max{dPϵn-1(fnmn,m),dPϵn-1(gnm,mn)}ϵn.

Such a sequence is said to realize the convergence of {Xn} to X in the pmGH topology.

We conclude this section by introducing the moduli space of pointed RCD(0,N)-structures on X.

Notation 2.3

We introduce the following spaces:

  • (i)

    RCDp(0,N) is the set of isomorphism classes of pointed RCD(0,N) spaces with full support, endowed with the pmGH-topology (seen as a subspace of Xp),

  • (ii)

    R0,Np(X) is the set of all pointed RCD(0,N)-structures on X,

  • (iii)

    M0,Np(X) is the quotient of R0,Np(X) by isomorphisms, endowed with the pmGH-topology (seen as a subspace of RCDp(0,N)).

We call M0,Np(X) the moduli space of pointed RCD(0,N)-structures on X.

Moduli space of equivariant pointed RCD(0,N)-structures

First of all, we introduce equivariant pointed RCD(0,N)-structures on X~. Here, in comparison with the definition of equivariant metric given by Fukaya and Yamaguchi in [20], both the topological space and the group action are fixed.

Definition 2.6

A pointed RCD(0,N)-structure (X~,d~,m~,~) on X~ is called equivariant if π¯1(X) acts by isomorphisms on (X~,d~,m~).

The following definition introduces equivariant isomorphisms between equivariant pointed RCD(0,N)-structures on X~.

Definition 2.7

For i{1,2}, let X~i=(X~,d~i,m~i,~i) be an equivariant pointed RCD(0,N)-structure on X~. We say that X~1 and X~2 are equivariantly isomorphic when there is an isomorphism ϕ of π¯1(X) and an isomorphism f:X~1X~2 of p.m.m.s. such that f(γx)=ϕ(γ)f(x), for every γπ¯1(X), and every xX~.

We now introduce the space (and moduli space) of equivariant pointed RCD(0,N)-structures on X~.

Notation 2.4

We introduce the following spaces:

  • (i)

    R0,Np,eq(X~) the set of equivariant pointed RCD(0,N)-structures on X~,

  • (ii)

    M0,Np,eq(X~) the quotient space of R0,Np,eq(X~) by equivariant ismormophisms.

We call M0,Np,eq(X~) the moduli space of equivariant pointed RCD(0,N)-structures on X~.

To define a topological structure on M0,Np,eq(X~) , we start by introducing the equivariant pointed distance on R0,Np,eq(X~).

Definition 2.8

Let ϵ>0, and, for i{1,2}, let X~i=(X~,d~i,m~i,~i) be an equivariant pointed RCD(0,N)-structure on X~. An equivariant pointed ϵ-isometry between X~1 and X~2 is a triple (f,g,ϕ) where f:X~X~ and g:X~X~ are Borel maps and ϕ is an isomorphism of π¯1(X) such that:

  • (i)

    f(~1)=~2 and g(~2)=~1,

  • (ii)

    for every γπ¯1(X) and xX~, f(γx)=ϕ(γ)f(x) and g(γx)=ϕ-1(γ)g(x),

  • (iii)

    for every x,yX~, d~1(x,y)ϵ-1|d~2(f(x),f(y))-d~1(x,y)|ϵ, and d~2(x,y)ϵ-1|d~1(g(x),g(y))-d~2(x,y)|ϵ,

  • (iv)

    for every xX~, d~1(gf(x),x)ϵ and d~2(fg(x),x)ϵ,

  • (v)

    max{dPϵ-1(fm~1,m~2),dPϵ-1(gm~2,m~1)}ϵ.

We define Dpeq(X~1,X~2) the equivariant pointed distance between X~1 and X~2 as the minimum between 1/24 and the infimum of all ϵ>0 such that there exists an equivariant pointed ϵ-isometry between X~1 and X~2.

The following result shows that we can endow M0,Np,eq(X~) with a metrizable topology.

Proposition 2.8

Dpeq induces a metrizable uniform structure on M0,Np,eq(X~).

Proof

See Appendix.

From now on, we endow M0,Np,eq(X~) with the topology induced by Dpeq, which we call the equivariant pmGH-topology.

Maps between moduli spaces

In Sect. 2.4.1, we are going to introduce the lift and push-forward maps. As we will explain at the end of that section, a consequence of Theorem A is that these maps are homeomorphisms and respectively inverse to each other (see Corollary A). Then, in Sect. 2.4.2, we will introduce the Albanese map and the soul map associated to X.

Lift and push-forward maps

Thanks to Corollary 2.1, we can define the lift of a pointed RCD(0,N)-structure.

Definition 2.9

Let (X,d,m,x) be a pointed RCD(0,N)-structure on X and let x~p-1(x). We define px~(X,d,m,x):=(X~,d~,m~,x~), where (X~,d~,m~) is the lift of (X,d,m).

Remark 2.3

For i{1,2}, let (X,di,mi,xi) be a pointed RCD(0,N)-structure on X, and let x~ip-1(xi). If (X,d1,m1,x1) and (X,d2,m2,x2) are isomorphic, then px~1(X1,d1,m1,x1) is equivariantly isomorphic to px~2(X2,d2,m2,x2).

Thanks to Remark 2.3, we can define the lift map associated to X.

Definition 2.10

(Lift map) The lift map associated to X is the unique map p:M0,Np(X)M0,Np,eq(X~) that satisfies p[X,d,m,x]=[px~(X,d,m,x)] for every (X,d,m,x)R0,Np(X) and x~p-1(x).

Thanks to Proposition 2.3, we can define the push-forward of an equivariant pointed RCD(0,N)-structure.

Definition 2.11

Let (X~,d~,m~,x~) be an equivariant pointed RCD(0,N)-structure on X~. We define p(X~,d~,m~,x~) as the unique pointed RCD(0,N)-structure on X such that p:(X~,d~,m~,x~)p(X~,d~,m~,x~) is a pointed local isomorphism.

Remark 2.4

For i{1,2}, let (X~,d~i,m~i,x~i) be an equivariant pointed RCD(0,N)-structure on X~. If (X~,d~1,m~1,x~1)(X~,d~2,m~2,x~2), then p(X~,d~1,m~1,x~1) is isomorphic to p(X~,d~2,m~2,x~2).

Thanks to Remark 2.4, we can define the push-forward map associated to X.

Definition 2.12

(Push-forward map) The push-forward map associated to X is the unique map

p:M0,Np,eq(X~)M0,Np(X)

satisfying p[X~,d~,m~,x~]=[p(X~,d~,m~,x~)] for every (X~,d~,m~,x~)R0,Np,eq(X~).

Thanks to Remark 2.1, we have the following proposition.

Proposition 2.9

The lift map p:M0,Np(X)M0,Np,eq(X~) and the push-forward map p:M0,Np,eq(X~)M0,Np(X) are respectively inverse to each other.

Observe that Corollary A immediately follows from Proposition 2.9 and Theorem A (which we will prove in Sect. 3.1).

Albanese and soul maps

First of all, we introduce the moduli space of flat metrics on the Albanese class A(X) (introduced in Definition 2.3). This moduli space will act as the codomain of the Albanese map.

Definition 2.13

(Mflat(A(X))) The moduli space of flat metrics on A(X) is the quotient of A(X) by isometry equivalence, endowed with the Gromov–Hausdorff distance dGH (see Definition 7.3.10 in [11]).

The following remark will be helpful in the proof of Theorem C. It is also interesting on its own as it gives a more explicit way to see the moduli space Mflat(A(X)).

Remark 2.5

Given any element ΓΓ(X), the moduli space of flat metrics on A(X) is isometric to the moduli space of flat metrics on the compact orbifold Rk/Γ (endowed with the Gromov–Hausdorff distance), which we denote Mflat(Rk/Γ) (see Section 4.2 in [8] for more details on Mflat(Rk/Γ)).

The next lemma is fundamental to introduce the Albanese and soul maps associated to X.

Lemma 2.1

For i{1,2}, let (X,di,mi) be an RCD(0,N)-structure on X, and let ϕi be a splitting of its lift (X~,d~i,m~i) with soul (X¯i,d¯i,m¯i). If (X,d1,m1) and (X,d2,m2) are isomorphic, then (X¯1,d¯1,m¯1) is isomorphic to (X¯2,d¯2,m¯2), and (Rk/Γ(ϕ1),dΓ(ϕ1)) is isometric to (Rk/Γ(ϕ2),dΓ(ϕ2)).

Proof

Let us fix an isomorphism ϕ:(X,d1,m1)(X,d2,m2). We can lift ϕ to the universal covers to get an isomorphism ϕ~:(X~,d~1,m~1)(X~,d~2,m~2) such that pϕ~=ϕp. Now, let μ:=ϕ2ϕ~ϕ1-1. Since, X¯1 and X¯2 are compact, μ is of the form μ=(μS,μR), where μS:(X¯,d¯1,m¯1)(X¯,d¯2,m¯2) is an isomorphism, and μRIso(Rk) (where k:=k(X)). In particular (X¯1,d¯1,m¯1) is isomorphic to (X¯2,d¯2,m¯2).

We are going to show that Γ(ϕ2)=μRΓ(ϕ1)μR-1. Let x¯1X¯1, let tRk, let απ¯1(X) and define z~:=ϕ1-1(x¯1,t). By definition of the soul and Euclidian homomorphisms associated to ϕ1 and ϕ2, we have:

μ(ρSϕ1(α)·x¯1,ρRϕ1(α)·t)=(ρSϕ2(η)·μS(x¯1),ρRϕ2(η)·μR(t))=(μS(ρSϕ1(α)·x¯1),μR(ρRϕ1(α)·t)),

where η:=ϕ~(α), and ϕ~ is the automorphism of π¯1(X) defined by ϕ~(α):=ϕ~αϕ~-1. In particular, for every tRk and απ¯1(X), we have μR(ρRϕ1(α)·t)=ρRϕ2ϕ~(α)·μR(t). Thus, for every απ¯1(X), we have μRρRϕ1(α)μR-1=ρRϕ2ϕ~(α). In particular, by definition of Γ(ϕ1) and Γ(ϕ2), and since ϕ~(π¯1(X))=π¯1(X), we have Γ(ϕ2)=μRΓ(ϕ1)μR-1. In conclusion, using Lemma 4.1 in [8], (Rk/Γ(ϕ1),dΓ(ϕ1)) is isometric to (Rk/Γ(ϕ2),dΓ(ϕ2)), which concludes the proof.

Thanks to Lemma 2.1, we can define the Albanese and soul maps.

Definition 2.14

(Albanese and soul maps) Given an RCD(0,N)-structure (X,d,m) on X, and given a splitting ϕ of (X~,d~,m~) with soul (X¯,d¯,m¯), we define:

A([X,d,m]):=[Rk/Γ(ϕ),dΓ(ϕ)]Mflat(A(X)),

and:

S([X,d,m]):=[X¯,d¯,m¯]RCD(0,N-k(X)).

The map A:M0,N(X)Mflat(A(X)) is called the Albanese map associated to X, and the map

S:M0,N(X)RCD(0,N-k(X))

is called the soul map associated to X.

We end this section with the following surjectivity result.

Proposition 2.10

The Albanese map associated to X is surjective from M0,N(X) onto Mflat(A(X)).

Proof

First of all, let (X,d0,m0) be a reference RCD(0,N)-structure on X, and let ϕ0 be a splitting of its lift (X~,d~0,m~0) with soul (X¯0,d¯0,m¯0). Now, let ΓΓ(X) and let us show that there is some (X,d,m)R0,N(X) such that A([X,d,m])=[Rk/Γ,dΓ].

Since Γ(ϕ0)Γ(X), there is αAff(Rk) such that Γ=αΓ(ϕ0)α-1. Now, let ψ:=(idX¯0,α)ϕ0, and consider the metric measure structure (d~,m~) defined as the pull back by ψ of (d¯0×deucli,m¯0Lk). Note that ψ is a homeomorphism, and (X¯0×Rk,d¯0×deucli,m¯0Lk) is an RCD(0,N) space; hence, (X~,d~,m~) is an RCD(0,N)-structure on X~.

Now, we are going to show that (X~,d~,m~) is the lift of some (X,d,m). Thanks to Remark 2.1, it is equivalent to show that π¯1(X)Isom.m.s.(X~,d~,m~), which is itself equivalent to ψ(π¯1(X))Isom.m.s.(X¯0×Rk,d¯0×deucli,m¯0Lk). Let ηπ¯1(X), then ψ(η)=ψηψ-1=(idX¯0,α)ϕ0(η)(idX¯0,α-1)=(ρSϕ0(η),αρRϕ0(η)α-1). Note that ρSϕ0(η)Isom.m.s.(X¯0,d¯0,m¯0) and αρRϕ0(η)α-1αΓ(ϕ0)α-1=ΓIso(Rk); hence, ψ(η)Isom.m.s.(X¯0×Rk,d¯0×deucli,m¯0Lk). In conclusion, there is an RCD(0,N)-structure (X,d,m)R0,N(X) whose lift is (X~,d~,m~). By construction, ψ is a splitting of (X~,d~,m~) with soul (X¯0,d¯0,m¯0). Moreover, we have seen above that, for every ηπ¯1(X), we have ρRψ(η)=pRψψ(η)=αρRϕ0(η)α-1. Hence, Γ(ψ)=αΓ(ϕ0)α-1=Γ, and we get A([X,d,m])=[Rk/Γ,dΓ].

Proof of the main results

Proof of Theorem A

First of all, let us introduce the systole associated to an RCD(0,N)-structure on X. Finding a uniform lower bound on the systoles associated to a sequence will be the key to prove Theorem A.

Definition 3.1

(Systole of an RCD(0,N)-structure) The systole associated to an RCD(0,N)-structure (X,d,m) on X is the quantity sys(X,d):=inf{d~(η·x~,x~)}, where the infimum is taken over all point x~X~ and ηπ¯1(X)\{id}. Whenever π¯1(X) is trivial, we define sys(X,d):=.

The following proposition relates the systole of an RCD(0,N)-structure (X,d,m) on X and the quantity δ(X,d) introduced in Theorem 2.1.

Proposition 3.1

Let (X,d,m) be an RCD(0,N)-structure on X. Then, sys(X,d)=2δ(X,d), where δ(X,d) is defined in Theorem 2.1.

Proof

Let δ<δ(X,d), let ηπ¯1(X)\{id}, and let x~X~. Then, by Proposition 2.2 and Theorem 2.1, p induces a homeomorphism from Bd~(x~,δ) (resp. Bd~(η·x~,δ)) onto Bd(x,δ), where x:=p(x~). Seeking for a contradiction, assume that there exists y~Bd~(x~,δ)Bd~(η·x~,δ). Then, d(η·y~,η·x~)=d(y~,x~)<δ. In particular, y~ and η·y~ are two distinct elements of Bd~(η·x~,δ) which have the same image under p, which is the contradiction we were looking for. Hence, Bd~(x~,δ)Bd~(η·x~,δ)=. In particular, d~(η·x~,x~)2δ; thus, 2δsys(X,d). Since that holds for every δ<δ(X,d), we have 2δ(X,d)sys(X,d).

Now assume that δ(X,d)<δ. Then, there is some xX such that Bd(x,δ) is not evenly covered by p. Therefore, given any x~p-1(x), there exists y~iBd~(x~,δ) (i{1,2}) such that py~1=py~2, y~1y~2. Hence, there exists γπ¯1(X)\{id} such that y~2=γy~1; thus, d~(y~1,y~2)=d~(γy~1,y~1)2δ. Therefore, we have sys(X,d)2δ. Thus, letting δ go to δ(X,d), we finally obtain sys(X,d)2δ(X,d).

The next result shows that we can find a positive uniform lower bound on the systoles associated to a converging sequence of RCD(0,N)-structures on X.

Proposition 3.2

Assume that {(X,dn,mn)} converges to (X,d,m) in the mGH-topology, where, for every nN{}, (X,dn,mn) is an RCD(0,N)-structure on X. Then 0<infnN{δ(X,dn)}.

Proof

First of all, observe that by Theorem 2.1, δ(X,dn)>0 for every nN. In particular, it is sufficient to prove that there exists a constant δ>0 such that δ(X,dn)δ whenever n is large enough.

We define ϵn:=dGH((X,dn),(X,d))0, δ2:=δ(X,d)/2, and δ1:=δ(X,d)/3. Whenever n is large enough, we have δ1>20ϵn and δ2>δ1+10ϵn; hence, by Theorem 3.4 of [33], there is a surjective group homomorphism ψn:G(δ1,dn)G(δ2,d). Moreover, since δ2<δ(X,d), then G(δ2,d) is isomorphic to π¯1(X) by Proposition 2.1. Now, fixing x~X~, and x1Xdnδ1, such that x:=p(x~)=pdnδ1(x1), we have a surjective homomorphism:

q:π1(X,x)/pπ1(X~,x~)(π1(X,x)/pπ1(X~,x~))/(pdnδ1π1(Xdnδ1,x1)/pπ1(X~,x~)).

However, the domain of q is isomorphic to π¯1(X), whereas its codomain is isomorphic to G(δ1,dn). Therefore, q gives rise to a surjective homomorphism νn from π¯1(X) onto G(δ1,dn). Hence, we have a surjective group homomorphism: graphic file with name 208_2022_2493_Figc_HTML.jpg However, π¯1(X) is a Hopfian group by Proposition 2.6; thus the homomorphism above has to be an isomorphism. In particular, νn:π¯1(X)G(δ1,dn) has to be injective; hence, it is an isomorphism, and it implies that q is also an isomorphism. In particular, we necessarily have pdnδ1π1(Xdnδ1,x1)=pπ1(X~,x~); hence, by the classification Theorem (see Theorem 2, Chapter 2, Section 5 in [34]), (Xdnδ1,X,pdnδ1) is equivalent to (X~,X,p). In particular, every ball of radius δ1 in (X,dn) is evenly covered by p; thus, δ(X,dn)δ1, which concludes the proof.

The following proposition is a converse to Proposition 3.2; it will be essential to prove the converse implication of Theorem A.

Proposition 3.3

Assume that {(X~,d~n,m~n,~n)} converges to (X~,d~,m~,~) in the equivariant pmGH-topology, where, for every nN{}, (X~,d~n,m~n,~n) is an equivariant pointed RCD(0,N)-structure on X~. Then 0<infnN{δ(X,dn)}, where (X,dn,mn) is the push-forward of (X~,d~n,m~n).

Proof

We fix a sequence {(f~n,g~n,ϕn,ϵn)} realizing the equivariant pointed convergence. Looking for a contradiction, assume that infnN{sys(X,dn)}=0. Without loss of generality, we can assume (passing to a subsequence if necessary) that there exist sequences {x~n} in X~ and {γn} in π¯1(X)\{id} such that d~n(γnx~n,x~n)0. However, when n is large enough so that d~n(γnx~n,x~n)ϵn-1, we have sys(X,d)d~(f~n(x~n),ϕn(γn)f~n(x~n))=d~(f~n(x~n),f~n(γnx~n))d~n(γnx~n,x~n)+ϵn0. Therefore, sys(X,d)=0=δ(X,d) (using Proposition 3.1), which is the contradiction we were looking for. Hence 0<infnN{sys(X,dn)}; therefore, thanks to Proposition 3.1, we have 0<infnN{δ(X,dn)}.

We can now prove Theorem A.

Proof of Theorem A, direct implication

Assume that {Xn=(X,dn,mn,n)} converges in the pmGH-topology to X=(X,d,m,). Let us prove that {X~n} converges in the equivariant pmGH-topology to X~.

Part I: Construction of the realizing sequence{f~n,g~n,ψn,ϵn}

First of all, we fix a sequence {fn,gn,ϵn} realizing the convergence of {Xn} to X in the pmGH-topology. Then, we define δ:=infnN{}{δ(X,dn)}, which satisfies δ>0 thanks to Proposition 3.2. By Proposition 3.1, we have μ0:=infnN{}{sys(X,dn)}=2δ>0. We define α:=δ/2, and we assume that n is large enough so that:

5ϵn<α<μ0/2-3ϵn/2. 3

Theorem 2.1 implies that (X~,d~,m~) is isomorphic to (Xdα,d,α,m,α), since α<δ(X,d). Now, thanks to Theorem 16 of [31] (and the construction in its proof), there exists a triple (f~n,g~n,ψn) such that:

  • f~n:(X~,~n)(X~,~) (resp. g~n:(X~,~)(X~,~n)) satisfy pf~n=fnp (resp. pg~n=gnp),

  • for every x~X~, we have d~n(g~nf~n(x~),x~)ϵn and d~(f~ng~n(x~),x~)ϵn,

  • for every x~X~ and ηπ¯1(X), we have f~n(η·x~)=ψn(η)·f~n(x~) and g~n(η·x~)=ψn-1(η)·g~n(x~).

Moreover, using inequality 3, Theorem 16 of [31] assures that, for every x~,y~X~, we have:

|d~(f~n(x~),f~n(y~))-d~n(x~,y~)|3ϵn(d~n(x~,y~)/α+1),

and:

|d~n(g~n(x~),g~n(y~))-d~(x~,y~)|3ϵn(d~(x~,y~)/α+1).

We fix C>0 such that C+3/αC2, and we define ϵn:=Cϵn. When n is large enough so that ϵnϵn, we have:

(f~n,g~n,ψn,ϵn)satisfies point (i) to (iv) of Definition2.8w.r.t.X~nandX~. 4

Let us prove that, when n is large enough, f~n and g~n are Borel maps. Let x~X~, and let r<δ/3. Thanks to Proposition 2.2 and property 4, we easily get:

f~n-1(B~(x~,r))=(fnp)-1(B(x,r))B~n(g~n(x~),r+2ϵn),

when n is large enough so that δ/3+4ϵn<δ/2<ϵn-1, and where x:=p(x~). However, fn is a Borel map, and p is continuous; therefore f~n-1(B~(x~,r)) is a Borel subset of X~. We have shown that when n is large enough, the pre-image by f~n of balls of radius r<δ/3 are Borel subsets of X~. Therefore, for n large enough, f~n is a Borel map, and the same is true for g~n with the same procedure.

Part II: Measured convergence

Our goal here is to prove that (making ϵn larger if necessary but keeping ϵn0), we have:

max{dP{ϵn-1}(f~nm~n,m~),dP{ϵn-1}(g~nm~,m~n)}ϵn.

This is implied by the fact that, {dP{R}(f~nm~n,m~)} and {dP{R}(g~nm~,m~n)} converge to 0 as n goes to infinity, for every R>0.

First of all, observe that limndP{R}(f~nm~n,m~)=0 for every R>0 if and only if {f~nm~n} converges to m~ in the weak- topology. Then, note that the space Mloc(X~,d~) of Radon measures on (X~,d~) endowed with the weak- topology is metrizable (see Theorem A2.6.III in [17]). Hence, it is sufficient to show that any subsequence of {f~nm~n} admits a subsequence converging to m~. Without loss of generality (reindexing the sequence if necessary), let us just show that {f~nm~n} admits a subsequence converging to m~.

First, let us show that {f~nm~n} is precompact, which is implied by the uniform boundedness of

{f~nm~n(Bd~(R))},

for every R>0 (see Theorem A2.6.IV and Theorem A2.4.I in [17]). We define

r0:=infnN{}{δ(X,dn)}/2andM:=supnN{}{mn(X)}.

Observe that r0 is positive thanks to Proposition 3.2, and M is finite since {mn(X)} converges to m(X), which is finite. Thanks to point (v) of Proposition 2.2, we have m~n(Bd~n(r0))=mn(Bdn(r0))M, for every nN. Then, thanks to property 4, we have f~n-1(Bd~(R))Bd~n(2R), for every R>0, and n sufficiently large. Now, consider the following two cases:

  • if Rr0/2, we get f~nm~n(Bd~(R))m~n(Bd~n(2R))m~n(Bd~n(r0))M, when n is sufficiently large,

  • if R>r0/2, thanks to Bishop–Gromov inequality for RCD(0,N) spaces (see Theorem 6.2 in [6]), we get f~nm~n(Bd~(R))m~n(Bd~n(2R))M(2R/r0)N, when n is sufficiently large.

In particular, for every R>0, the sequence {f~nm~n(Bd~(R))} is uniformly bounded; hence {f~nm~n} is precompact.

Now, passing to a subsequence if necessary, we can assume that {f~nm~n} is converging to some mMloc(X~,d~). Let us show that m=m~. Note that it is sufficient to prove that, for every x~X~ and 0<r<r0, we have m(Bd~(x~,r))=m~(Bd~(x~,r)); since small balls generate the Borel σ-algebra of X~.

First, observe that, for every nN, we have m~n(Bd~n(r0))=mn(Bdn(r0))m, where

m:=infnN{}{mn(Bdn(r0))}.

In addition, m is positive since {mn(Bdn(r0))} is a sequence of positive numbers converging to m(Bd(r0)), which is positive. Therefore, {X~n} is a sequence of pointed RCD(0,N) spaces with measures uniformly bounded from below; hence (thanks to Theorem 7.2 in [23]), any limit point in the pmGH-topology is a full support RCD(0,N) space. However, the sequence converges in the pmGH-topology to (X~,d~,m,~). Thus, (X~,d~,m,~) is a full support RCD(0,N) space. In particular, thanks to Theorem 30.11 in [39], we have m(Bd~(x~,R))=0 for every R>0 and x~X~. Hence, thanks to Proposition A2.6.II in [17], for every R>0 and x~X~ we have:

m(Bd~(x~,R))=limnf~nm~n(Bd~(x~,R)), 5

Now, let x~X~ and 0<r<r0, and let us show that we have m(Bd~(x~,r))=m~(Bd~(x~,r)). First, when n is large enough so that rϵn-1, we can use property 4 to get:

Bd~n(g~n(x~),r-2ϵn)f~n-1(Bd~(x~,r))Bd~n(g~n(x~),r+2ϵn)).

In particular, defining A:=m(Bd~(x~,r)) and using equation 5, we have:

lim supnm~n(Bd~n(g~n(x~),r-2ϵn))Alim infnm~n(Bd~n(g~n(x~),r+2ϵn)).

Moreover, when n is large enough, we have r+2ϵn<r0<δ/2; hence, point (v) of Proposition 2.2 implies:

lim supnmn(Bdn(gn(x),r-2ϵn))Alim infnmn(Bdn(gn(x),r+2ϵn)),

where x:=p(x~). Now, observe that when n is large enough so that r+4ϵnϵn-1, we can use property 4 to get:

Bdn(gn(x),r+2ϵn)fn-1(Bd(x,r+4ϵn)),fn-1(Bd(x,r-4ϵn))Bdn(gn(x),r-2ϵn).

In particular, for every η>0, we have:

lim supnfnmn(Bd(x,r-η))Alim infnfnmn(Bd(x,r+η)).

However, since {fnmn} converges to m, and since X is a full support RCD(0,N) space, we can apply Theorem 30.11 in [39] and Proposition A2.3.II in [17] to get:

lim supnfnmn(Bd(x,r-η))=m(Bd(x,r-η)),lim infnfnmn(Bd(x,r+η))=m(Bd(x,r+η)).

Hence, for every η>0, we have:

m(Bd(x,r-η))m(Bd~(x~,r))m(Bd(x,r+η));

and, letting η go to 0, we have m(Bd~(x~,r))=m(Bd(x,r))=m~(Bd~(x~,r)) (using r<r0<δ/2 for the last equality). Therefore {f~nm~n} converges to m~.

For every R>0, we define ϵ(n,R):=dP{R}(f~nm~n,m~). Thanks to the discussion above, we have limnϵ(n,R)0, for every R>0. Let R>0, and let us show that limndP{R}(g~nm~,m~n)=0.

Let AB¯d~n(R), and observe that we have m~n(A)f~nm~n(f~n(A)). Also, when n is large enough, we can use property 4 to get f~n(A)B¯d~(2R). Therefore, we have m~n(A)m~({f~n(A)}ϵ(n,2R))+ϵ(n,2R). Then, when n is large enough, we can use property 4 to obtain {f~n(A)}ϵ(n,2R)g~n-1({A}2ϵn+ϵ(n,2R)). Thus, we have m~n(A)g~nm~({A}2ϵn+ϵ(n,2R))+ϵ(n,2R). Applying the same arguments, we also have g~nm~(A)m~n({A}2ϵn+ϵ(n,2R))+ϵ(n,2R). Therefore, dP{R}(g~nm~,m~n)ϵ(n,2R)+2ϵn; in particular, limndP{R}(g~nm~,m~n)=0. This concludes the proof.

Proof of Theorem A, converse implication

Assume that {X~n=(X~,d~n,m~n,~n)} converges in the equivariant pmGH-topology to X~=(X~,d~,m~,~). Let us prove that {Xn=(X,dn,mn,n)} converges in the pmGH-topology to X=(X,d,m,).

Let {f~n,g~n,ϕn,ϵn} be a sequence realizing the convergence of {X~n} to X~ in the equivariant pmGH-topology. Thanks to the equivariant requirement, there exists pointed Borel maps fn:(X,n)(X,) and gn:(X,)(X,n) such that pf~n=fnp and pg~n=gnp.

Let us fix xX and x~p-1(x). Observe that dn(gn(fn(x)),x)=inf{d~n(y~,x~)}, where the infimum is taken over all y~X~ such that p(y~)=gn(fn(x)). However, we have p(g~n(f~n(x~)))=gn(fn(x)). Therefore, we have dn(gn(fn(x)),x)d~n(g~n(f~n(x~)),x~)ϵn. The same argument shows that d(fn(gn(x)),x)ϵn.

Let yiX (i{1,2}) and let y~i such that p(y~i)=yi and d~(y~1,y~2)=d(y1,y2). Assume that D:=Diam(X,d)ϵn-1 and observe that, since p(g~n(y~i))=gn(yi), we have:

dn(gn(y1),gn(y2))-d(y1,y2)d~n(g~n(y~1),g~n(y~2))-d~(y~1,y~2) 6
ϵn. 7

Then, let xiX (i{1,2}) such that dn(x1,x2)ϵn-1, and fix x~iX~ such that p(x~i)=xi and d~n(x~1,x~2)=dn(x1,x2). Observe that we have p(f~n(x~i))=fn(xi), therefore:

d(fn(x1),fn(x2))-dn(x1,x2)d~(f~n(x~1),f~n(x~2))-d~n(x~1,x~2) 8
ϵn. 9

Let us show that {Dn:=Diam(X,dn)} is a bounded sequence. Let xiX (i{1,2}) and observe that thanks to inequality 6, we have dn(gn(fn(x1)),gn(fn(x2)))ϵn+D (when Dϵn-1). However, we have |dn(x1,x2)-dn(gn(fn(x1)),gn(fn(x2)))|dn(gn(fn(x1)),x1)+dn(gn(fn(x2)),x2)2ϵn. Therefore, dn(x1,x2)D+3ϵn. We can conclude that {Dn} is bounded.

Since {Dn} is bounded, we have (thanks to inequality 8):

x1,x2X,d(fn(x1),fn(x2))-dn(x1,x2)ϵn,

when n is large enough. Also, we have dn(x1,x2)dn(gn(fn(x1)),gn(fn(x2)))+2ϵn. Therefore, using inequality 6 we obtain, dn(x1,x2)-d(fn(x1),fn(x2))3ϵn. Hence, we can conclude that Dis(fn)3ϵn. The same argument also gives Dis(gn)3ϵn, which concludes the proof of the second metric requirement.

Finally, using Lemma 3.3, and applying exactly the same procedure as in Part II of the direct implication, we can prove that (making ϵn smaller if necessary but keeping ϵn0) we have:

max{dP(fnmn,m),dP(gnm,mn)}ϵn.

Hence, {fn,gn,ϵn} is a sequence realizing the convergence of {Xn} to X in the pmGH-topology, which concludes the proof.

Proof of Theorem B

In this section, we give a proof of Theorem B. Let {(X,dn,mn)} be a sequence converging in the mGH-topology to (X,d,m), where for every nN{}, (X,dn,mn) is an RCD(0,N)-structure on X. For every nN{}, we fix ϕn a splitting of (X~,d~n,m~n) with soul (X¯n,d¯n,m¯n), and we denote k:=k(X) the splitting degree of X (see Corollary 2.2 for the definition of the splitting degree). To conclude, we are going to prove that:

{X¯n,d¯n,m¯n}converges to(X¯,d¯,m¯)in the mGH-topology, 10

and:

{Rk/Γ(ϕn),dΓ(ϕn)}converges to(Rk/Γ(ϕ),dΓ(ϕ))in the GH-topology. 11

Observe that, since X is compact, we can find a family {n}nN{} of points in X such that {(X,dn,mn,n)} converges to (X,d,m,) in the pmGH-topology. Then, for every nN{}, let us fix ~np-1(n). Observe that without loss of generality, we can assume that, for every nN{}, we have pRk(ϕn(~n))=0. For every nN{}, we denote:

  • Xn:=(X,dn,mn,n),

  • X~n:=(X~,d~n,m~n,~n)=p~n(Xn),

  • X¯n:=(X¯n,d¯n,m¯n,¯n), where ¯n:=pX¯n(ϕn(~n)).

Thanks to Theorem A, {X~n} converges to X~ in the equivariant pmGH-topology. Thanks to the proof of Theorem A, there exists a sequence {fn,gn,ϵn} (resp. {f~n,g~n,ϕn,ϵn}) realizing the convergence of {Xn} to X (resp. of {X~n} to X~) in the pmGH-topology (resp. in the equivariant pmGH-topology), such that pf~n=fnp and pg~n=gnp. Finally, for every nN, we define:

  • kn:=ϕf~nϕn-1, knR:=pRkkn(¯n,·), and knS:=pX¯kn(·,0),

  • ln:=ϕng~nϕ-1, lnR:=pRkln(¯,·), and lnS:=pX¯nln(·,0).

The main difficulty of the argument will be to prove that kn and ln almost split. More precisely, we will show that kn(knS,knR) and ln(lnS,lnR) (where we will give a precise meaning to ). Then, we will deduce property 10 and property 11 from that.

First of all, we prove that {Diam(X¯n,d¯n)} is bounded.

Proposition 3.4

The sequence {Diam(X¯n,d¯n)} is bounded.

Proof

Looking for a contradiction, let us suppose that lim supnDiam(X¯n,d¯n)=. Passing to a subsequence if necessary, we can assume that Diam(X¯n,d¯n)>2n+1, for every nN. Hence, there are sequences {x¯n} and {z¯n} such that, for every nN, we have x¯n,z¯nX¯n, and d¯n(x¯n,z¯n)=2n+1.

For every nN, let γ¯n:[-2n,2n]X¯n be a minimizing geodesic parametrized by arc length from x¯n to z¯n, and let us denote γ~n:=(γ¯n,0). Thanks to Proposition 2.4, there exists ηπ¯1(X) such that ηγ~n(0)BX¯n×Rk((¯n,0),D), where D:=supnN{}{Diam(X,dn)}< (D being finite because (X,dn) converges to (X,d) in the GH-topology).

Then, let us define β~n:=ηγ~n, and denote β¯n:=pX¯n(β~n), and vn:=pRk(β~n)=ρRϕn(η)(0). Observe that the sequence {β~n} consists of isometric embeddings such that β~n(0)BX¯n×Rk((¯n,0),D). Moreover, {kn,ln} realizes the convergence of {X¯n×(Rk,0)} to X¯×(Rk,0) in the pmGH-topology. Therefore, thanks to Arzelà–Ascoli Theorem (see Proposition 27.20 in [39]), we can assume (passing to a subsequence if necessary) that {knβ~n} converges locally uniformly to an isometric embedding β~:RX¯×Rk. However, (X¯,d¯) is compact; thus, applying Lemma 1 of [32], there exists a,bRk and y¯X¯ such that, for every tR, β~(t)=(y¯,at+b), and a=1.

Now, we define y¯n:=β¯n(0) and, for uRk, Φn(u):=(y¯n,u)X¯n×Rk. Observe that {Φn} is a sequence of isometric embeddings such that, for every nN, we have Φn(0)BX¯n×Rk((¯n,0),D). Therefore, thanks to Arzelà–Ascoli Theorem (see Proposition 27.20 in [39]), we can assume (passing to a subsequence if necessary) that {knΦn} converges locally uniformly to an isometric embedding Φ:RkX¯×Rk. Moreover, since X¯ is compact, we can easily deduce from Lemma 1 of [32] that there exists ϕIso(Rk) and z¯X¯ such that Φ(t)=(z¯,ϕ(t)), for every tRk.

Notice that β~n(0)=(y¯n,vn)=Φn(vn). Moreover, observe that |vn|D; hence, passing to a subsequence if necessary, we can assume that vnvRk. Thus, we have dX¯×Rk(Φ(v),β~(0))=limndX¯×Rk(knΦn(vn),knβ~n(0))=0. In particular, y¯=z¯, and ϕ(v)=b. Now, let cRk such that [ϕ-ϕ(0)](c)=a; thus, we have Φ(ct+v)=at+b=β~(t), for every tR.

Now, observe that we have 0=dX¯×Rk(Φ(c+v),β~(1)). Therefore:

dX¯n×Rk(Φn(c+vn),β~n(1))=(1+c2)1/20.

Therefore, we have 0=(1+c2)1/2>0, which is the contradiction we were looking for.

Thanks to Proposition 3.4, we can introduce the following notations.

Notation 3.1

We denote D:=supnN{Diam(X,dn)}< (finiteness being granted by the convergence of {X,dn} to (X,d) in the GH-topology). We also denote D¯:=supnN{}{Diam(X¯n,d¯n)}<.

Our first goal will be to obtain a convergence result on the following “splitting quantities”.

Notation 3.2

(Splitting quantities) Given nN and R>0, we define:

  • (i)

    α(n,R):=sup{dX¯×Rk(kn(y¯n,t),(knS(y¯n),knR(t)))}, the supremum being taken over y¯nX¯n and |t|R,

  • (ii)

    β(n,R):=sup{dX¯n×Rk(ln(y¯,t),(lnS(y¯),lnR(t)))}, the supremum being taken over y¯X¯ and |t|R.

The following next two technical lemmas will be our main ingredients in the proof of the convergence result of the splitting quantities.

Lemma 3.1

Let {y¯n} be a sequence such that, for every nN, y¯nX¯n. For every nN, and tRk, we define Φn:tRk(y¯n,t)X¯n×Rk. Then, the sequence of maps {knΦn:RkX¯×Rk} admits a subsequence converging locally uniformly to a map Φ:RkX¯×Rk. Moreover, for any such limit Φ, there exists y¯, and ϕOk(R) such that tRk,Φ(t)=(y¯,ϕ(t)).

Proof

Observe that, for every nN, Φn is an isometric embedding that satisfies

Φn(0)BX¯n×Rk((¯n,0),D¯).

Therefore, applying Arzelà–Ascoli Theorem (see Proposition 27.20 in [39]) as in the proof of Proposition 3.4, we can assume without loss of generality that {knΦn} converges locally uniformly to an isometric embedding Φ:RkX¯×Rk. Moreover, using Lemma 1 of [32], there exist ϕIso(Rk) and y¯X¯ such that Φ(t)=(y¯,ϕ(t)), for every tRk. To conclude, we need to show that ϕ(0)=0. First, observe that:

d¯(¯,y¯)dX¯×Rk((¯,0),Φ(0))d¯n(¯n,y¯n)+un,

whenever n is large enough (so that D¯ϵn-1), and where un=ϵn+dX¯×Rk(Φ(0),knΦn(0))0. Now, let tRk such that ϕ(t)=0, and observe that:

d¯n(¯n,y¯n)dX¯n×Rk((¯n,0),(y¯n,t))dX¯×Rk((¯,0),Φ(t))+vn=d¯(¯,y¯)+vn,

when n is large enough (so that (D¯2+|t|2)1/2ϵn-1), and where vn:=ϵn+dX¯×Rk(knΦn(t),Φ(t))0. Hence, combining the two inequalities above, we obtain:

|dX¯×Rk((¯,0),Φ(0))-d¯(¯,y¯)|un+vn0.

In particular, d¯2(¯,y¯)=d¯2(¯,y¯)+|ϕ(0)|2. In conclusion, ϕ(0)=0.

Lemma 3.2

Let {y¯n} and {z¯n} be sequences such that, for every nN, y¯n,z¯nX¯n. For every nN and tRk, we define Φn(t):=(y¯n,t) and Ψn(t):=(z¯n,t). Assume that (passing to a subsequence if necessary), the sequences of maps {knΦn} and {knΨn} converge locally uniformly, respectively to Φ=(y¯,ϕ) and Ψ=(z¯,ψ), where y¯,z¯X¯ and ϕ,ψOk(R). Then, we necessarily have ϕ=ψ.

Proof

Looking for a contradiction, let us suppose that ϕψ. In that case, there exists tRk\{0} such that ϕ(t)ψ(t), which implies limsdX¯×Rk(Φ(st),Ψ(st))=. In particular, there exists sR such that dX¯×Rk(Φ(st),Ψ(st))D¯+1. However, using d¯n(y¯n,z¯n)=dX¯n×Rk(Φn(st),Ψn(st)), we have:

dX¯×Rk(Φ(st),Ψ(st))d¯n(y¯n,z¯n)+unD¯+un, 12

where un=ϵn+dX¯×Rk(Φ(st),knΦn(st))+dX¯×Rk(Ψ(st),knΨn(st)), and when n is large enough (so that D¯ϵn-1). Now, observe that limnun=0; therefore, passing to the limit in inequality 12, we have dX¯×Rk(Φ(st),Ψ(st))D¯, which contradicts dX¯×Rk(Φ(st),Ψ(st))D¯+1.

We can now state the convergence result on the splitting quantities.

Lemma 3.3

For every R>0, we have limnα(n,R)=limnβ(n,R)=0.

Proof

Part I:limnα(n,R)=0

Looking for a contradiction, we assume that limnα(n,R)0. Passing to a subsequence if necessary, there exist ϵ>0, and sequences {y¯n} and {tn} such that:

ϵdX¯×Rk(kn(y¯n,tn),(knS(y¯n),knR(tn))), 13

y¯nX¯n, and |tn|R. Moreover, since {tn} is bounded, we can assume that tnt. Now, applying Lemma 3.1, and passing to a subsequence if necessary, we can assume that {knΦn} converges locally uniformly to Φ=(y¯,ϕ), where Φn(s):=(y¯n,s), ϕOk(R) and y¯X¯. In particular, we have:

limnd¯(pX¯kn(y¯n,tn),knS(y¯n))=d¯(pX¯Φ(t),pX¯Φ(0))=0. 14

Now, applying Lemma 3.1 and Lemma 3.2, we can also assume that {knΨn} converges locally uniformly to Ψ=(z¯,ϕ), where Ψn(s):=(¯n,s), and z¯X¯. Thus, we have:

limndeucli(pRkkn(y¯n,tn),knR(tn))=deucli(pRkΦ(t),pRkΨ(t))=0. 15

Hence, using equations 14 and 15, we have limndX¯×Rk(kn(y¯n,tn),(knS(y¯n),knR(tn)))=0, which contradicts inequality 13.

Part II:limnβ(n,R)=0

Looking for a contradiction, we assume that limnβ(n,R)0. Passing to a subsequence if necessary, there exist ϵ>0, and sequences {y¯(n)}X¯N and |tn|R such that:

ϵdX¯n×Rk(ln(y¯(n),tn),(lnS(y¯(n)),lnR(tn))). 16

Observe that {tn} is bounded and X¯ is compact; therefore, we can assume that tnt and y¯(n)y¯.

Now, let us define (y¯n,sn):=ln(y¯(n),tn), and Φn(u):=(y¯n,u), uRk. Applying Lemma 3.1, we can assume (passing to a subsequence if necessary) that {knΦn} converges locally uniformly to (z¯,ϕ), where z¯X¯ and ϕOk(R). Observe that {sn} is bounded since {(y¯(n),tn)} converges. Therefore, we can assume that sns. However, we have limnknΦn(sn)=Φ(s)=limnknln(y¯(n),tn)=(y¯,t). Thus, z¯=y¯, and ϕ(s)=t. Now, observe that d¯n(pX¯nln(y¯(n),tn),lnS(y¯(n)))dX¯n×Rk(Φn(0),ln(Φ(0)))+dX¯n×Rk(ln(y¯,0),ln(y¯(n),0)). Therefore:

limnd¯n(pX¯nln(y¯(n),tn),lnS(y¯(n)))=0. 17

Then, applying Lemma 3.1 and Lemma 3.2, we can also assume that {knΨn} converges locally uniformly to Ψ=(z¯,ϕ), where Ψn(u):=(¯n,u), and z¯X¯. Moreover, Ψ(0)=limnknΨn(0)=(¯,0); therefore, z¯=¯. Now, using pRkln(y¯(n),tn)=sn, and lnR(tn)=pRkln(Ψ(ϕ-1(tn))), observe that deucli(pRkln(y¯(n),tn),lnR(tn))deucli(sn,ϕ-1(tn))+dX¯n×Rk(Ψn(ϕ-1(tn)),ln(Ψ(ϕ-1(tn)))). Therefore, using limnsn=s=ϕ-1(t)=limnϕ-1(tn), we obtain:

limndeucli(pRkln(y¯(n),tn),lnR(tn))=0. 18

Finally, observe that equations 17 and 18 contradict inequality 16, which concludes the proof.

The continuity of the soul map is a consequence of the following proposition, which gives us property 10 as a corollary.

Proposition 3.5

The sequence {knS,lnS} (resp. {knR,lnR}) realizes the convergence of {X¯n,d¯n,m¯n} (resp. {Rk,deucli,Lk,0}) to (X¯,d¯,m¯) (resp. (Rk,deucli,Lk,0)) in the mGH-topology (resp. pmGH-topology).

Proof

Part I:{knR,lnR}realizes the convergence of{Rk,deucli,Lk,0}to(Rk,deucli,Lk,0)

We are going to show that there exists a map ϵR:N×R0R0 such that for every R>0:

  • (i)

    limnϵR(n,R)=0,

  • (ii)

    for every |t|R, we have deucli(knRlnR(t),t)ϵR(n,R), and deucli(lnRknR(t),t)ϵR(n,R) (when n is large enough),

  • (iii)

    max{Dis(knR|BRk(0,R)),Dis(lnR|BRk(0,R))}ϵR(n,R) (when n is large enough).

Then we will prove that {knRLk} converges to Lk for the weak- topology.

Let R>0, and let tRk such that |t|R. Observe that dX¯×Rk(kn(¯n,t),(¯,knR(t)))α(n,R). Hence, thanks to Lemma 3.3, we get dX¯n×Rk(lnkn(¯n,t),ln(¯,knR(t)))α(n,R)+ϵnR (when n is large enough). In addition, we have dX¯×Rk((¯,0),kn(¯n,t))R+ϵn (when n is large enough). In particular, this implies |knR(t)|R+α(n,R)+ϵn2R. Therefore, we get dX¯n×Rk(ln(¯,knR(t)),(¯n,lnRknR(t)))β(n,2R). In conclusion, when n is large enough, we obtain deucli(t,lnRknR(t))2ϵn+α(n,R)+β(n,2R). The same strategy leads to deucli(t,knRlnR(t))2ϵn+β(n,R)+α(n,2R), for n large enough. Therefore, we get point (ii) if we set ϵR(n,R):=2(ϵn+α(n,2R)+β(n,2R)). Moreover, thanks to Lemma 3.3, we have limnϵR(n,R)=0, for every R>0.

Now, given t1,t2Rk such that |t1|R and |t2|R, we define:

A:=||deucli(knR(t1),knR(t2))-deucli(t1,t2)|-|dX¯×Rk(kn(¯n,t1),kn(¯n,t2))-deucli(t1,t2)||.

Using knS(¯n)=¯, we get:

A|deucli(knR(t1),knR(t2))-dX¯×Rk(kn(¯n,t1),kn(¯n,t2))|dX¯×Rk(kn(¯n,t1),(¯,knR(t1)))+dX¯×Rk(kn(¯n,t2),(¯,knR(t2)))2α(n,R).

Hence, for n large enough, we have:

|deucli(knR(t1),knR(t2))-deucli(t1,t2)|A+|dX¯×Rk(kn(¯n,t1),kn(¯n,t2))-deucli(t1,t2)|2α(n,R)+ϵn.

In particular, this implies Dis(knR|BRk(0,R))2α(n,R)+ϵn. Moreover, kn and ln playing symmetric roles, we also have Dis(lnR|BRk(0,R))2β(n,R)+ϵn. We replace ϵR(n,R) by ϵR(n,R)+2α(n,R)+2β(n,R)+ϵn. This concludes the proof of (i), (ii), and (iii) (thanks to Lemma 3.3).

Let us prove that {knRLk} converges to Lk for the weak- topology. Here, the strategy will be the same as in the proof of Theorem A. More precisely, the weak- topology on Mloc(Rk) is metrizable; therefore, it is equivalent to prove that every subsequence of {knRLk} admits a subsequence converging to Lk in the weak- topology. Let us just prove that {knRLk} admits a subsequence converging to Lk in the weak- topology (the proof for a subsequence being exactly the same). First of all, observe that thanks to Lemma 3.1, we can assume (passing to a subsequence if necessary) that {tRkkn(¯n,t)X¯×Rk} converges locally uniformly to (y¯,ϕ) for some y¯X¯ and ϕOk(R). In particular, {knR} converges locally uniformly to ϕ. Then, notice that y¯=limnpX¯kn(¯n,0)=¯. Now let R>0, and let fCc(Rk) be a continuous function such that Spt(f)BRk(0,R), and let us show that limnRkfdknRLk=RkfdLk.

First, observe that if fknR(t)0, then |knR(t)|R. Therefore, deucli(knR(t),knR(0))R+deucli(knR(0),0). In particular, we have:

dX¯×Rk(kn(¯n,t),kn(¯n,0))R~n,

where R~n:=(D¯2+(R+deucli(knR(0),0))2)1/2, and D¯ is defined in Notation 3.1. Hence, if fknR(t)0, then we have |t|=dX¯n×Rk((¯n,t),(¯n,0))3ϵn+R~n2R~:=2(D¯2+R2)1/2 (then n is sufficiently large). Hence, whenever n is large enough, we have:

|RkfdknRLk-RkfdϕLk|B(2R~)|fknR(t)-fϕ(t)|dLk(t)Lk(B(2R~))ωn,

where ωn:=sup|x-y|νn{|f(x)-f(y)|} and νn:=suptB(2R~){|knR(t)-ϕ(t)|}. Observe that knR converges locally uniformly to ϕ; thus νn0. In particular, since f has compact support, ωn0; hence limnRkfdknRLk=RkfdϕLk. In conclusion, passing to a subsequence if necessary, {knRLk} converges in the weak- topology to ϕLk=Lk (using ϕOk(R)).

Part II:{knS,lnS}realizes the convergence of{X¯n,d¯n,m¯n}to(X¯,d¯,m¯)

Let y¯nX¯n, and observe that dX¯×Rk(kn(y¯n,0),(knS(y¯n),0))α(n,0). Thanks to Lemma 3.3, we have dX¯n×Rk(lnkn(y¯n,0),ln(knS(y¯n),0))α(n,0)+ϵn (when n is large enough). However, we have dX¯n×Rk(ln(knS(y¯n),0),(lnSknS(y¯n),0))β(n,0). In conclusion, we have:

d¯n(y¯n,lnSknS(y¯n))=dX¯n×Rk((y¯n,0),(lnSknS(y¯n),0))α(n,0)+β(n,0)+2ϵn=:ϵnS. 19

Since kn and ln play symmetric roles, we also have:

d¯(y¯,knSlnS(y¯))ϵnS, 20

for every y¯X¯. Observe that, thanks to Lemma 3.3, we have limnϵnS=0.

Now, let y1,y2X¯n, and define:

A:=||d¯(knS(y1),knS(y2))-d¯n(y1,y2)|-|dX¯×Rk(kn(y1,0),kn(y2,0))-d¯n(y1,y2)||.

Then, we have:

A|d¯(knS(y1),knS(y2))-dX¯×Rk(kn(y1,0),kn(y2,0))|dX¯×Rk(kn(y1,0),(knS(y1),0))+dX¯×Rk(kn(y2,0),(knS(y2),0))2α(n,0).

Hence, we finally get (for n large enough):

|d¯(knS(y1),knS(y2))-d¯n(y1,y2)|A+|dX¯×Rk(kn(y1,0),kn(y2,0))-d¯n(y1,y2)||2α(n,0)+ϵn.

In particular, we have Dis(knS)2α(n,0)+ϵn. Then, kn and ln playing symmetric roles, we also have Dis(lnS)2β(n,0)+ϵn. Finally, replacing ϵnS by ϵnS+ϵn+2max{α(n,0),β(n,0)}, and applying Lemma 3.3, we have ϵnS0; therefore, using inequalities 19 and 20, we can conclude that {X¯n,d¯n} converges to (X¯,d¯) in the GH-topology.

Now let us prove that {knSm¯n} converges to m¯ in the weak- topology. As we’ve seen in Part I, it is sufficient to show that {knSm¯n} admits a subsequence converging to m¯ in the weak- topology (the weak- topology being metrizable).

First, let us show that {knSm¯n} is precompact in the space M(X¯) of Radon measure on X¯, which is implied by the uniform boundedness of the sequence {m¯n(X¯n)}. Let us fix r0(0,δ/2), where δ:=infnN{}{δ(X,dn)} (δ being positive thanks to Proposition 3.2). Then, using point (v) of Proposition 2.2 and Theorem 2.1, observe that m¯nLk(BX¯n×Rk((¯n,0),r0))=mn(Bdn(n,r0))M, where M:=supnN{}{mn(X)} is finite. Moreover, notice that Bd¯n(¯n,r0/2)×BRk(0,r0/2)BX¯n×Rk((¯n,0),r0). Hence, m¯n(Bd¯n(¯n,r0/2))(2/r0)kM/ωk, where ωk:=Lk(BRk(0,1)). In particular, for every nN, we can apply Bishop–Gromov inequality (see Theorem 6.2 in [6]), and get m¯n(X¯n)(2/r0)ND¯N-kM/ωk=:M¯, where D¯ is defined in Notation 3.1. In conclusion, {m¯n(X¯n)} is uniformly bounded; thus {knSm¯n} is precompact in the weak- topology.

Now, passing to a subsequence if necessary, we can assume that {knSm¯n} converges to some Radon measure m¯ on X¯. We need to prove that m¯=m¯. Observe that it is equivalent to prove that m¯Lk=m¯Lk. First, thanks to the first part of the proof, {knRLk} converges to Lk in the weak- topology; hence {(knS,knR)[m¯nLk]} converges to m¯Lk in the weak- topology. In addition, thanks to Theorem A, {kn[m¯nLk]} converges to m¯Lk. Now, let ϕCc(X¯×Rk) and R>0 such that Spt(ϕ)BX¯×Rk(R). Then, proceeding as in Part I of the proof, we obtain:

Spt(ϕ(knS,knR))Spt(ϕkn)X¯×BRk(0,2R~),

when n is sufficiently large and where R~:=(D¯2+R2)1/2. In particular, we have:

|X¯n×Rkϕ(knS(x¯),knR(t))-ϕ(kn(x¯,t))dm¯nLk(x¯,t)|(2R~)kωkM¯ωϕ(α(n,2R~)),

where ωϕ is the modulus of uniform continuity associated to ϕ. Then, thanks to Lemma 3.3, we have limnωϕ(α(n,2R~))=0. Thus, for every ϕCc(X¯×Rk), we have:

limnX¯n×Rkϕd(knS,knR)[m¯nLk]=limnX¯n×Rkϕdkn[m¯nLk].

In particular, {(knS,knR)[m¯nLk]} and {kn[m¯nLk]} have the same limit, i.e. m¯Lk=m¯Lk. This concludes the proof.

Inspired by the proof of Theorem 5.4 in [37], we introduce the following "shrunk" metrics.

Definition 3.2

Given nN{}, and mN, let d~n,m:=ϕn(2-md¯n×deucli), and let (X,dn,m,mn) be the push-forward of (X~,d~n,m,m~n) (see Proposition 2.3).

Remark 3.1

Given nN{}, mN, and x~,y~X~, we have d~n,m(x~,y~)d~n(x~,y~). Indeed, defining (x¯,tx):=ϕn(x) and (y¯,ty):=ϕn(y), we have:

d~n,m(x~,y~)2=2-2md¯n2(x¯,y¯)+deucli2(tx,ty)d¯n2(x¯,y¯)+deucli2(tx,ty)=d~n(x~,y~)2.

In particular, this implies dn,mdn.

The following lemma shows that the “shrunk” metrics associated to the sequence {X,dn,mn} are close to the corresponding Albanese varieties.

Lemma 3.4

We have dGH([X,dn,m],A([X,dn,mn]))2-m+1D¯, for every mN, and nN{} (where D¯ is defined in Notation 3.1).

Proof

First of all, observe that there exists a continuous map a:X(Rk/Γ(ϕn),dΓ(ϕn)) such that ap=qpRkϕn, where q:RkRk/Γ(ϕn) is the quotient map. Notice that, qpRkϕn is surjective; hence, a is also surjective. Now, let y,zX and let us show that |dn,m(y,z)-dΓ(ϕn)(a(y),a(z))|2-mD¯.

First, thanks to Proposition 2.3, there exists y~p-1(y) and z~p-1(z) such that dn,m(y,z)=d~n,m(y~,z~). Let (y¯,ty):=ϕn(y~) and (z¯,tz):=ϕn(z~). Observe that a(y)=q(ty) and a(z)=q(tz), and dn,m2(y,z)=2-2md¯n2(y¯,z¯)+deucli2(ty,tz). Now, note that, by definition of dΓ(ϕn), we have dΓ(ϕn)(a(y),a(z))deucli(ty,tz); in particular:

0dn,m(y,z)-dΓ(ϕn)(a(y),a(z)). 21

Notice that, by definition of dΓ(ϕn), there exists ηπ¯1(X) such that dΓ(ϕn)(a(y),a(z))=deucli(ty,t2), where t2:=ρRϕn(η)·tz. Then, observe that d~n,m(y~,z~)=dn,m(y,z)d~n,m(y~,η·z~); hence:

dn,m(y,z)-dΓ(ϕn)(a(y),a(z))(2-2md¯n2(y¯,ρSϕn(η)·z¯)+deucli(ty,t2))1/2-deucli(ty,t2)2-md¯n(y¯,ρSϕn(η)·z¯)+deucli(ty,t2)-deucli(ty,t2)2-mD¯.

In particular, thanks to inequality 21, we have |dn,m(y,z)-dΓ(ϕn)(a(y),a(z))|2-mD¯. Therefore, recalling that a is surjective, and using Corollary 7.3.28 of [11], we get dGH([X,dn,m],A([X,dn,mn]))2-m+1D¯.

To prove property 11, we will need to obtain a convergence result on the following quantities.

Notation 3.3

Given R>0, n,mN, we denote:

  • (i)

    ϵ(n,m,R):=sup{|d~,m(f~n(y~1),f~n(y~2))-d~n,m(y~1,y~2)|}, the supremum being taken over y~iB~n(R),

  • (ii)

    ϵ(n,m,R):=sup{|d~n,m(g~n(y~1),g~n(y~2))-d~,m(y~1,y~2)|}, the supremum being taken over y~iB~(R).

Lemma 3.5

For every R>0, limn,mϵ(n,m,R)=limn,mϵ(n,m,R)=0.

Proof

Let us only prove that limn,mϵ(n,m,R)=0, the proof for ϵ(n,m,R) being exactly the same. Let y~iB~n(R), i{1,2}. For i{1,2}, we denote (y¯i,ti):=ϕn(y~i), y~i:=f~n(y~i), (y¯i,ti):=ϕ(y~i), and A:=|d~,m(f~n(y~1),f~n(y~2))-d~n,m(y~1,y~2)|. Using the fact that, for every x,yR0, we have |x-y||x-y|, we get:

A(|2-2m(d¯2(y¯1,,y¯2)-d¯n2(y¯1,,y¯2))+(deucli2(t1,t2)-deucli2(t1,t2))|)1/22-m(|d¯2(y¯1,,y¯2)-d¯n2(y¯1,,y¯2)|)1/2+(|deucli2(t1,t2)-deucli2(t1,t2)|)1/2.

However, note that |(d¯2(y¯1,y¯2)-d¯n2(y¯1,y¯2))|2D¯2, where D¯ is introduced in Notation 3.1. Then, observe that deucli(t1,t2)d~(y~1,y~2)d~n(y~1,y~2)+ϵn2R+ϵn, when n is large enough. Therefore, deucli(t1,t2)+deucli(t1,t2)4R+ϵn. Then, using |ti|R, and denoting B:=|deucli(t1,t2)-deucli(t1,t2)|, we have:

B|deucli(t1,t2)-deucli(knR(t1),knR(t2))|+|deucli(knR(t1),knR(t2))-deucli(t1,t2)|Dis(kn|BRk(0,R)R)+deucli(knR(t1),t1)+deucli(knR(t2),t2)Dis(kn|BRk(0,R)R)+2α(n,R).

In conclusion, we obtain:

A2-m+1/2D¯+(4R+ϵn)1/2(Dis(kn|BRk(0,R)R)+2α(n,R))1/2=:ϵ~(n,m,R).

Therefore, passing to the supremum as y~iB~n(R) (i{1,2}), we obtain ϵ(n,m,R)ϵ~(n,m,R). Thanks to Lemma 3.3 and Proposition 3.5, we have limn,mϵ~(n,m,R)=0.

Therefore, limn,mϵ(n,m,R)=0, which concludes the proof.

We conclude this section with the following proposition, which states the continuity of the Albanese map by proving property 11.

Proposition 3.6

The sequence {[Rk/Γ(ϕn),dΓ(ϕn)]=A(X,dn,mn)} converges in the GH-topology to [Rk/Γ(ϕ),dΓ(ϕ)]=A(X,d,m).

Proof

Observe that for every nN{}, we have dGH([X,dn,n],A(X,dn,mn))2-n+1D¯, thanks to Lemma 3.4 (where dn,n is defined in Definition 3.2 and D¯ is introduced in Notation 3.1). In particular, using the triangle inequality for the Gromov–Hausdorff distance dGH, we obtain:

dGH(A(X,dn,mn),A(X,d,m))2-n+2D¯+dGH([X,dn,n],[X,d,n]).

Therefore, to conclude, it is sufficient to prove that:

limn,mdGH([X,dn,m],[X,d,m])=0,

which is what we are going to prove.

Let n,mN, and y1,y2X. There exists y~1p-1(y1), and y~2p-1(y2), such that d~n(~n,y~1)=dn(n,y1), and d~n,m(y~1,y~2)=dn,m(y1,y2). Then, for i{1,2}, we denote (y¯i,ti):=ϕn(y~i). Observe that d~n(y~1,y~2)=(d¯n2(y¯1,y¯2)+deucli2(t1,t2))1/2, where d¯n(y¯1,y¯2)D¯, and deucli(t1,t2)dn,m(y1,y2)dn(y1,y2)D (using Remark 3.1). Therefore, we get:

d~n(~n,y~2)D+(D¯2+D2)1/2=:D~. 22

Now, using dn,m(y1,y2)=d~n,m(y~1,y~2), and f~n(y~i)p-1(fn(yi)), we have:

d,m(fn(y1),fn(y2))-dn,m(y1,y2)d~,m(f~n(y~1),f~n(y~2))-d~n,m(y~1,y~2)ϵ(n,m,D~),

where ϵ(n,m,R) is introduced in Notation 3.3. Since f~n and g~n play symmetric roles, we also have:

y1,y2X,dn,m(gn(y1),gn(y2))-d,m(y1,y2)ϵ(n,m,D~),

where ϵ(n,m,D~) is also introduced in Notation 3.3. In particular, this implies:

y1,y2X,dn,m(gnfn(y1),gnfn(y2))-d,m(fn(y1),fn(y2))ϵ(n,m,D~).

Hence, defining A:=dn,m(y1,y2)-d,m(fn(y1),fn(y2)), we have:

Adn,m(y1,y2)-dn,m(gnfn(y1),gnfn(y2))+ϵ(n,m,D~)dn,m(gnfn(y1),y1)+dn,m(gnfn(y2),y2)+ϵ(n,m,D~)dn(gnfn(y1),y1)+dn(gnfn(y2),y2)+ϵ(n,m,D~)2ϵn+ϵ(n,m,D~).

In conclusion, we have:

|d,m(fn(y1),fn(y2))-dn,m(y1,y2)|2ϵn+ϵ(n,m,D~)+ϵ(n,m,D~). 23

Moreover, since d,md, and since fn is an ϵn-isometry from (X,dn) onto (X,d), we have:

xX,yX,d,m(x,fn(y))ϵn. 24

Hence, thanks to inequalities 23 and 24, fn is a 2ϵn+ϵ(n,m,D~)+ϵ(n,m,D~)-isometry from (X,dn,m) to (X,d,m). Therefore, using Corollary 7.3.28 of [11], we have:

dGH([X,dn,m],[X,d,m])2(2ϵn+ϵ(n,m,D~)+ϵ(n,m,D~)).

However, thanks to Lemma 3.5, we have limn,m2ϵn+ϵ(n,m,D~)+ϵ(n,m,D~)=0, which concludes the proof.

Proof of Theorem C

The proof of Theorem C is inspired by the proof of Theorem 1.1 in [37] and uses some of the computations realized in [29].

First of all, using Theorem B, we are going to prove the following result.

Proposition 3.7

Let N[1,), let X be a compact topological space that admits an RCD(0,N)-structure such that π¯1(X)=0 (see Theorem 1.2 for the definition of π¯1(X)), and let Γ be a Bieberbach subgroup of Rk (k2). Then, the moduli space M0,N+k(X×Rk/Γ) retracts onto Mflat(Rk/Γ).

Proof

Let us describe the crystallographic class Γ(X×Rk/Γ) (introduced in Proposition 2.7). First, observe that, since π¯1(X)=0, then the universal cover of X×Rk/Γ is X×Rk, and the covering map is just idX×q, where q:RkRk/Γ is the usual quotient map. Now, let g be the flat Riemannian metric on Rk/Γ such that q is a local isometry, and fix an RCD(0,N)-structure (X,d0,m0) on X. Observe that (X,d0,m0)×(Rk/Γ,dg,mg) is an RCD(0,N+k)-structure on X×Rk/Γ, where dg and mg are respectively the Riemannian distance and measure associated to g. Moreover, the lifted RCD(0,N+k)-structure on X×Rk is equal to (X,d0,m0)×(Rk,deucli,Lk). In particular, the identity map idX×Rk is a splitting of (X,d0,m0)×Rk. Moreover, since π¯1(X×Rk/Γ) acts trivially on X, we have Γ(idX×Rk)=Γ. Hence, the crystallographic class Γ(X×Rk/Γ) is equal to the set of crystallographic subgroups of Iso(Rk) that are isomorphic to Γ. This implies, thanks to Remark 2.5, that Mflat(A(X×Rk/Γ)) is isometric to Mflat(Rk/Γ).

Now, thanks to Theorem B, the Albanese map associated to X×Rk/Γ is continuous from M0,N+k(X×Rk/Γ) onto Mflat(A(X×Rk/Γ)). Hence, it gives rise to a continuous surjective map ϕ from M0,N+k(X×Rk/Γ) onto Mflat(Rk/Γ). Given [Rk/Γ,d]Mflat(Rk/Γ), we define:

s([Rk/Γ,d]):=[(X,d0,m0)×(Rk/Γ,d,Hd)]M0,N+k(X×Rk/Γ),

where Hd is the Hausdorff measure associated to (Rk/Γ,d). Observe that s is a section of ϕ; therefore, we only have to show that s is continuous in order to conclude the proof.

Let us show that s:Mflat(Rk/Γ)M0,N+k(X×Rk/Γ) is continuous. To do so, let {(Rk/Γ,dn)} converge in the Gromov–Hausdorff sense to (Rk/Γ,d), where, for every nN{}, dn is a flat metric on Rk/Γ. Let us prove that {(X,d0,m0)×(Rk/Γ,dn,Hdn)} converges in the measured Gromov–Hausdorff sense to (X,d0,m0)×(Rk/Γ,d,Hd). Observe that it is sufficient to prove that {(Rk/Γ,dn,Hdn)} converges in the mGH sense to (Rk/Γ,d,Hd). However, since d is a flat metric on Rk/Γ, the Hausdorff dimension of (Rk/Γ,d) is equal to k. In particular, by Theorem 1.2 of [18], {(Rk/Γ,dn,Hdn)} converges in the mGH sense to (Rk/Γ,d,Hd). In conclusion s is continuous.

Proposition 3.7 implies that the homotopy groups of Mflat(Rk/Γ) inject in those of M0,N+k(X×Rk/Γ). Therefore, the topology of M0,N+k(X×Rk/Γ) is, in a way, at least as complicated as the topology of Mflat(Rk/Γ). Thankfully, informations on moduli spaces of flat metrics have been derived in [37] (in the case of the torus Tk, with k4 and k8,9,10) and in [29] (in the case of 3 and 4-dimensional closed flat Riemannian manifolds). We are now able to prove Theorem C.

Proof of Theorem C

Observe that thanks to Theorem 3.4.3 of [29] and Proposition 5.5 of [37] the moduli space Mflat(N) has non-trivial higher rational homotopy groups. Therefore, Proposition 3.7 concludes the proof.

Let us now prove Corollary B.

Proof

First of all, observe that thanks to Theorem 3.4.3 of [29], the moduli space of flat metrics on X3:=S1×K2 is homotopy equivalent to a circle (where K2 is the Klein bottle). Then, let us define X4:=[0,1]×X3, and XN:=SN-3×X3 (N5). Thanks to Proposition 3.7, for every N3, M0,N(XN) retracts onto Mflat(X3). In particular, for every N3, M0,N(XN) has non trivial fundamental group.

To conclude the proof, we apply the same idea, using the fact that π3(T4)QQ, and π5(T5)QQ (see Proposition 5.5 of [37]).

Acknowledgements

Both the authors are supported by the European Research Council (ERC), under the European Union Horizon 2020 research and innovation programme, via the ERC Starting Grant CURVATURE, Grant agreement No. 802689. The authors wish to thank Gérard Besson for stimulating conversations on the topics of the paper.

Appendix

Before proving Proposition 2.8, let us point out the following technicality on the Prokhorov distance.

Remark a

There are various notions of distances between restricted measures. Indeed, given a pointed complete separable metric space (Y,d,) endowed with two boundedly finite measures m1 and m2, and R>0, we can define: graphic file with name 208_2022_2493_Figd_HTML.jpg or: graphic file with name 208_2022_2493_Fige_HTML.jpg where m1R (resp. m2R) is the restriction of m1 (resp. m2) to B¯R. Let us see how these two notions differ.

First of all, we easily obtain:

dPR(m1,m2)dP(m1R,m2R). 25

Then, if dPR(m1,m2)ϵ, we have:

dP(m1R,m2R)ϵ+m1(B¯R+ϵ\B¯R)+m2(B¯R+ϵ\B¯R). 26

Therefore, both notions lead to the same convergence. Indeed, given a boundedly finite measure m and a sequence {mk} of boundedly finite measures, we have the following equivalence thanks to inequalities 25 and 26:

mkmR>0,dP(mkR,mR)0R>0,dPR(mk,m)0.

In Definition 2.8, we are using dPR(m1,m2) instead of dP(m1R,m2R) because the first one is a non-decreasing function of R whereas the other one is not. Doing so, we are losing the triangle inequality. In fact, given three boundedly finite measures mi (i{1,2,3}) such that dPR(m1,m2)δ and dPR(m2,m3)η, we only have:

dPR-(δ+η)(m1,m3)δ+η. 27

This will be sufficient to investigate the properties of Dpeq.

Proof

Part I: Properties ofDpeq

Observe that Dpeq is symmetric, nonnegative and invariant under equivariant isomorphisms. However, Dpeq does not satisfy the triangle inequality a priori. Nevertheless, it will be sufficient for our purposes to show that Dpeq satisfy the following modified triangle inequality:

Dpeq(X~1,X~3)4(Dpeq(X~1,X~2)+Dpeq(X~2,X~3)) 28

for any three equivariant pointed RCD(0,N)-structures X~i=(X~,d~i,m~i,~i)R0,Np,eq(X~) (i{1,2,3}).

Observe that the inequality is trivially true whenever Dpeq(X~1,X~2) or Dpeq(X~2,X~3) is equal to 1/24.

Now, assume that for i{1,2}, we have Dpeq(X~i,X~i+1)ϵi,i+1, where ϵi,i+1(0,1/24), and let (fi,i+1,gi,i+1,ϕi,i+1) be an associated equivariant pointed ϵi,i+1-isometry. We define f:=f23f12, g:=g12g23, and ϕ:=ϕ23ϕ12. We want to show that (f,g,ϕ) is an equivariant pointed 4(ϵ12+ϵ23)-isometry between X~1 and X~3.

First, point (i) and point (ii) of Definition 2.8 are trivially satisfied.

Then, let x,yX~ such that d1(x,y)[4(ϵ12+ϵ23)]-1. Notice that [4(ϵ12+ϵ23)]-1ϵ12-1. Therefore d~2(f12(x),f12(y))ϵ12+[4(ϵ12+ϵ23)]-1ϵ23-1 (where we used ϵi,i+1<1/24). Hence, we have |d~3(f(x),f(y))-d~2(f12(x),f12(y))|ϵ23 and |d~2(f12(x),f12(y))-d~1(x,y)|ϵ12. Thus, we get:

|d~3(f(x),f(y))-d~1(x,y)|ϵ12+ϵ234(ϵ12+ϵ23).

Using the same argument for g, we can conclude that point (iii) of Definition 2.8 is satisfied.

Now, let xX~ and observe that d~2(g23f23f12(x),f12(x))ϵ23ϵ12-1. Hence, we have:

d~1(gf(x),x)d~1(x,g12f12(x))+d~1(g12g23f23f12(x),g12f12(x))ϵ12+ϵ23+ϵ124(ϵ12+ϵ23).

Arguing the same way for fg, we can conclude that point (iv) of Definition 2.8 is satisfied.

Now, let us show that point (v) of Definition 2.8 is also satisfied. First of all, let us found an upper bound on dP[2(ϵ12+ϵ23)]-1(fm~1,f23m~2). Let A be a closed subset of B¯3([2(ϵ12+ϵ23)]-1). We easily get that f23-1(A)B¯2(ϵ12-1), which implies:

fm~1(A)=f23(f12m~1)(A)m~2((f23-1(A))ϵ12)+ϵ12.

Then, note that we have (f23-1(A))ϵ12f23-1((A)ϵ12+ϵ23). Therefore, we have fm~1(A)f23m~2(Aϵ12+ϵ23)+ϵ12. Doing the same in the opposite direction gives us:

dP[2(ϵ12+ϵ23)]-1(fm~1,f23m~2)ϵ12+ϵ23.

Finally, observe that [4(ϵ12+ϵ23)]-1[2(ϵ12+ϵ23)]-1-(ϵ12+2ϵ23). Thus, applying inequality 27 of Remark a, we have:

dP[4(ϵ12+ϵ23)]-1(fm~1,m~3)dP[2(ϵ12+ϵ23)]-1-(ϵ12+2ϵ23)(fm~1,m~3)dP[2(ϵ12+ϵ23)]-1(f23m~2,m~3)+dP[2(ϵ12+ϵ23)]-1(f23m~2,fm~1)ϵ12+2ϵ234(ϵ12+ϵ23).

Applying the same argument, we also get dP[4(ϵ12+ϵ23)]-1(gm~3,m~1)4(ϵ12+ϵ23). Therefore, point (v) of Definition 2.8 is satisfied.

This concludes the proof of the modified triangle inequality 28.

Part II: Hausdorff uniform structure

Let B:={{Dpeq2-n},nN}P(M0,Np,eq(X~)×M0,Np,eq(X~)). Thanks to the fact that Dpeq is well defined on M0,Np,eq(X~), symmetric, nonnegative, and satisfies the modified triangle inequality 28, we can easily check the axioms introduced p.141 of [9]. Therefore, B is a fundamental system of neighborhood of a uniform structure on M0,Np,eq(X~).

Let us now show that the uniform structure is Hausdorff. Let X~i=(X~,d~i,m~i,~i)R0,Np,eq(X~) (i{1,2}) such that Dpeq(X~1,X~2)=0. We need to prove that X~1 and X~2 are equivariantly isomorphic.

First, since Dpeq(X~1,X~2)=0, there is a sequence of equivariant pointed ϵn-isometries (fn,gn,ϕn) such that ϵn0. Let us fix a countable dense subset D1 in (X~,d~1) and observe that for every xD1, the sequence {fn(x)} is bounded. Therefore, applying Cantor’s diagonal argument, we can assume that there is a map f:D1(X~,d~2) such that, for every xD1, we have fn(x)f(x). Then, observe that given x,yD1, and n large enough, we have |d~2(fn(x),fn(y))-d~1(x,y)|ϵn0. Therefore, f is an isometric embedding, hence, can be extended in a unique way into an isometric embedding f:(X~,d~1)(X~,d~2). Then, it is not hard to prove that, given any sequence {xn} in X~ converging to xX~, we have fn(xn)f(x). Now, we can apply the same procedure to the sequence {gn}, and get an isometric embedding g:(X~,d~2)(X~,d~1) such that, given any sequence {xn} in X~ converging to xX~, we have gn(xn)g(x). In particular, given xX~, we have limnd~1(gn(fn(x)),x)=0=d~1(g(f(x)),x). Therefore, f and g are respectively inverse to each other. Moreover, we have f(~1)=limnfn(~1)=~2. The same argument gives us g(~2)=~1. Hence, f:(X~,d~1,~1)(X~,d~2,~2) is an isomorphism of pointed metric space. Now, observe that dPϵn-1(fnm~1,m~2)ϵn0. Hence, thanks to Remark a, {fnm~1} converges to m~2 in the weak- topology. Let us show that fm~1=m~2. We fix R>0 and hC0(X~) such that Spt(h)B¯2(R). Observe that we have X~hdm~2=limnX~hdfnm~1. However, hfn is point-wise converging to hf. Also, whenever n is large enough, we have Spt(hfn)B¯1(2R). Hence, applying the dominated convergence theorem, we obtain:

X~hdm~2=limnX~hdfnm~1=X~hdfm~1.

Therefore, since m~2 and fm~1 are Radon measures, we necessarily have m~2=fm~1. Finally, given γπ¯1(X) and xX~, we have:

p(f(x))=limnp(fn(x))=limnp(fn(γx))=p(f(γx)).

Thus, p(fγf-1)=p. We define ϕIso(π¯1(X)) by ϕ(γ):=fγf-1, which satisfies f(γx)=ϕ(γ)f(x).

Hence, we can conclude that X~1 is equivariantly isomorphic to X~2.

Part III: Metrizable uniform structure

We have seen that Dpeq induces a Hausdorff uniform structure on M0,Np,eq(X~). Moreover, B is a countable system of fundamental neighborhoods for this uniform structure. Therefore, thanks to Proposition 2 p.126 of [10], there exists a distance d:M0,Np,eq(X~)×M0,Np,eq(X~)[0,+] such that d induces the same uniform structure as Dpeq. Observe that we can assume, without loss of generality, that d is finite (replacing d by min{1,d} if necessary), which concludes the proof.

Declarations

Conflict of interest

On behalf of all authors, the corresponding author states that there is no conflict of interest.

Footnotes

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Contributor Information

Andrea Mondino, Email: Andrea.Mondino@maths.ox.ac.uk.

Dimitri Navarro, Email: navarro@maths.ox.ac.uk.

References

  • 1.Abraham R, Delma J-F, Hoscheit P. A note on the Gromov–Hausdorff–Prokhorov distance between (locally) compact metric measure spaces. Electron. J. Probab. 2013;18(14):1–21. [Google Scholar]
  • 2.Ambrosio L, Gigli N, Mondino A, Rajala T. Riemannian Ricci curvature lower bounds in metric measure spaces with -finite measure. Trans. Am. Math. Soc. 2015;367:4661–4701. doi: 10.1090/S0002-9947-2015-06111-X. [DOI] [Google Scholar]
  • 3.Ambrosio L, Gigli N, Savaré G. Metric measure spaces with Riemannian Ricci curvature bounded from below. Duke Math. J. 2014;163(7):1405–1490. doi: 10.1215/00127094-2681605. [DOI] [Google Scholar]
  • 4.Ambrosio L, Gigli N, Savaré G. Bakry-émery curvature-dimension condition and riemannian Ricci curvature bounds. Ann. Probab. 2015;43(1):339–404. doi: 10.1214/14-AOP907. [DOI] [Google Scholar]
  • 5.Ambrosio L, Mondino A, Savaré G. Nonlinear diffusion equations and curvature conditions in metric measure spaces. Mem. Am. Math. Soc. 2019;262(1270):v+121. [Google Scholar]
  • 6.Bacher K, Sturm K-T. Localization and tensorization properties of the curvature-dimension condition for metric measure spaces. J. Funct. Anal. 2010;259(1):28–56. doi: 10.1016/j.jfa.2010.03.024. [DOI] [Google Scholar]
  • 7.Belegradek I. Gromov–Hausdorff hyperspace of nonnegatively curved 2-spheres. Proc. Am. Math. Soc. 2018;146:1757–1764. doi: 10.1090/proc/13910. [DOI] [Google Scholar]
  • 8.Bettiol R, Derdzinski A, Piccione P. Teichmüller theory and collapse of flat manifolds. Ann. Math. Pura Appl. (1923 -) 2018;197(4):1247–1268. doi: 10.1007/s10231-017-0723-7. [DOI] [Google Scholar]
  • 9.Bourbaki N. Topologie générale: Chapitres 1 à 4. Berlin: Bourbaki, N. Springer; 2007. [Google Scholar]
  • 10.Bourbaki N. Topologie générale: Chapitres 5 à 10. Berlin: Bourbaki, N. Springer; 2007. [Google Scholar]
  • 11.Burago, D., Burago, Y., Ivanov, S.: A course in metric geometry. Graduate studies in mathematics. v. 33. American Mathematical Society, Providence, R.I. (2001)
  • 12.Cavalletti F, Milman E. The globalization theorem for the curvature dimension condition. Invent. Math. 2021;226(1):1–137. doi: 10.1007/s00222-021-01040-6. [DOI] [Google Scholar]
  • 13.Charlap LS. Bieberbach groups and flat manifolds. Universitext. New York: Springer-Verlag; 1986. [Google Scholar]
  • 14.Cheeger J, Colding T. On the structure of spaces with Ricci curvature bounded below. I. J. .ential Geometry. 1997;46(3):406–480. [Google Scholar]
  • 15.Cheeger J, Colding TH. On the structure of spaces with Ricci curvature bounded below. II. J. Differ. Geometry. 2000;54(1):13–35. [Google Scholar]
  • 16.Cheeger J, Colding TH. On the structure of spaces with Ricci curvature bounded below. III. J. Differ. Geometry. 2000;54(1):37–74. [Google Scholar]
  • 17.Daley D , Vere-Jones D. An Introduction to the Theory of Point Processes: Volume I: Elementary Theory and Methods Probability and Its Applications. New York: Springer-Verlag; 2003. [Google Scholar]
  • 18.De Philippis G, Gigli N. Non-collapsed spaces with Ricci curvature bounded from below. J. l’Ecole Polytechnique -Math. 2018;5:613–650. doi: 10.5802/jep.80. [DOI] [Google Scholar]
  • 19.Erbar M, Kuwada K, Sturm K-T. On the equivalence of the entropic curvature-dimension condition and Bochner’s inequality on metric measure spaces. Invent. Math. 2015;201:993–1071. doi: 10.1007/s00222-014-0563-7. [DOI] [Google Scholar]
  • 20.Fukaya K, Yamaguchi T. The fundamental groups of almost nonnegatively curved manifolds. Ann. Math. 1992;136(2):253–333. doi: 10.2307/2946606. [DOI] [Google Scholar]
  • 21.Gigli N. An overview of the proof of the splitting theorem in spaces with non-negative Ricci curvature. Anal. Geom. Metr. Spaces. 2014;2(1):169–213. doi: 10.2478/agms-2014-0006. [DOI] [Google Scholar]
  • 22.Gigli N. On the differential structure of metric measure spaces and applications. Mem. Am. Math. Soc. 2015;236:91. [Google Scholar]
  • 23.Gigli N, Mondino A, Savaré G. Convergence of pointed non-compact metric measure spaces and stability of Ricci curvature bounds and heat flows. Proc. Lond. Math. Soc. 2015;111(5):1071–1129. [Google Scholar]
  • 24.Gigli N, Rajala T, Sturm Optimal maps and exponentiation on finite-dimensional spaces with Ricci curvature bounded from below. J. Gometr. Anal. 2016;26:2914–2929. doi: 10.1007/s12220-015-9654-y. [DOI] [Google Scholar]
  • 25.Lott J, Villani C. Ricci curvature for metric-measure spaces via optimal transport. Ann. Math. 2009;169(3):903–991. doi: 10.4007/annals.2009.169.903. [DOI] [Google Scholar]
  • 26.Mondello I, Mondino A, Perales R. An upper bound on the revised first betti number and a torus stability result for RCD spaces. Comment. Math. Helv. 2022;97:555–609. doi: 10.4171/CMH/540. [DOI] [Google Scholar]
  • 27.Mondino A, Wei G. On the universal cover and the fundamental group of an RCD*(K, N)-space. J. für die reine und angewandte Mathematik (Crelles Journal) 2019;2019(753):211–237. doi: 10.1515/crelle-2016-0068. [DOI] [Google Scholar]
  • 28.Navarro, D.: Contractibility of moduli spaces of RCD(0,2)-structures. arXiv e-prints (Feb. 2022). arXiv:2202.06659
  • 29.Pérez, A., Tuschmann, W.: Spaces and Moduli Spaces of Flat Riemannian Metrics on Closed Manifolds. Karlsruher Institut für Technologie (KIT) (2020)
  • 30.Petersen P. Riemannian Geometry. Graduate Texts in Mathematics. New York: Springer; 2006. [Google Scholar]
  • 31.Reviron G. Rigidité topologique sous l’hypothèse “entropie majorée” et applications. Commentarii Mathematici Helvetici. 2008;83:815–846. doi: 10.4171/CMH/144. [DOI] [Google Scholar]
  • 32.Shen Z, Wei G. On Riemannian-manifolds of almost nonnegative curvature. Indiana Univ. Math. J. 1991;40(2):551–565. doi: 10.1512/iumj.1991.40.40026. [DOI] [Google Scholar]
  • 33.Sormani C, Wei G. Hausdorff convergence and universal covers. Trans. Am. Math. Soc. 2001;353(9):3585–3602. doi: 10.1090/S0002-9947-01-02802-1. [DOI] [Google Scholar]
  • 34.Spanier, E.H.: Algebraic topology. Springer-Verlag, New York, 1st corr. springer ed. edition (1981)
  • 35.Sturm K-T. On the geometry of metric measure spaces. Acta Math. 2006;196(1):65–131. doi: 10.1007/s11511-006-0002-8. [DOI] [Google Scholar]
  • 36.Sturm K-T. On the geometry of metric measure spaces. II. Acta Math. 2006;196(1):133–177. doi: 10.1007/s11511-006-0003-7. [DOI] [Google Scholar]
  • 37.Tuschmann, W., Wiemeler, M.: On the topology of moduli spaces of non-negatively curved Riemannian metrics. Mathematische Annalen, (10.1007/s00208-021-02327-y) (2021)
  • 38.Tuschmann W, Wraith D. Moduli Spaces of Riemannian Metrics. Oberwolfach Seminars. Basel: Springer; 2015. [Google Scholar]
  • 39.Villani C. Optimal transport: old and new. Grundlehren der mathematischen Wissenschaften. Berlin: Springer; 2009. [Google Scholar]
  • 40.Wilking B. On fundamental groups of manifolds of nonnegative curvature. Differ. Geom. Appl. 2000;13(2):129–165. doi: 10.1016/S0926-2245(00)00030-9. [DOI] [Google Scholar]

Articles from Mathematische Annalen are provided here courtesy of Springer

RESOURCES