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

Rigidity and regularity for almost homogeneous spaces with Ricci curvature bounds

Xin Qian School of Mathematical Sciences, Fudan University, Shanghai China [email protected]
Abstract.

We say that a metric space XX is (ϵ,G)(\epsilon,G)-homogeneous if GIso(X)G\leq\operatorname{Iso}(X) is a discrete group of isometries with diam(X/G)ϵ\operatorname{diam}(X/G)\leq\epsilon. A sequence of (ϵi,Gi)(\epsilon_{i},G_{i})-homogeneous spaces XiX_{i} with ϵi0\epsilon_{i}\to 0 is called a sequence of almost homogeneous spaces.

In this paper we show that the Gromov-Hausdorff limit of a sequence of almost homogeneous RCD(K,N)(K,N) spaces must be a nilpotent Lie group with RicK\text{Ric}\geq K. We also obtain a topological rigidity theorem for (ϵ,G)(\epsilon,G)-homogeneous RCD(K,N)(K,N) spaces, which generalizes a recent result by Wang. Indeed, if XX is an (ϵ,G)(\epsilon,G)-homogeneous RCD(K,N)(K,N) space and GG is an almost-crystallographic group, then X/GX/G is bi-Hölder to an infranil orbifold. Moreover, we study (ϵ,G)(\epsilon,G)-homogeneous spaces in the smooth setting and prove rigidity and ϵ\epsilon-regularity theorems for Riemannian orbifolds with Einstein metrics and bounded Ricci curvatures respectively.

1. Introduction

A classical result of Gromov [30] (refined by Ruh [56]) on almost flat manifolds states that for any integer n2n\geq 2, there exists ϵ(n),C(n)>0\epsilon(n),C(n)>0 such that if a closed nn-manifold (M,g)(M,g) satisfies diam(M,g)ϵ(n)\operatorname{diam}(M,g)\leq\epsilon(n) and |secg|1|\operatorname{sec}_{g}|\leq 1, then MM is diffeomorphic to a infranilmanifold 𝒩/Γ\mathcal{N}/\Gamma, where 𝒩\mathcal{N} is a simply connected nilpotent Lie group, and Γ\Gamma is a torsion free discrete subgroup of affine group 𝒩Aut(𝒩)\mathcal{N}\rtimes\operatorname{Aut}(\mathcal{N}) such that [Γ:Γ𝒩]C(n)[\Gamma:\Gamma\cap\mathcal{N}]\leq C(n). This topological control fails if one works on manifolds with bounded Ricci curvature, since even the topology of compact Ricci-flat manifold can exhibit considerable complexity.

However, the nilpotent structure still occurs at the level of fundamental group for manifolds with lower bounded Ricci curvature. In fact, it was proved by Kapovitch-Wilking [40] that there exists ϵ(n),C(n)>0\epsilon(n),C(n)>0 such that for any closed nn-manifold (M,g)(M,g) with diam(M,g)ϵ(n)\operatorname{diam}(M,g)\leq\epsilon(n) and Ricg(n1)\operatorname{Ric}_{g}\geq-(n-1), its fundamental group π1(M)\pi_{1}(M) contains a nilpotent subgroup NN with index [π1(M):N]C(n)[\pi_{1}(M):N]\leq C(n) and rank(N)n\operatorname{rank}(N)\leq n. This theorem is now called the generalized Margulis lemma and it has recently been generalized in the non-smooth setting (i.e., for RCD spaces) in [20]. It is known that any finitely generated nilpotent group is polycyclic and rank(N)\operatorname{rank}(N) is defined as the number of \mathbb{Z} factors in the polycyclic series of NN. In general, following [69], we can define the rank for any finitely generated group GG as the infimum of rank(N)\operatorname{rank}(N) among all finite index polycyclic subgroups NGN\leq G (see Definition 2.24). In particular, for a closed nn-manifold (M,g)(M,g) with diam(M,g)ϵ(n)\operatorname{diam}(M,g)\leq\epsilon(n) and Ricg(n1)\operatorname{Ric}_{g}\geq-(n-1), rank(π1(M))n\operatorname{rank}(\pi_{1}(M))\leq n.

By Naber-Zhang’s results in [52], if π1(M)\pi_{1}(M) attains maximal rank nn, then the universal cover M~\widetilde{M} is volume non-collapsing; Rong termed this the manifold MM satisfying the bounded covering geometry (see [35, 55]). It was proved in [35] that the bounded covering geometry will result in MM being an infra-nilmanifold, which generalized Gromov-Ruh’s theorem on almost flat manifolds. Combining the results in [52, 35], we have the following theorem, which was also mentioned in a recent work by Si-Xu [59].

Theorem 1.1 (52, 35).

There is ϵ(n)>0,v(n)>0\epsilon(n)>0,v(n)>0 such that for any nn-manifold (M,g)(M,g) with Ricg(n1)\operatorname{Ric}_{g}\geq-(n-1) and diam(M)ϵ(n)\operatorname{diam}(M)\leq\epsilon(n), the followings are equivalent:

  1. (1)

    MM is diffeomorphic to an infranilmanifold;

  2. (2)

    rank(π1(M))\operatorname{rank}(\pi_{1}(M)) is equal to nn;

  3. (3)

    (M,g)(M,g) satisfies (1,v(n))(1,v(n))-bounded covering geometry, i.e., vol(B1(x~))v(n)\operatorname{vol}(B_{1}(\tilde{x}))\geq v(n), where x~\tilde{x} is a point in the Riemannian universal cover M~\widetilde{M}.

In [59], Si-Xu proved that if in addition (M,g)(M,g) is Einstein in Theorem 1.1, then (M,g)(M,g) must be flat. Moreover, based on the results of the two most recent works by Zamora-Zhu [69] and Wang [66], which will be reviewed in Section 2, Theorem 1.1 can actually be extended to the RCD setting as following.

Theorem 1.2 (66, 69).

For each KK\in\mathbb{R} and N1N\geq 1, there is ϵ=ϵ(K,N),v=v(K,N)\epsilon=\epsilon(K,N),v=v(K,N) such that for any RCD(K,NK,N) space (X,d,𝔪)(X,d,\mathfrak{m}) with diam(X)ϵ\operatorname{diam}(X)\leq\epsilon, the followings are equivalent:

  1. (1)

    XX is bi-Hölder homeomorphic to an NN-dimensional infranilmanifold 𝒩/Γ\mathcal{N}/\Gamma, where 𝒩\mathcal{N} is a simply connected nilpotent Lie group endowed with a left invariant metric;

  2. (2)

    rank(π1(X))\operatorname{rank}(\pi_{1}(X)) is equal to NN;

  3. (3)

    N(B1(x~))v\mathcal{H}^{N}(B_{1}(\tilde{x}))\geq v, where x~\tilde{x} is a point in the universal cover X~\widetilde{X}.

In this paper, we aim to study such questions in a more general situation. Accordingly, we first introduce the following definition (see [68]).

Definition 1.3.

We say a proper geodesic space XX is (ϵ,G)(\epsilon,G)-homogeneous if GIso(X)G\leq\operatorname{Iso}(X) is a discrete group of isometries with diam(X/G)ϵ\operatorname{diam}(X/G)\leq\epsilon. A sequence of (ϵi,Gi)(\epsilon_{i},G_{i})-homogeneous spaces XiX_{i} with ϵi0\epsilon_{i}\to 0 is called a sequence of almost homogeneous spaces.

For example, if XX is a compact geodesic space with diam(X)ϵ\operatorname{diam}(X)\leq\epsilon, then any normal cover X^\hat{X} is (ϵ,G(X^))(\epsilon,G(\hat{X}))-homogeneous where G(X^)G(\hat{X}) is the group of deck transformations. In particular, the universal cover X~\widetilde{X} is (ϵ,π1(X))(\epsilon,\pi_{1}(X))-homogeneous. In general, the action of GG may not be free. Notice that we require discreteness for the group GG in Definition 1.3 and hence, a homogeneous space is not necessarily an (ϵ,G)(\epsilon,G)-homogeneous space.

The goal of this paper is to study the structure of (ϵ,G)(\epsilon,G)-homogeneous spaces with Ricci curvature bounds. In the case of lower Ricci curvature bounds, we investigate this in the general non-smooth setting (i.e., the RCD setting). However, for two-sided Ricci curvature bounds or Einstein metrics, we are limited to a smooth setting, focusing on Riemannian manifolds and orbifolds. Our first main result is to classify the pointed measured GH-limit of a sequence of almost homogeneous RCD(K,NK,N) spaces.

Theorem 1.4.

Let K,N[1,)K\in\mathbb{R},N\in[1,\infty) and (Xi,di,𝔪i,pi)(X_{i},d_{i},\mathfrak{m}_{i},p_{i}) be a sequence of (ϵi,Gi)(\epsilon_{i},G_{i})-homogeneous pointed RCD(K,N)(K,N) spaces converging to (X,d,𝔪,p)(X,d,\mathfrak{m},p) in the pmGH-sense with ϵi0\epsilon_{i}\to 0. Assume that XX is not a point. Then the followings hold:

  1. (1)

    (X,d)(X,d) is isometric to an nn-dimensional nilpotent Lie group with a left invariant Riemannian metric for some nNn\leq N;

  2. (2)

    if the actions of GiG_{i} are measure-preserving, then 𝔪=cn\mathfrak{m}=c\mathcal{H}^{n} for some c>0c>0 and (X,d,n)(X,d,\mathcal{H}^{n}) is an RCD(K,n)(K,n) space. In particular, RicXK\operatorname{Ric}_{X}\geq K if n2n\geq 2;

  3. (3)

    if XiX_{i} are compact (or equivalently, GiG_{i} are finite), then 𝔪\mathfrak{m} and n\mathcal{H}^{n} are mutually abolutely continuous and (X,d,n)(X,d,\mathcal{H}^{n}) is an RCD(K,n)(K,n) space. In particular, RicXK\operatorname{Ric}_{X}\geq K if n2n\geq 2;

  4. (4)

    if XX is compact, then XX is isometric to a flat torus 𝕋n\mathbb{T}^{n}.

For a metric measure space (Y,d,𝔪)(Y,d,\mathfrak{m}), we say an isometric action gIso(Y)g\in\operatorname{Iso}(Y) is measure preserving if g#𝔪=𝔪g_{\#}\mathfrak{m}=\mathfrak{m}. Notice that when 𝔪\mathfrak{m} is the Hausdorff measure, any isometric action is measure preserving. Also, if we consider a normal covering (Y^,d^,𝔪^)(Y,d,𝔪)(\hat{Y},\hat{d},\hat{\mathfrak{m}})\to(Y,d,\mathfrak{m}), then the construction of the lifted measure 𝔪^\hat{\mathfrak{m}} trivially implies that the deck transformations are measure preserving.

Theorem 1.4 implies that the pGH-limit of a sequence of pointed almost homogeneous manifolds with RicK\text{Ric}\geq K must be isometric to one of the following: a point, S1(r)S^{1}(r), \mathbb{R}, or a nilpotent Lie group with left invariant Riemannian metric and RicK\text{Ric}\geq K.

Based on Gigli’s splitting theorem on RCD(0,N0,N) spaces (see Theorem 2.14), we also obtain the following result.

Theorem 1.5.

Let (Xi,di,𝔪i,pi)(X_{i},d_{i},\mathfrak{m}_{i},p_{i}) be a sequence of almost homogeneous RCD(δi,N)(-\delta_{i},N) spaces converging to (X,d,𝔪,p)(X,d,\mathfrak{m},p) in the pmGH-sense with δi0\delta_{i}\to 0. Then XX is isometric to k×𝕋nk\mathbb{R}^{k}\times\mathbb{T}^{n-k}, where 𝕋nk\mathbb{T}^{n-k} is a flat torus and 0knN0\leq k\leq n\leq N.

We can also prove the following corollary for non-collapsed RCD spaces. Recall that an RCD(K,NK,N) space (X,d,𝔪)(X,d,\mathfrak{m}) is called non-collapsed if 𝔪=N\mathfrak{m}=\mathcal{H}^{N} (see [18]). In this case, NN must be an integer.

Corollary 1.6.

Given v>0,D>0,Kv>0,D>0,K\in\mathbb{R} and N1N\geq 1, there exists ϵ=ϵ(K,N,v,D)\epsilon=\epsilon(K,N,v,D) such that if (X,d,N)(X,d,\mathcal{H}^{N}) is an (ϵ,G)(\epsilon,G)-homogeneous RCD(K,NK,N) space with N(X)v\mathcal{H}^{N}(X)\geq v and diam(X)D{\text{diam}}(X)\leq D, then XX is bi-Hölder homeomorphic to a flat torus 𝕋N\mathbb{T}^{N}. Moreover, if XX is a Riemannian manifold, then XX is diffeomorphic to 𝕋N\mathbb{T}^{N}.

More generally, one may consider those (X,d,N)(X,d,\mathcal{H}^{N}) satisfying local (r,v)(r,v)-bounded covering geometry; that is, for any xXx\in X, N(Br/2(x~))v\mathcal{H}^{N}(B_{r/2}(\tilde{x}))\geq v, where x~\tilde{x} is a pre-image of xx in the (incomplete) universal covering Br(x)~Br(x)\widetilde{B_{r}(x)}\to B_{r}(x). Notice that on an RCD(K,N)(K,N) space, any ball Br(x)B_{r}(x) admits a universal cover by Wang’s result [67] (see Theorem 2.16). Employing Wang’s latest results [66], we have the following fibration theorem for almost homogeneous RCD(K,N)(K,N) spaces satisfying local bounded covering geometry.

Theorem 1.7.

Given v>0,D>0,Kv>0,D>0,K\in\mathbb{R} and N1N\geq 1, there exists ϵ=ϵ(K,N,v,D)\epsilon=\epsilon(K,N,v,D) such that if (X,d,N)(X,d,\mathcal{H}^{N}) is an (ϵ,G)(\epsilon,G)-homogeneous RCD(K,NK,N) space satisfying local (1,v)(1,v)-bounded covering geometry and diam(X)D{\text{diam}}(X)\leq D, then there is a flat torus 𝕋k\mathbb{T}^{k} with 0kN0\leq k\leq N and a continuous fiber bundle map f:X𝕋kf:X\to\mathbb{T}^{k}, which is also an Φ(ϵ|K,N,v,D)\Phi(\epsilon|K,N,v,D)-GHA, where Φ(ϵ|K,N,v,D)0\Phi(\epsilon|K,N,v,D)\to 0 as ϵ0\epsilon\to 0 while K,N,vK,N,v and DD are fixed. Moreover, the ff-fiber with the induced metric is bi-Hölder to an (Nk)(N-k)-dimensional infranilmanifold and in particular, XX is homeomorphic to an infranilmanifold.

In addition, if XX is a Riemannian manifold, then ff is a smooth bundle map and the ff-fiber is diffeomorphic to an (Nk)(N-k)-dimensional infranilmanifold. In particular, XX is diffeomorphic to an infranilmanifold

Notice that diam(X)ϵ{\text{diam}}(X)\leq\epsilon is equivalent to XX being (ϵ,{e})(\epsilon,\left\{e\right\})-homogeneous. In this case, the flat torus 𝕋k\mathbb{T}^{k} in Theorem 1.7 will be a point, and thus this theorem can be seen as a generalization of [35, Theorem A] and [66, Theorem A], which are the “(3) implies (1)” parts of Theorem 1.1 and Theorem 1.2, respectively.

For a proper geodesic space XX, GIso(X)G\leq\operatorname{Iso}(X) is discrete if and only if its action on XX has discrete orbits and is almost free (any isotropy group is finite). If XX is a manifold, then X/GX/G admits a natural orbifold structure [62]. Recall that an infranil orbifold is the quotient of a simply connected nilpotent Lie group 𝒩\mathcal{N} by an almost-crystallographic group (see Definition 2.31). We can then generalize Theorem 1.2 as following.

Theorem 1.8.

For each KK\in\mathbb{R} and N1N\geq 1, there is ϵ=ϵ(K,N),v=v(K,N)\epsilon=\epsilon(K,N),v=v(K,N) such that for any (ϵ,G)(\epsilon,G)-homogeneous RCD(K,NK,N) space (X,d,𝔪)(X,d,\mathfrak{m}), the followings are equivalent:

  1. (1)

    XX is homeomorphic to N\mathbb{R}^{N};

  2. (2)

    XX is a contractible topological NN-manifold without boundary;

  3. (3)

    rank(G)\operatorname{rank}(G) is equal to NN;

  4. (4)

    XX is simply connected and N(B1(x))v\mathcal{H}^{N}(B_{1}(x))\geq v for some xXx\in X;

  5. (5)

    π1(X)\pi_{1}(X) is finite and N(B1(x))v\mathcal{H}^{N}(B_{1}(x))\geq v for some xXx\in X.

Moreover, if GG does not contain a non-trivial finite normal subgroup, then the above conditions are equivalent to the following:

  1. (6)

    X/GX/G is bi-Hölder homeomorphic to an NN-dimensional infranil orbifold 𝒩/Γ\mathcal{N}/\Gamma, where 𝒩\mathcal{N} is a simply connected nilpotent Lie group endowed with a left invariant metric and GG is isomorphic to Γ\Gamma.

If XX/GX\to X/G is a covering map, then GG is torsion free when XX is a contractible manifold. Thus, Theorem 1.8 will indeed yield Theorem 1.2 (see Remark 4.11). In general, we need to assume that GG does not contain a non-trivial finite normal subgroup to identify GG as an almost-crystallographic group (see Theorem 2.32).

The construction of the bi-Hölder homeomorphism in Theorem 1.8 (6) is based on the recent work of Wang [66] and can be adapted to prove diffeomorphism in the smooth case. In fact, the proof of Theorem 1.8 allows us to obtain the following orbifold version of Theorem 1.1. Basic notions and terminology regarding smooth (Riemannian) orbifolds will be reviewed in Section 2.5.

Theorem 1.9.

There is ϵ(n)>0,v(n)>0\epsilon(n)>0,v(n)>0 such that if (𝒪,g)(\mathcal{O},g) is a Riemannian nn-orbifold with Ric(n1)\operatorname{Ric}\geq-(n-1) and diam(𝒪)ϵ(n)\operatorname{diam}(\mathcal{O})\leq\epsilon(n), then the followings are equivalent:

  1. (1)

    |𝒪~||\tilde{\mathcal{O}}| is homeomorphic to n\mathbb{R}^{n}, where |𝒪~||\tilde{\mathcal{O}}| is the underlying topological space of the universal orbifold cover 𝒪~\tilde{\mathcal{O}};

  2. (2)

    rank(π1orb(𝒪))=n\operatorname{rank}(\pi_{1}^{orb}(\mathcal{O}))=n, where π1orb(𝒪)\pi_{1}^{orb}(\mathcal{O}) is the orbifold fundamental group of 𝒪\mathcal{O};

  3. (3)

    vol(B1(x~))v(n)\operatorname{vol}(B_{1}(\tilde{x}))\geq v(n), where x~\tilde{x} is a point in the universal orbifold covering 𝒪~\widetilde{\mathcal{O}}.

Moreover, if 𝒪\mathcal{O} is good and π1orb(𝒪)\pi_{1}^{orb}(\mathcal{O}) does not contain a non-trivial finite normal subgroup, then the above conditions are equivalent to the following:

  1. (4)

    𝒪\mathcal{O} is diffeomorphic to an infranil orbifold.

Recall that a smooth orbifold is called good if it is the global quotient of a smooth manifold by some dicrete group. In the last statement of Theorem 1.9, without assuming 𝒪\mathcal{O} is good, we know from Theorem 1.8 that the underlying space |𝒪||\mathcal{O}| is homeomorphic to an infranil orbifold; with this assumption, we obtain diffeomorphism.

For two-sided Ricci curvature bounds or Einstein metrics, there is currently no comprehensive and rigorous synthetic theory on metric measure spaces. So we are limited to a smooth setting and the following theorem is about the rigidity case in Theorem 1.9 for Einstein orbifolds, which can be seen as an orbifold version of [59, Theorem 0.2].

Theorem 1.10 (Rigidity for Einstein orbifolds).

There is ϵ(n)>0,v(n)>0\epsilon(n)>0,v(n)>0 such that for any closed Einstein nn-orbifold (𝒪,g)(\mathcal{O},g) satisfying Ricλg\operatorname{Ric}\equiv\lambda g with λ(n1)\lambda\geq-(n-1) and diam(𝒪)ϵ(n)\operatorname{diam}(\mathcal{O})\leq\epsilon(n), the followings are equivalent:

  1. (1)

    𝒪\mathcal{O} is a closed flat orbifold;

  2. (2)

    𝒪\mathcal{O} is diffeomorphic to an infranil orbifold;

  3. (3)

    rank(π1orb(𝒪))=n\operatorname{rank}(\pi_{1}^{orb}(\mathcal{O}))=n;

  4. (4)

    vol(B1(x~))v(n)\operatorname{vol}(B_{1}(\tilde{x}))\geq v(n), where x~\tilde{x} is a point in the universal orbifold covering 𝒪~\widetilde{\mathcal{O}}.

Compared with Theorem 1.9, we removed the assumption that 𝒪\mathcal{O} is good and π1orb(𝒪)\pi_{1}^{orb}(\mathcal{O}) does not contain a non-trivial finite normal subgroup, since we can get a flat metric in the Einstein case. Recall that by a result of Thurston [62], a closed flat orbifold must be the quotient orbifold n/Γ\mathbb{R}^{n}/\Gamma, where Γ\Gamma is a crystallographic group.

It is known that almost flat orbifolds are infranil orbifolds (see [21, Proposition 1.4]). Then it follows from Theorem 1.10 that almost flat Einstein orbifolds must be flat.

Corollary 1.11.

There exists ϵ(n)>0\epsilon(n)>0 such that if (𝒪,g)(\mathcal{O},g) is an Einstein nn-orbifold satisfying diam(𝒪)2|secg|ϵ(n)\operatorname{diam}(\mathcal{O})^{2}|\sec_{g}|\leq\epsilon(n), then (𝒪,g)(\mathcal{O},g) is flat.

For the case of bounded Ricci curvature, we have the following ϵ\epsilon-regularity theorem.

Theorem 1.12.

Given v>0v>0 and p(1,)p\in(1,\infty), there exists C=C(n,v,p),ϵ=ϵ(n,v,p)C=C(n,v,p),\epsilon=\epsilon(n,v,p) such that if an (ϵ,G)(\epsilon,G)-homogeneous pointed nn-orbifold (𝒪,g,x)(\mathcal{O},g,x) satisfies |Ricg|n1|\operatorname{Ric}_{g}|\leq n-1 and vol(B1(x))v\operatorname{vol}(B_{1}(x))\geq v, then B1(x)|Rm|pC\int_{B_{1}(x)}|Rm|^{p}\leq C. If, in addition, (𝒪,g)(\mathcal{O},g) is a manifold, then there exists r0=r0(n,v)r_{0}=r_{0}(n,v) such that rh(y)r0r_{h}(y)\geq r_{0} for any yB1(x)y\in B_{1}(x), where rh(y)r_{h}(y) is the C1C^{1}-harmonic radius at yy.

For a manifold with bounded Ricci curvature, a lower bound on C1C^{1}-harmonic radius implies a lower bound on C1,αC^{1,\alpha}-harmonic radius and a local LpL^{p}-bound on curvature, due to elliptic estimates. For Einstein manifolds, this will further lead to higher-order control on curvature. However, under our conditions, the Einstein metric exhibits very strong rigidity and will, in fact, be flat by Theorem 1.10.

Moreover, for manifolds with bounded Ricci curvature, we obtain the following result analogous to Theorem 1.1.

Theorem 1.13.

Given p>n/2p>n/2, there is C(n,p),ϵ(n,p),v(n)C(n,p),\epsilon(n,p),v(n) such that for any nn-manifold (M,g)(M,g) with |Ricg|n1|\operatorname{Ric}_{g}|\leq n-1 and diam(M)ϵ(n,p)\operatorname{diam}(M)\leq\epsilon(n,p), the followings are equivalent:

  1. (1)

    RmLp(M)C(n,p)||Rm||_{L^{p}(M)}\leq C(n,p);

  2. (2)

    MM is diffeomorphic to an nn-dimensional infra-nilmanifold;

  3. (3)

    rank(π1(M))\operatorname{rank}(\pi_{1}(M)) is equal to nn;

  4. (4)

    (M,g)(M,g) satisfies (1,v(n))(1,v(n))-bounded covering geometry, i.e., vol(B1(x~))v(n)\operatorname{vol}(B_{1}(\tilde{x}))\geq v(n), where x~\tilde{x} is a point in the Riemannian universal cover M~\widetilde{M}.

The paper is organized as follows. In Section 2, we cover the preliminary material. In Section 3, we prove Theorem 1.4, Theorem 1.5 and derive a series of consequences. In Section 4, the topological rigidity theorems (Theorem 1.8 and Theorem 1.9) will be proved using results in [69] and [66]. In Section 5, we study the rigidity and ϵ\epsilon-regularity for Einstein orbifolds and orbifolds with bounded Ricci curvature respectively.

Acknowledgements: The author would like to thank Ruobing Zhang for bringing Jikang Wang’s recent paper [66] to his attention.

2. Preliminaries

In this paper, a metric measure space is a triple (X,d,𝔪)(X,d,\mathfrak{m}), where (X,d)(X,d) is a complete, separable and proper metric space and 𝔪\mathfrak{m} is a locally finite non-negative Borel measure on XX with supp𝔪=X{\text{supp}}\mathfrak{m}=X. We will also always assume (X,d)(X,d) to be geodesic, i.e., any couple of points is joined by a length minimizing geodesic.

Throughout this paper, we assume the reader is familiar with the notion and basic theory of RCD(K,N)\text{RCD}(K,N) spaces (N<N<\infty). We refer the reader to [47, 60, 61, 2, 1, 28] for the relevant notions. Notice that a large body of literature has studied the so-called RCD(K,N){}^{*}(K,N) spaces, which are now known to be equivalent to RCD(K,N)(K,N) spaces by the work of Cavelletti-Milman [12] and Li [46]. We refer to [22] for an overview of equivalent definitions of RCD(K,N)\text{RCD}(K,N) spaces.

2.1. Gromov–Hausdorff topology

Definition 2.1.

Let (X,p),(Y,q)(X,p),(Y,q) be pointed geodesic spaces. A map f:XYf:X\to Y is called a pointed ϵ\epsilon-Gromov-Hausdorff approximation (or pointed ϵ\epsilon-GHA) if

(2.1) d(f(p),q)ϵ,d(f(p),q)\leq\epsilon,
(2.2) supx1,x2Bϵ1(p)|d(f(x1),f(x2))d(x1,x2)|ϵ,\sup_{x_{1},x_{2}\in B_{\epsilon^{-1}}(p)}|d(f(x_{1}),f(x_{2}))-d(x_{1},x_{2})|\leq\epsilon,
(2.3) supyBϵ1(q)infxBϵ1(p)d(f(x),y)ϵ.\sup_{y\in B_{\epsilon^{-1}}(q)}\inf_{x\in B_{\epsilon^{-1}}(p)}d(f(x),y)\leq\epsilon.
Definition 2.2.

Let (Xi,pi)(X_{i},p_{i}) be a sequence of pointed proper geodesic spaces. We say that it converges in the pointed Gromov–Hausdorff sense (or pGH-sense) to a pointed proper geodesic space (X,p)(X,p) if there is a sequence of pointed ϵi\epsilon_{i}-GHAs ϕi:XiX\phi_{i}:X_{i}\to X with ϵi0\epsilon_{i}\to 0 as ii\to\infty.

If in addition to that, (Xi,di,𝔪i)(X_{i},d_{i},\mathfrak{m}_{i}), (X,d,𝔪)(X,d,\mathfrak{m}) are metric measure spaces, the maps ϕi\phi_{i} are Borel measurable and Xfd(ϕi)#𝔪iXf𝑑𝔪\int_{X}f\cdot d(\phi_{i})_{\#}\mathfrak{m}_{i}\to\int_{X}f\cdot d\mathfrak{m}, for all continuous f:Xf:X\to\mathbb{R} with compact support, then we say (Xi,di,𝔪i,pi)(X_{i},d_{i},\mathfrak{m}_{i},p_{i}) converges to (X,d,𝔪,p)(X,d,\mathfrak{m},p) in the pointed measured Gromov–Hausdorff sense (or pmGH-sense).

Remark 2.3.

The maps ϕi:XiX\phi_{i}:X_{i}\to X above are called Gromov-Hausdorff approximations and for any xXx\in X, there is a sequence xiXix_{i}\in X_{i} such that ϕi(xi)x\phi_{i}(x_{i})\to x.

Remark 2.4.

If there is a sequence of groups Γi\Gamma_{i} acting on XiX_{i} by (measure preserving) isometries with diam(Xi/Γi)D(X_{i}/\Gamma_{i})\leq D for some D>0D>0, then one could ignore the points pip_{i} when one talks about p(m)GH convergence, as any pair of limits are going to be isomorphic as metric (measure) spaces. In particular, if diam(Xi)D\operatorname{diam}(X_{i})\leq D for some D>0D>0, then we simply say that XiX_{i} converges to XX in the (m)GH-sense.

One of the main features of the RCD(K,NK,N) condition is that it is stable under pmGH convergence, i.e., the pmGH-limit of a sequence of RCD(K,NK,N) spaces is also an RCD(K,NK,N) space (see [29]). Combined with Gromov’s precompactness criterion and Prokhorov’s compactness theorem (see [64, Chapter 27] for instance), the class of RCD(K,NK,N) spaces with normalized measure is compact under the pmGH-topology.

Definition 2.5.

We say that a pointed metric measure space (Y,dY,mY,y)(Y,d_{Y},m_{Y},y) is a tangent cone of (X,d,𝔪)(X,d,\mathfrak{m}) at xx if there exists a sequence ri0+r_{i}\to 0^{+} such that

(X,ri1d,𝔪(Bri(x))1𝔪,x)pmGH(Y,dY,mY,y).(X,r_{i}^{-1}d,\mathfrak{m}(B_{r_{i}}(x))^{-1}\mathfrak{m},x)\xrightarrow{pmGH}(Y,d_{Y},m_{Y},y).

The collection of all tangent cones of (X,d,𝔪)(X,d,\mathfrak{m}) at xx is denoted by Tanx(X,d,𝔪)\text{Tan}_{x}(X,d,\mathfrak{m}).

For an RCD(K,N)(K,N) space (X,d,𝔪)(X,d,\mathfrak{m}), the compactness yields that Tanx(X,d,𝔪)\text{Tan}_{x}(X,d,\mathfrak{m}) is non-empty and any tangent cone is RCD(0,N0,N). We are now in the position to introduce the notions of kk-regular set and essential dimension as follows.

Definition 2.6 (kk-regular set).

For any integer k[1,N]k\in[1,N], we denote by k\mathcal{R}_{k} the set of all points xXx\in X such that Tanx(X,d,𝔪)={(k,dk,(ωk)1k,0k)}\text{Tan}_{x}(X,d,\mathfrak{m})=\left\{(\mathbb{R}^{k},d_{\mathbb{R}^{k}},(\omega_{k})^{-1}\mathcal{H}^{k},0^{k})\right\}, where ωk\omega_{k} is the volume of the unit ball in k\mathbb{R}^{k}. We call k\mathcal{R}_{k} the kk-regular set of XX.

The following result is proved by Bruè-Semola in [10].

Theorem 2.7 (10).

Let (X,d,𝔪)(X,d,\mathfrak{m}) be an RCD(K,N)(K,N) space with KandN(1,)K\in\mathbb{R}\ and\ N\in(1,\infty). Then there exists a unique integer k[1,N]k\in[1,N], called the essential dimension of (X,d,𝔪)(X,d,\mathfrak{m}), denoted by dimess(X)\dim_{ess}(X), such that 𝔪(Xk)=0\mathfrak{m}(X\setminus\mathcal{R}_{k})=0.

2.2. Equivariant Gromov–Hausdorff convergence

There is a well studied notion of convergence of group actions in this setting, first introduced by Fukaya-Yamaguchi in [24]. For a pointed proper metric space (X,p)(X,p), we equip its isometry group Iso(X)\operatorname{Iso}(X) with the metric d0pd_{0}^{p} given by

(2.4) d0p(h1,h2):=infr>0{1r+supxBr(p)d(h1x,h2x)}.d_{0}^{p}(h_{1},h_{2}):=\inf_{r>0}\left\{\frac{1}{r}+\sup_{x\in B_{r}(p)}d(h_{1}x,h_{2}x)\right\}.

for h1,h2Iso(X)h_{1},h_{2}\in\operatorname{Iso}(X). Obviously, we get a left invariant metric that induces the compact-open topology and makes Iso(X)\operatorname{Iso}(X) a proper metric space.

Recall that if a sequence of pointed proper metric spaces (Xi,pi)(X_{i},p_{i}) converges in the pGH sense to the pointed proper metric space (X,p)(X,p), one has a sequence of pointed ϵi\epsilon_{i}-GHAs ϕi:XiX\phi_{i}:X_{i}\to X with ϵi0\epsilon_{i}\to 0.

Definition 2.8.

Consider a sequence of pointed proper metric spaces (Xi,pi)(X_{i},p_{i}) that converges in the pGH sense to a pointed proper metric space (X,p)(X,p), a sequence of closed groups of isometries GiIso(Xi)G_{i}\leq\operatorname{Iso}(X_{i}) and a closed group GIso(X)G\leq\operatorname{Iso}(X). Equip GiG_{i} with the metric d0pid^{p_{i}}_{0} and GG with the metric d0pd^{p}_{0}. We say that the sequence GiG_{i} converges equivariantly to GG if there is a sequence of Gromov-Hausdorff approximations fi:GiGf_{i}:G_{i}\to G such that for each R>0R>0, one has

limisupgBRGi(IdXi)supxBRXi(pi)d(ϕi(gx),fi(g)ϕi(x))=0.\lim\limits_{i\to\infty}\sup_{g\in B^{G_{i}}_{R}(Id_{X_{i}})}\sup_{x\in B^{X_{i}}_{R}(p_{i})}d(\phi_{i}(gx),f_{i}(g)\phi_{i}(x))=0.

Let us recall two basic properties of equivariant convergence proved in [24].

Lemma 2.9 (24).

Let (Yi,qi)(Y_{i},q_{i}) be a sequence of proper geodesic spaces that converges in the pGH sense to (Y,q)(Y,q), and ΓiIso(Yi)\Gamma_{i}\leq\operatorname{Iso}(Y_{i}) a sequence of closed groups of isometries that converges equivariantly to a closed group ΓIso(Y)\Gamma\leq\operatorname{Iso}(Y). Then the sequence (Yi/Γi,[qi])(Y_{i}/\Gamma_{i},[q_{i}]) converges in the pGH sense to (Y/Γ,[q])(Y/\Gamma,[q]).

Lemma 2.10 (24).

Let (Yi,qi)(Y_{i},q_{i}) be a sequence of proper geodesic spaces that converges in the pGH sense to (Y,q)(Y,q), and take a sequence ΓiIso(Yi)\Gamma_{i}\leq\operatorname{Iso}(Y_{i}) of closed groups of isometries. Then there is a subsequence (Yik,qik,Γik)k(Y_{i_{k}},q_{i_{k}},\Gamma_{i_{k}})_{k\in\mathbb{N}} such that Γik\Gamma_{i_{k}} converges equivariantly to a closed group ΓIso(Y)\Gamma\leq\operatorname{Iso}(Y).

A sequence of groups, which converges equivariantly to the trivial group, is called a sequence of small groups. The explicit definition is the following.

Definition 2.11.

Let (Xi,pi)(X_{i},p_{i}) be a sequence of proper geodesic spaces. We say a sequence of groups WiIso(Xi)W_{i}\leq\text{Iso}(X_{i}) consists of small subgroups if for each R>0R>0 we have

limisupgWisupxBR(pi)d(gx,x)=0.\lim_{i\to\infty}\,\sup_{g\in W_{i}}\,\sup_{x\in B_{R}(p_{i})}d(gx,x)=0.

Equivalently, the groups WiW_{i} are small if d0pi(gi,IdXi)0d_{0}^{p_{i}}(g_{i},Id_{X_{i}})\to 0 for any choice of giWig_{i}\in W_{i}.

Obviously, if WiIso(Xi)W_{i}\leq\text{Iso}(X_{i}) is a sequence of discrete small subgroups, then WiW_{i} is a finite group for any large ii. It is proved in [58, Theorem 93] that a non-collapsing sequence of RCD(K,N)RCD(K,N) spaces cannot have small groups of measure preserving isometries.

Theorem 2.12 (58).

Let (Xi,di,𝔪i,pi)(X_{i},d_{i},\mathfrak{m}_{i},p_{i}) be a sequence of pointed RCD(K,N)RCD(K,N) spaces of essential dimension nn and let HiIso(Xi)H_{i}\leq Iso(X_{i}) be a sequence of small subgroups acting by measure preserving isometries. Assume the sequence (Xi,di,𝔪i,pi)(X_{i},d_{i},\mathfrak{m}_{i},p_{i}) converges in the pmGH sense to an RCD(K,N)RCD(K,N) space (X,d,𝔪,p)(X,d,\mathfrak{m},p) of essential dimension nn. Then HiH_{i} is trivial for large ii.

2.3. Properties of RCD(K,N) spaces

In this subsection, we review some properties on RCD(K,NK,N) spaces that we will need in this paper.

Combining Theorem 2.7 and [3, Theorem 4.1], we have the following theorem.

Theorem 2.13.

Let (X,d,𝔪)(X,d,\mathfrak{m}) be an RCD(K,N)(K,N) space with k=dimess(X)k=\dim_{ess}(X). Set

k={xk:limr0+𝔪(Br(x))ωkrk exists and(0,)}.\mathcal{R}_{k}^{*}=\left\{x\in\mathcal{R}_{k}:\ \lim\limits_{r\to 0+}\frac{\mathfrak{m}(B_{r}(x))}{\omega_{k}r^{k}}\text{ exists and}\in(0,\infty)\right\}.

Then we have the following:

  1. (1)

    𝔪(Xk)=0\mathfrak{m}(X\setminus\mathcal{R}_{k}^{*})=0;

  2. (2)

    𝔪k\mathfrak{m}\llcorner\mathcal{R}_{k}^{*} and kk\mathcal{H}^{k}\llcorner\mathcal{R}_{k}^{*} are mutually absolutely continuous;

  3. (3)

    limr0+𝔪(Br(x))ωkrk=d𝔪kdkk(x)\lim\limits_{r\to 0+}\dfrac{\mathfrak{m}(B_{r}(x))}{\omega_{k}r^{k}}=\dfrac{d\mathfrak{m}\llcorner\mathcal{R}_{k}^{*}}{d\mathcal{H}^{k}\llcorner\mathcal{R}_{k}^{*}}(x), for 𝔪\mathfrak{m}-a.e. xkx\in\mathcal{R}_{k}^{*}.

The well known Cheeger–Gromoll splitting theorem [15] was extended by Gigli to the setting of RCD(0,N)(0,N) spaces [27].

Theorem 2.14 (27).

Let (X,d,𝔪)(X,d,\mathfrak{m}) be an RCD(0,N)(0,N) space. Then there is a metric measure space (Y,dY,ν)(Y,d_{Y},\nu) where (Y,dY)(Y,d_{Y}) contains no line, such that (X,d,𝔪)(X,d,\mathfrak{m}) is isomorphic to the product (k×Y,dk×dY,kν)(\mathbb{R}^{k}\times Y,d_{\mathbb{R}^{k}}\times d_{Y},\mathcal{H}^{k}\otimes\nu). Moreover, if Nk[0,1)N-k\in[0,1) then YY is a point, and in general, (Y,dY,ν)(Y,d_{Y},\nu) is an RCD(0,Nk)(0,N-k) space.

Remark 2.15.

If (X,d,𝔪)(X,d,\mathfrak{m}) is an RCD(0,N)(0,N) space and GG is a discrete subgroup of Iso(X)\operatorname{Iso}(X) with diam(X/G)<{\text{diam}}(X/G)<\infty, then in the above splitting theorem, the space YY can be taken to be compact (see [51]). In addition, Iso(X)=Iso(k)×Iso(Y)\operatorname{Iso}(X)=\operatorname{Iso}(\mathbb{R}^{k})\times\operatorname{Iso}(Y).

Let (X,d,𝔪)(X,d,\mathfrak{m}) be an RCD(K,N)(K,N) space and ρ:YX\rho:Y\to X be a covering space. YY has a natural geodesic structure such that for any curve γ:[0,1]Y\gamma:[0,1]\to Y one has

length(ργ)=length(γ).\text{length}(\rho\circ\gamma)=\text{length}(\gamma).

Set

𝒲:={WY open bounded | ρ|W:Wρ(W) is an isometry}\mathcal{W}:=\{W\subset Y\text{ open bounded }|\text{ }\rho_{|_{W}}:W\to\rho(W)\text{ is an isometry}\}

and define a measure 𝔪Y\mathfrak{m}_{Y} on YY by setting 𝔪Y(A):=𝔪(ρ(A))\mathfrak{m}_{Y}(A):=\mathfrak{m}(\rho(A)) for each Borel set AA contained in 𝒲\mathcal{W}. The measure 𝔪Y\mathfrak{m}_{Y} makes ρ:YX\rho:Y\to X a local isomorphism of metric measure spaces, so by the local-to-global property of RCD(K,N)(K,N) space [22], (Y,dY,𝔪Y)(Y,d_{Y},\mathfrak{m}_{Y}) is an RCD(K,N)(K,N) space, and its group of deck transformations acts by measure-preserving isometries (see [51] for more details). Wang proved the following in [67].

Theorem 2.16 (67).

Let (X,d,𝔪)(X,d,\mathfrak{m}) be an RCD(K,N)(K,N) space. Then for any xXx\in X and R>0R>0, there exists r>0r>0 so that any loop in Br(x)B_{r}(x) is contractible in BR(x)B_{R}(x). In particular, XX is semi-locally-simply-connected and its universal cover X~\tilde{X} is simply connected.

Due to Theorem 2.16, for an RCD(K,N)(K,N) space (X,d,𝔪)(X,d,\mathfrak{m}) we can think of its fundamental group π1(X)\pi_{1}(X) as the group of deck transformations of the universal cover X~\tilde{X}.

Recall that an RCD(K,N)(K,N) space (X,d,𝔪)(X,d,\mathfrak{m}) is called non-collapsed if 𝔪=N\mathfrak{m}=\mathcal{H}^{N} (see [18]). There are some equivalent conditions for the non-collapse condition up to a scaling on measure (see [7, Theorem 2.20], [69, Theorem 2.3] and references therein).

Theorem 2.17.

Let (X,𝖽,𝔪)(X,\mathsf{d},\mathfrak{m}) be an RCD(K,N)(K,N) space. Then the following five conditions are equivalent.

  1. (1)

    XX has essential dimension NN.

  2. (2)

    XX has topological dimension NN.

  3. (3)

    XX has Hausdorff dimension NN.

  4. (4)

    𝔪=cN\mathfrak{m}=c\mathcal{H}^{N} for some constant c>0c>0.

  5. (5)

    NN\in\mathbb{N} and XX has Hausdorff dimension greater than N1N-1.

De Philippis-Gigli [18] studied the GH-limit of non-collapsed RCD spaces and obtained the following result, generalizing Cheeger-Colding’s result on Ricci-limit spaces [13].

Theorem 2.18 (18).

Let (Xi,di,N,pi)(X_{i},d_{i},\mathcal{H}^{N},p_{i}) be a sequence of pointed RCD(K,N)\text{RCD}(K,N) spaces and (Xi,di,pi)pGH(X,d,p)(X_{i},d_{i},p_{i})\xrightarrow{pGH}(X,d,p). Then precisely one of the following holds:

  1. (1)

    lim supiN(B1(pi))>0\limsup_{i\to\infty}\mathcal{H}^{N}(B_{1}(p_{i}))>0. In this case, (Xi,di,N,pi)pmGH(X,d,N,p)(X_{i},d_{i},\mathcal{H}^{N},p_{i})\xrightarrow{pmGH}(X,d,\mathcal{H}^{N},p) and the limit limiN(B1(pi))\lim_{i\to\infty}\mathcal{H}^{N}(B_{1}(p_{i})) exists and equal to N(B1(p))\mathcal{H}^{N}(B_{1}(p)).

  2. (2)

    limiN(B1(pi))=0\lim_{i\to\infty}\mathcal{H}^{N}(B_{1}(p_{i}))=0. In this case, dim(X)N1\dim_{\mathcal{H}}(X)\leq N-1.

Let us recall the following topological stability theorem for non-collapsed RCD(K,N)(K,N) spaces, proved by Kapovitch-Mondino [39, Theorem 3.3], based on Cheeger-Colding’s Reifenberg type theorem [13].

Theorem 2.19 (39).

Let (Xi,di,N,pi)(X_{i},d_{i},\mathcal{H}^{N},p_{i}) be a sequence of pointed RCD(K,N)\text{RCD}(K,N) spaces such that the sequence (Xi,pi)(X_{i},p_{i}) converges in the pGH-sense to (MN,p)(M^{N},p) where MNM^{N} is a smooth Riemannian manifold. Then for any R>0R>0, there is a sequence of pointed ϵi\epsilon_{i}-GHAs fi:(Xi,pi)(MN,p)f_{i}:(X_{i},p_{i})\to(M^{N},p) with ϵi0\epsilon_{i}\to 0, such that for all ii large enough depending on RR, the restriction of fif_{i} to BR(pi)B_{R}(p_{i}) is a bi-Hölder homeomorphism onto its image, and

BR4ϵi(p)fi(BR(pi)).B_{R-4\epsilon_{i}}(p)\subset f_{i}(B_{R}(p_{i})).

In particular, if MM is compact then XiX_{i} is bi-Hölder homeomorphic to MM for all large ii.

The bi-Hölder homeomorphism can be constructed via harmonic splitting map. We refer to [9] for the notion of harmonic (k,ϵ)(k,\epsilon)-splitting map on RCD spaces.

Theorem 2.20 (16, 33).

Assume that (X,d,N,p)(X,d,\mathcal{H}^{N},p) is a pointed RCD(ϵ,N)(-\epsilon,N) space and u:B2(p)Nu:B_{2}(p)\to\mathbb{R}^{N} is a harmonic (N,ϵ)(N,\epsilon)-splitting map. Then for any x,yB1(p)x,y\in B_{1}(p) we have

(1Φ(ϵ|N))d(x,y)1+Φ(ϵ|N)d(f(x),f(y))(1+Φ(ϵ|N))d(x,y),(1-\Phi(\epsilon|N))d(x,y)^{1+\Phi(\epsilon|N)}\leq d(f(x),f(y))\leq(1+\Phi(\epsilon|N))d(x,y),

where Φ(ϵ|N)0\Phi(\epsilon|N)\to 0 as ϵ0\epsilon\to 0 while NN is fixed. Moreover, if XX is a smooth NN-manifold with Ricϵ\text{Ric}\geq-\epsilon, then for any xB1(p)x\in B_{1}(p), du:TxXNdu:T_{x}X\to\mathbb{R}^{N} is nondegenerate.

2.4. Nilpotent and polycyclic groups

Definition 2.21.

For a group GG, let G(0):=GG^{(0)}:=G and define inductively G(j+1):=[G(j),G]G^{(j+1)}:=[G^{(j)},G]GG is called nilpotent if G(s)G^{(s)} is the trivial group for some ss\in\mathbb{N}GG is called virtually nilpotent if there exists a nilpotent subgroup NGN\leq G of finite index.

Definition 2.22.

A finitely generated group Λ\Lambda is said to be polycyclic if there is a finite subnormal series

Λ=ΛmΛ0=1\Lambda=\Lambda_{m}\trianglerighteq\cdots\trianglerighteq\Lambda_{0}={1}

with Λj/Λj1\Lambda_{j}/\Lambda_{j-1} cyclic for each jj. Such a subnormal series is called a polycyclic series. The polycyclic rank is defined as the number of jj’s for which Λj/Λj1\Lambda_{j}/\Lambda_{j-1} is isomorphic to , which is independent of the choice of the polycyclic series and denoted by rank(Λ){\text{rank}}(\Lambda).

From the definition, we immediately know that any finite index subgroup of a polycyclic group is a polycyclic group with the same rank..

It is well-known that any finitely generated nilpotent group is polycyclic (see [41] for instance), and the rank of a finitely generated nilpotent group is defined to be the polycyclic rank. The following lemma (see [52, Lemma 2.22 and Lemma 2.24]) gives the definition of the rank of a finitely generated virtually nilpotent group.

Lemma 2.23.

Let GG be a finitely generated virtually nilpotent group. Then:

  1. (1)

    Every nilpotent subgroup NGN\leq G of finite index has the same rank. The common rank is called the rank of GG and also denoted by rank(G)\operatorname{rank}(G).

  2. (2)

    If Γ\Gamma is a finite index subgroup of GG, then rank(Γ)=rank(G)\operatorname{rank}(\Gamma)=\operatorname{rank}(G).

Following [69], we can define the rank for an arbitrary finitely generated group.

Definition 2.24.

For a finitely generated group GG, we define

rank(G):=inf{rank(Λ):Λ is a finite index polycyclic subgroup of G}.{\text{rank}}(G):=\inf\left\{{\text{rank}}(\Lambda):\ \Lambda\text{ is a finite index polycyclic subgroup of }G\right\}.

The infimum of the empty set is defined to be ++\infty.

By Definition 2.22 and Lemma 2.23, if GG is polycyclic or finitely generated virtually nilpotent, there is no conflict between the distinct definitions of rank(G){\text{rank}}(G). Also, if Λ\Lambda is a finite index subgroup of GG, then rank(Λ)=rank(G)\operatorname{rank}(\Lambda)=\operatorname{rank}(G).

We also note that if GG is a discrete group of isometries on a proper geodesic space XX with diam(X/G)(0,){\text{diam}}(X/G)\in(0,\infty), then by [68, Lemma 2.5], GG is finitely generated.

2.5. Riemannian orbifolds

In this subsection, we review the basic theory of orbifolds. An orbifold 𝒪\mathcal{O} is, roughly speaking, a topological space that is locally homeomorphic to a quotient of n\mathbb{R}^{n} by some finite group. We recall the definitions from [45] (see also [25]).

Definition 2.25.

A local model of dimension nn is a pair (U^,G)(\hat{U},G), where U^\hat{U} is an open, connected subset of a Euclidean space n\mathbb{R}^{n}, and GG is a finite group acting smoothly and effectively on U^\hat{U}.

A smooth map (U^1,G1)(U^2,G2)(\hat{U}_{1},G_{1})\to(\hat{U}_{2},G_{2}) between local models (U^i,Gi)(\hat{U}_{i},G_{i}), i=1,2i=1,2, is a homomorphism f:G1G2f_{\sharp}:G_{1}\to G_{2} together with a ff_{\sharp}-equivariant smooth map f^:U^1U^2\hat{f}:\hat{U}_{1}\to\hat{U}_{2}, i.e., f^(γu^)=f(γ)f^(u^)\hat{f}(\gamma\cdot\hat{u})=f_{\sharp}(\gamma)\cdot\hat{f}(\hat{u}), for all γG1\gamma\in G_{1} and u^U^1\hat{u}\in\hat{U}_{1}.

Given a local model (U^,G)(\hat{U},G), denote by UU the quotient U^/G\hat{U}/G. The smooth map between local models is called an embedding if f^\hat{f} is an embedding. In this case, the effectiveness of the actions in the local models implies that ff_{\sharp} is injective.

Definition 2.26.

An nn-dimensional orbifold local chart (Ux,U^x,Gx,πx)(U_{x},\hat{U}_{x},G_{x},\pi_{x}) around a point xx in a topological space XX consists of:

  1. (1)

    A neighborhood UxU_{x} of xx in XX;

  2. (2)

    A local model (U^x,Gx)(\hat{U}_{x},G_{x}) of dimension nn;

  3. (3)

    A GxG_{x}-equivariant projection πx:U^xUx\pi_{x}:\hat{U}_{x}\to U_{x} that induces a homeomorphism U^x/GxUx\hat{U}_{x}/G_{x}\to U_{x}.

If πx1(x)\pi_{x}^{-1}(x) consists of a single point, x^\hat{x}, then (Ux,U^x,Gx,πx)(U_{x},\hat{U}_{x},G_{x},\pi_{x}) is called a good local chart around xx. In particular, x^\hat{x} is fixed by the action of GxG_{x} on U^x\hat{U}_{x}.

Definition 2.27.

An nn-dimensional orbifold atlas for a topological space XX is a collection of nn-dimensional local charts 𝒜={Uα}α\mathcal{A}=\{U_{\alpha}\}_{\alpha} such that the local charts Uα𝒜U_{\alpha}\in\mathcal{A} give an open covering of XX and for any xUαUβx\in U_{\alpha}\cap U_{\beta}, there is a local chart Uγ𝒜U_{\gamma}\in\mathcal{A} with xUγUαUβx\in U_{\gamma}\subset U_{\alpha}\cap U_{\beta} and embeddings (U^γ,Gγ)(U^α,Gα)(\hat{U}_{\gamma},G_{\gamma})\to(\hat{U}_{\alpha},G_{\alpha}), (U^γ,Gγ)(U^β,Gβ)(\hat{U}_{\gamma},G_{\gamma})\to(\hat{U}_{\beta},G_{\beta}).

Two nn-dimensional atlases are called equivalent if they are contained in a third atlas.

Definition 2.28.

An nn-dimensional (smooth) orbifold, denoted by 𝒪n\mathcal{O}^{n} or simply 𝒪\mathcal{O}, is a second-countable, Hausdorff topological space |𝒪||\mathcal{O}|, called the underlying topological space of 𝒪\mathcal{O}, together with an equivalence class of nn-dimensional orbifold atlases.

Given an orbifold 𝒪\mathcal{O} and any point x|𝒪|x\in|\mathcal{O}|, one can always find a good local chart UxU_{x} around xx. Moreover, the corresponding group GxG_{x} does not depend on the choice of good local chart around xx, and is referred to as the local group at xx. From now on, only good local charts will be considered.

Each point x|𝒪|x\in|\mathcal{O}| with Gx={e}G_{x}=\{e\} is called a regular point. The subset |𝒪|reg|\mathcal{O}|_{reg} of regular points is called regular part; it is a a smooth manifold that forms an open dense subset of |𝒪||\mathcal{O}|. A point which is not regular is called singular.

If a discrete group Γ\Gamma acts properly discontinuously on a manifold MM, then the quotient space M/ΓM/\Gamma can be naturally endowed with an orbifold structure. For simplicity, we still use the terminology M/ΓM/\Gamma to mean M/ΓM/\Gamma as an orbifold. An orbifold 𝒪\mathcal{O} is good if 𝒪=M/Γ\mathcal{O}=M/\Gamma for some manifold MM and some discrete group Γ\Gamma.

Similarly, suppose that a discrete group Γ\Gamma acts by diffeomorphisms on an orbifold 𝒪\mathcal{O}. We say that it acts properly discontinuously if the action of Γ\Gamma on |𝒪||\mathcal{O}| is properly discontinuous. Then there is a quotient orbifold 𝒪/Γ\mathcal{O}/\Gamma, with |𝒪/Γ|=|𝒪|/Γ|\mathcal{O}/\Gamma|=|\mathcal{O}|/\Gamma.

A smooth map f:𝒪1𝒪2f:\mathcal{O}_{1}\rightarrow\mathcal{O}_{2} between orbifolds is given by a continuous map |f|:|𝒪1||𝒪2||f|:|\mathcal{O}_{1}|\rightarrow|\mathcal{O}_{2}| with the property that for each p|𝒪1|p\in|\mathcal{O}_{1}|, there are local models (U^1,G1)(\hat{U}_{1},G_{1}) and (U^2,G2)(\hat{U}_{2},G_{2}) for pp and f(p)f(p) respectively, and a smooth map f^:(U^1,G1)(U^2,G2)\hat{f}:(\hat{U}_{1},G_{1})\rightarrow(\hat{U}_{2},G_{2}) between local models so that the diagram

(2.5) U^1f^U^2U1|f|U2\begin{matrix}\hat{U}_{1}&\stackrel{{\scriptstyle\hat{f}}}{{\rightarrow}}&\hat{U}_{2}\\ \downarrow&&\downarrow\\ U_{1}&\stackrel{{\scriptstyle|f|}}{{\rightarrow}}&U_{2}\end{matrix}

commutes. A diffeomorphism f:𝒪1𝒪2f\>:\>\mathcal{O}_{1}\rightarrow\mathcal{O}_{2} is a smooth map with a smooth inverse. In this case, GpG_{p} is isomorphic to Gf(p)G_{f(p)}.

An orbifold covering π:𝒪^𝒪\pi:\hat{\mathcal{O}}\rightarrow\mathcal{O} is a surjective smooth map such that

  1. (1)

    for each x|𝒪|x\in|\mathcal{O}|, there is an orbifold local chart (U,U~,H,ϕ)(U,\tilde{U},H,\phi) around xx such that |π|1(U)|\pi|^{-1}(U) is a disjoint union of open subsets Vi|𝒪^|V_{i}\subset|\hat{\mathcal{O}}|;

  2. (2)

    each ViV_{i} admits an orbifold local chart of the type (Vi,U~,Hi,ϕi)(V_{i},\tilde{U},H_{i},\phi_{i}) where Hi<HH_{i}<H, such that |π||\pi| locally lifts to the identity U~U~\tilde{U}\to\tilde{U} with inclusion HiHH_{i}\to H.

A universal orbifold covering of 𝒪\mathcal{O} is an orbifold covering π:𝒪~𝒪\pi:\tilde{\mathcal{O}}\rightarrow\mathcal{O} such that for every orbifold covering ψ:𝒪^𝒪\psi:\hat{\mathcal{O}}\rightarrow\mathcal{O}, there is an orbifold covering ϕ:𝒪~𝒪^\phi:\tilde{\mathcal{O}}\rightarrow\hat{\mathcal{O}} so that ψϕ=π\psi\circ\phi=\pi. It is due to Thurston [62] that any connected orbifold 𝒪\mathcal{O} admits a universal orbifold covering π:𝒪^𝒪\pi:\hat{\mathcal{O}}\rightarrow\mathcal{O}. The orbifold fundamental group of 𝒪\mathcal{O}, denoted by π1orb(𝒪)\pi_{1}^{orb}(\mathcal{O}), is defined to be the deck transformation group of its universal orbifold covering. The universal orbifold covering π:𝒪~𝒪\pi:\tilde{\mathcal{O}}\rightarrow\mathcal{O} induces a diffeomorphism 𝒪~/π1orb(𝒪)𝒪\tilde{\mathcal{O}}/\pi_{1}^{orb}(\mathcal{O})\to\mathcal{O}.

In general, an orbifold covering is not locally homeomorphism and hence, not a covering. In addition, π1orb(𝒪)\pi_{1}^{orb}(\mathcal{O}) is different from π1(|𝒪|)\pi_{1}(|\mathcal{O}|) and there is actually a epimorphism π1orb(𝒪)π1(|𝒪|)\pi_{1}^{orb}(\mathcal{O})\to\pi_{1}(|\mathcal{O}|) (see [31]).

Definition 2.29 (Riemannian metric on an orbifold).

A Riemannian metric gg on an orbifold 𝒪\mathcal{O} is given by a collection of Riemannian metrics on the local models U^α\hat{U}_{\alpha} so that the following conditions hold:

  1. (1)

    The local group GαG_{\alpha} acts isometrically on U^α\hat{U}_{\alpha}.

  2. (2)

    The embeddings (U^3,G3)(U^1,G1)(\hat{U}_{3},G_{3})\to(\hat{U}_{1},G_{1}) and (U^3,G3)(U^2,G2)(\hat{U}_{3},G_{3})\to(\hat{U}_{2},G_{2}) in the definition of orbifold atlas are isometries (with respect to the Riemannian metric).

Note that the Riemannian metric gg induces a natural metric dd on |𝒪||\mathcal{O}| that is locally isometric to the quotient metric of (U^x,d^x)(\hat{U}_{x},\hat{d}_{x}) by GxG_{x}, where d^x\hat{d}_{x} is induced by the Riemannian metric g^x\hat{g}_{x} on U^x\hat{U}_{x}. In the absence of ambiguity, we sometimes directly treat (𝒪,g)(\mathcal{O},g) as the metric space (|𝒪|,d)(|\mathcal{O}|,d) and apply the terminology from metric spaces to (𝒪,g)(\mathcal{O},g).

For any Riemannian orbifold (𝒪,g)(\mathcal{O},g), there is a natural volume measure volg\operatorname{vol}_{g} given on the local orbifold charts (Ux,U^x,Gx,πx)(U_{x},\hat{U}_{x},G_{x},\pi_{x}) by volg|Ux:=1|Gx|(πx)#volg^x\operatorname{vol}_{g}|_{U_{x}}:=\frac{1}{|G_{x}|}(\pi_{x})_{\#}\operatorname{vol}_{\hat{g}_{x}}, where volg^x\operatorname{vol}_{\hat{g}_{x}} is the Riemannian volume measure on (U^x,g^x)(\hat{U}_{x},\hat{g}_{x}).

The regular part |𝒪|reg|\mathcal{O}|_{reg} inherits a Riemannian metric. The corresponding volume form equals the nn-dimensional Hausdorff measure on |𝒪|reg|\mathcal{O}|_{reg}. In particular, volg(𝒪)\operatorname{vol}_{g}(\mathcal{O}) coincides with the volume of the Riemannian manifold |𝒪|reg|\mathcal{O}|_{reg}, which equals the nn-dimensional Hausdorff measure of the metric space |𝒪||\mathcal{O}|.

The Levi-Civita connection on (𝒪,g)(\mathcal{O},g) can be defined via the local models. We can then define the curvature tensor RmRm on 𝒪\mathcal{O} and derived curvature notions, such as sectional and Ricci curvatures, are defined accordingly. We adopt the same notation for corresponding geometric quantities as is used on Riemannian manifolds.

Letting greg=g||𝒪|regg_{reg}=g|_{|\mathcal{O}|_{reg}}, we have that (|𝒪|reg,greg)(|\mathcal{O}|_{reg},g_{reg}) is a smooth open Riemannian manifold. By density, it is clear that (|𝒪|reg,greg)(|\mathcal{O}|_{reg},g_{reg}) satisfies RicgregK\text{Ric}_{g_{reg}}\geq K if and only if (𝒪,g)(\mathcal{O},g) satisfies RicgK\text{Ric}_{g}\geq K.

Also, the following result was proved by Galaz-Kell-Mondino-Sosa [25, Theorem 7.10].

Theorem 2.30 (25).

Let (𝒪,g)(\mathcal{O},g) be an nn-dimensional Riemannian orbifold. Then RicgK\text{Ric}_{g}\geq K if and only if (𝒪,g,volg)(\mathcal{O},g,\operatorname{vol}_{g}) is an RCD(K,n)(K,n) space.

Finally, let us review some facts about closed flat orbifolds. Recall that a group ΓIso(n)\Gamma\leq\operatorname{Iso}(\mathbb{R}^{n}) is called crystallographic if it is discrete and cocompact, so that n/Γ\mathbb{R}^{n}/\Gamma is a closed flat orbifold. Conversely, by a result of Thurston [62], if (𝒪,g)(\mathcal{O},g) is a closed flat orbifold, then it is good, its universal orbifold cover is n\mathbb{R}^{n} and π1orb(𝒪)\pi_{1}^{orb}(\mathcal{O}) is isomorphic to a crystallographic group.

2.6. Almost-crystallographic groups and infranil orbifolds

In this subsection, we review some basic notions of almost-crystallographic groups and infranil orbifolds.

Definition 2.31.

Let 𝒩\mathcal{N} be a connected and simply connected nilpotent Lie group, and consider a maximal compact subgroup CC of Aut(𝒩)\operatorname{Aut}(\mathcal{N}). A cocompact and discrete subgroup Γ\Gamma of 𝒩Aut(𝒩)\mathcal{N}\rtimes\operatorname{Aut}(\mathcal{N}) is called an almost-crystallographic group (modeled on 𝒩\mathcal{N}). The dimension of Γ\Gamma is defined to be that of 𝒩\mathcal{N}. If moreover, Γ\Gamma is torsion free, then Γ\Gamma is called an almost-Bieberbach group.

An infranil orbifold (resp. infranilmanifold) is a quotient space 𝒩/Γ\mathcal{N}/\Gamma, where Γ\Gamma is an almost-crystallographic (resp. almost-Bieberbach) group modeled on 𝒩\mathcal{N}. If further Γ𝒩\Gamma\subset\mathcal{N} (so Γ\Gamma acts freely on 𝒩\mathcal{N}), then we say that 𝒩/Γ\mathcal{N}/\Gamma is a nilmanifold.

Let Γ\Gamma be an almost-crystallographic group modeled on 𝒩\mathcal{N}. Due to the generalized first Bieberbach Theorem proved by Auslander [5], G=𝒩ΓG=\mathcal{N}\cap\Gamma is a lattice of 𝒩\mathcal{N} and Γ/G\Gamma/G is finite. Therefore, an infranil orbifold is actually the quotient of a nilmanifold by a finite group of affine diffeomorphisms. In addition, rank(Γ)=rank(G)=dim(𝒩)\operatorname{rank}(\Gamma)=\operatorname{rank}(G)=\dim(\mathcal{N}).

Let us recall a well-known algebraic characterization of almost-crystallographic groups (see [19, Theorem 4.2] for instance).

Theorem 2.32.

Let EE be a finitely generated virtually nilpotent group. Then the followings are equivalent.

  1. (1)

    EE is isomorphic to an almost-crystallographic group.

  2. (2)

    EE contains a torsion free nilpotent normal subgroup GG, such that GG is maximal nilpotent in EE and [E:G]<[E:G]<\infty.

  3. (3)

    EE does not contain a non-trivial finite normal subgroup.

To prove “(3) implies (1)” part in the above theorem, one may first find a normal subgroup NN of finite index in EE such that NN is a finitely generated torsion free nilpotent group. It was shown by Mal’cev [49] that such NN can always be embedded as a lattice in a simply connected nilpotent Lie group 𝒩\mathcal{N}, which is unique up to isomorphism and now called the Mal’cev completion of NN. Then EE can be identified to an almost-crystallographic group modeled on 𝒩\mathcal{N}. Specifically, we have the following proposition (see [19] for the detailed proof).

Proposition 2.33.

Let EE be a finitely generated virtually nilpotent group which does not contain a non-trivial finite normal subgroup. Let NN be a torsion free nilpotent normal subgroup of finite index in EE. Then EE is isomorphic to an almost-crystallographic group modeled on 𝒩\mathcal{N}, where 𝒩\mathcal{N} is the Mal’cev completion of NN.

2.7. Limits of almost homogeneous spaces

Let (Xi,pi)(X_{i},p_{i}) be a sequence of pointed almost homogeneous spaces, which converges in the pGH sense to (X,p)(X,p). By Lemma 2.9 and Lemma 2.10, there is a closed group GIso(X)G\leq\text{Iso}(X) acting transitively on XX; that is, XX is GG-homogeneous. Indeed, the limit of almost homogeneous spaces was specifically studied by Zamora in [68], where he utilized the results of Breuillard-Green-Tao [8] and proved the following theorem.

Theorem 2.34 (68).

Let (Xi,pi)(X_{i},p_{i}) be a sequence of pointed almost homogeneous spaces, which converges in the pGH sense to (X,p)(X,p). If XX is semi-locally-simply-connected, then XX is a nilpotent Lie group equipped with a sub-Finsler invariant metric, and π1(X)\pi_{1}(X) is a torsion free subgroup of a quotient of π1(Xi)\pi_{1}(X_{i}) for sufficiently large ii.

Indeed, the fundamental group of any connected nilpotent Lie group is finitely generated torsion free abelian (see [68, Corollary 2.11]). In the above theorem, XX will be simply connected when π1(Xi)\pi_{1}(X_{i}) are finite groups.

It is well-known that any compact connected nilpotent Lie group is abelian (see [68, Corollary 2.13] for instance). Thus, if the above limit space XX is compact, then it must be a torus. We note that this result on compact limits of almost homogeneous spaces can also be obtained in the finite dimensional case by using an old theorem of Turing [63], and Gelander [26] proved a more general result which covers the infinite dimensional case (see also [68, Theorem 1.4]).

A key to proving Theorem 2.34 lies in finding a nilpotent group of isometries acting transitively on XX, which was further applied in the recent work of Zamora-Zhu [69]. The following three lemmas stem from [69].

Lemma 2.35 (69).

Let (Xi,pi)(X_{i},p_{i}) be a sequence of pointed proper geodesic spaces that converges to a pointed proper semi-locally-simply-connected geodesic space (X,p)(X,p) in the pointed Gromov–Hausdorff sense, and GiIso(Xi)G_{i}\leq\text{Iso}(X_{i}) a sequence of discrete groups with diam(Xi/Gi)0{\text{diam}}(X_{i}/G_{i})\to 0. Then there exists ss\in\mathbb{N} and a sequence of finite index normal subgroups GiGiG_{i}^{\prime}\leq G_{i} with

limisupxXisupg(Gi)(s)d(gx,x)=0 and lim supi[Gi:Gi]<.\lim\limits_{i\to\infty}\,\sup_{x\in X_{i}}\,\sup_{g\in(G_{i}^{\prime})^{(s)}}d(gx,x)=0\text{ and }\limsup\limits_{i\to\infty}\,[G_{i}:G_{i}^{\prime}]<\infty.

Note the basic fact that any subgroup of bounded index contains a normal subgroup of bounded index (see [66, Lemma 4.8] for instance). Thus based on [69, Lemma 2.23], we can further assume that GiG_{i}^{\prime} is a normal subgroup.

Lemma 2.36 (69).

Let (Xi,pi)(X_{i},p_{i}), (X,p)(X,p), GiG_{i}, GiG_{i}^{\prime} be as in Lemma 2.35. Then after passing to a subsequence the groups GiG_{i}^{\prime} converge equivariantly to a connected nilpotent group GIso(X)G\leq\text{Iso}(X) acting freely and transitively.

Lemma 2.37 (69).

Let (Xi,pi)(X_{i},p_{i}), (X,p)(X,p), GiG_{i}, GiG_{i}^{\prime} be as in Lemma 2.35. If GiG_{i} satisfy that any sequence of small subgroups is trivial for large ii, then GiG_{i}^{\prime} acts freely for large ii.

We further introduce the following definition and lemma from [8, 68, 69], which gives an explicit description of the groups GiG_{i}^{\prime}.

Definition 2.38.

Let GG be a group, u1,u2,,urGu_{1},u_{2},\ldots,u_{r}\in G, and N1,N2,,NrN_{1},N_{2},\ldots,N_{r} +\in\mathbb{R}^{+}. The set P(u1,,ur;N1,,Nr)GP(u_{1},\ldots,u_{r};N_{1},\ldots,N_{r})\subset G is defined to be the set of elements that can be expressed as words in the uiu_{i}’s and their inverses such that the number of appearances of uiu_{i} and ui1u_{i}^{-1} is not more than NiN_{i}. We then say that P(u1,,ur;N1,,Nr)P(u_{1},\ldots,u_{r};N_{1},\ldots,N_{r}) is a nilprogression in CC-normal form for some C>0C>0 if it also satisfies the following properties:

  1. (1)

    For all 1ijr1\leq i\leq j\leq r, and all choices of signs, we have

    [ui±1,uj±1]P(uj+1,,ur;CNj+1NiNj,,CNrNiNj).[u_{i}^{\pm 1},u_{j}^{\pm 1}]\in P\left(u_{j+1},\ldots,u_{r};\dfrac{CN_{j+1}}{N_{i}N_{j}},\ldots,\dfrac{CN_{r}}{N_{i}N_{j}}\right).

  2. (2)

    The expressions u1n1urnru_{1}^{n_{1}}\ldots u_{r}^{n_{r}} represent distinct elements as n1,,nrn_{1},\ldots,n_{r} range over the integers with |n1|N1/C,,|nr|Nr/C|n_{1}|\leq N_{1}/C,\ldots,|n_{r}|\leq N_{r}/C.

  3. (3)

    One has

    1C(2N1+1)(2Nr+1)|P|C(2N1+1)(2Nr+1).\frac{1}{C}(2\lfloor N_{1}\rfloor+1)\cdots(2\lfloor N_{r}\rfloor+1)\leq|P|\leq C(2\lfloor N_{1}\rfloor+1)\cdots(2\lfloor N_{r}\rfloor+1).

For a nilprogression PP in CC-normal form, and ε(0,1)\varepsilon\in(0,1), the set P(u1,,ur;P(u_{1},\ldots,u_{r}; εN1,,εNr)\varepsilon N_{1},\ldots,\varepsilon N_{r}) also satisfies conditions (1) and (2), and we denote it by εP\varepsilon P. We define the thickness of PP as the minimum of N1,,NrN_{1},\ldots,N_{r} and we denote it by thick(P)\text{thick}(P). The set {u1n1urnr||ni|Ni/C}\{u_{1}^{n_{1}}\ldots u_{r}^{n_{r}}||n_{i}|\leq N_{i}/C\} is called the grid part of PP, and is denoted by G(P)G(P).

Lemma 2.39.

Let (Xi,pi)(X_{i},p_{i}), (X,p)(X,p), GiG_{i}, GiG_{i}^{\prime} be as in Lemma 2.35 and n=dimtop(X)n=\dim_{top}(X). Then for ii large enough, there are N1,i,,Nn,i+N_{1,i},\ldots,N_{n,i}\in\mathbb{R}^{+} and torsion free nilpotent groups Γ~i\tilde{\Gamma}_{i} generated by elements u~1,i,,u~n,iΓ~i\tilde{u}_{1,i},\ldots,\tilde{u}_{n,i}\in\tilde{\Gamma}_{i} with the following properties:

  1. (1)

    There are polynomials Qi:n×nnQ_{i}:\mathbb{R}^{n}\times\mathbb{R}^{n}\to\mathbb{R}^{n} of degree d(n)\leq d(n) giving the group structures on n\mathbb{R}^{n} by x1x2=Qi(x1,x2)x_{1}\cdot x_{2}=Q_{i}(x_{1},x_{2}) such that for each ii, Γ~i\tilde{\Gamma}_{i} is isomorphic to the group (,nQi|×nn)({}^{n},Q_{i}|_{{}^{n}\times{}^{n}}) and the group (n,Qi)(\mathbb{R}^{n},Q_{i}) is isomorphic to the Mal’cev completion of Γ~i\tilde{\Gamma}_{i}.

  2. (2)

    There are small normal subgroups WiGiW_{i}\trianglelefteq G_{i}^{\prime} and surjective group morphisms

    Φi:Γ~iΓi:=Gi/Wi\Phi_{i}:\tilde{\Gamma}_{i}\to\Gamma_{i}:=G_{i}^{\prime}/W_{i}

    such that Ker(Φi)\operatorname{Ker}(\Phi_{i}) is a quotient of π1(Xi)\pi_{1}(X_{i}) and contains an isomorphic copy of π1(X)\pi_{1}(X) for ii large enough.

  3. (3)

    There is C>0C>0 such that if uj,i:=Φi(u~j,i)u_{j,i}:=\Phi_{i}(\tilde{u}_{j,i}) for each j{1,,n}j\in\{1,\ldots,n\}, the set

    Pi:=P(u1,i,,un,i;N1,i,,Nn,i)ΓiP_{i}:=P(u_{1,i},\ldots,u_{n,i};N_{1,i},\ldots,N_{n,i})\subset\Gamma_{i}

    is a nilprogression in CC-normal form with thick(Pi)\text{thick}(P_{i})\to\infty.

  4. (4)

    For each ε>0\varepsilon>0 there is δ>0\delta>0 such that

    G(δPi){gΓi|\displaystyle G(\delta P_{i})\subset\{g\in\Gamma_{i}\,|\, d(g[pi],[pi])ε},\displaystyle d(g[p_{i}],[p_{i}])\leq\varepsilon\},
    {gΓi|d(g[pi],[pi\displaystyle\{g\in\Gamma_{i}\,|\,d(g[p_{i}],[p_{i} ])δ}G(εPi)\displaystyle])\leq\delta\}\subset G(\varepsilon P_{i})

    for ii large enough, where we are considering the action of Γi\Gamma_{i} on Xi/WiX_{i}/W_{i}.

Proof.

Note that [69, Lemma 2.30] provides (2), (3), (4), except that in (2) we additionally obtain that Ker(Φi)\operatorname{Ker}(\Phi_{i}) is a quotient of π1(Xi)\pi_{1}(X_{i}). This is due to [68, Proposition 8.4]. Moreover, (1) is just the Mal’cev Embedding Theorem and the construction of the polynomial QiQ_{i} can be found in [11, Section 5.1] (see also [68, Section 8]). ∎

Remark 2.40.

As noted in [69, Remark 2.31], one obtains that

(2.6) rank(Gi)=rank(Γi)=rank(Γ~i)rank(Ker(Φi))n.{\text{rank}}(G_{i}^{\prime})={\text{rank}}(\Gamma_{i})={\text{rank}}(\tilde{\Gamma}_{i})-{\text{rank}}({\text{Ker}}(\Phi_{i}))\leq n.

If rank(Gi)=n\operatorname{rank}(G_{i}^{\prime})=n, then Ker(Φi){\text{Ker}}(\Phi_{i}) is trivial and hence, XX is simply connected and Γ~i=Γi\tilde{\Gamma}_{i}=\Gamma_{i}.

Remark 2.41.

If π1(Xi)\pi_{1}(X_{i}) are finite groups, then Ker(Φi){\text{Ker}}(\Phi_{i}) will also be trivial, so XX is simply connected and Γ~i=Γi=Gi/Wi\tilde{\Gamma}_{i}=\Gamma_{i}=G_{i}^{\prime}/W_{i}. In addition, rank(Gi)=rank(Gi)=rank(Γ~i)=n{\text{rank}}(G_{i})={\text{rank}}(G_{i}^{\prime})={\text{rank}}(\tilde{\Gamma}_{i})=n.

2.8. Topological rigidity for RCD spaces with bounded covering geometry

As we have noted in Section 1, Theorem 1.1 can be extended to the RCD setting as the following theorem by the recent works of Zamora-Zhu [69] and Wang [66].

Theorem 2.42 (Theorem 1.2).

For each KK\in\mathbb{R} and N1N\geq 1, there is ϵ=ϵ(K,N),v=v(K,N)\epsilon=\epsilon(K,N),v=v(K,N) such that for any RCD(K,NK,N) space (X,d,N)(X,d,\mathcal{H}^{N}) with diam(X)ϵ\operatorname{diam}(X)\leq\epsilon, the followings are equivalent:

  1. (1)

    XX is bi-Hölder homeomorphic to an NN-dimensional infranilmanifold 𝒩/Γ\mathcal{N}/\Gamma, where 𝒩\mathcal{N} is a simply connected nilpotent Lie group endowed with a left invariant metric;

  2. (2)

    rank(π1(X))\operatorname{rank}(\pi_{1}(X)) is equal to NN;

  3. (3)

    N(B1(x~))v\mathcal{H}^{N}(B_{1}(\tilde{x}))\geq v, where x~\tilde{x} is a point in the universal cover X~\widetilde{X}.

In the above theorem, (1) trivially implies (2). It was proved by Zamora-Zhu [69] that (2) implies XX being homeomorphic to an NN-dimensional infranilmanifold and their proof implicitly shows that (2) implies (3). In [66], Wang proved that (3) implies (1).

We say that a non-collapsed RCD(K,N)(K,N) space (X,d,N)(X,d,\mathcal{H}^{N}) satisfies (global) (1,v)(1,v)-bounded covering geometry if condition (3) of Theorem 2.42 holds.

Indeed, a local version of this term was proposed by Rong on Riemannian manifolds. Specifically, let MM be a compact nn-manifold with RicM(n1)\text{Ric}_{M}\geq-(n-1). We say that MM satisfies local (r,v)(r,v)-bounded covering geometry if for any xMx\in M, vol(Br/2(x~))v\operatorname{vol}(B_{r/2}(\tilde{x}))\geq v where x~\tilde{x} is a pre-image of xx in the (incomplete) Riemannian universal covering

π:(Br(x)~,x~)(Br(x),x).\pi:(\widetilde{B_{r}(x)},\tilde{x})\to(B_{r}(x),x).

If MM has small diameter, then local bounded covering geometry is equivalent to (global) bounded covering geometry. We refer to [35, 55] and the survey paper [36] for more detailed descriptions on this terminology.

Naturally, one can define a non-collapsed RCD(K,N)(K,N) space (X,d,N)(X,d,\mathcal{H}^{N}) satisfying local bounded covering geometry in a similar fashion. Note that by Theorem 2.16, any rr-ball Br(x)XB_{r}(x)\subset X is semi-locally-simply-connected and hence, admits a universal cover.

The following fibration theorem summarizes the contributions from [34] and [55].

Theorem 2.43 (34, 55).

Let (Min,gi)(M_{i}^{n},g_{i}) converge to (Nk,g)(N^{k},g) in the pGH-sense, where NkN^{k} is a compact smooth manifold with knk\leq n. Suppose that (Min,gi)(M_{i}^{n},g_{i}) satisfy Ricgi(n1)\text{Ric}_{g_{i}}\geq-(n-1) and local (1,v)(1,v)-bounded covering geometry for some v>0v>0. Then for all large ii, there exist smooth fiber bundle maps fi:MiNf_{i}:M_{i}\to N which are also ϵi\epsilon_{i}-GHAs with ϵi0\epsilon_{i}\to 0. Moreover, any fif_{i}-fiber is diffeomorphic to an (nk)(n-k)-dimensional infranilmanifold.

In [34], Huang constructed the smooth fiber bundle map and Rong [55] further identified the fibers to infranilmanifolds. This theorem is a generalization of Fukaya’s fibration theorem in [23] on collapsed manifolds with bounded sectional curvature (see also [14]). Indeed, any nn-manifold with |sec|1|\sec|\leq 1 satisfies local (r,v)(r,v)-bounded covering geometry with rr and vv depending only on nn (see [14]).

Recently, Wang [66] generalized Theorem 2.43 to RCD(K,N)(K,N) spaces (X,d,N)(X,d,\mathcal{H}^{N}).

Theorem 2.44 (66).

Let (Xi,di,N)(X_{i},d_{i},\mathcal{H}^{N}) be a sequence of RCD(K,N)(K,N) spaces and (Xi,di)(X_{i},d_{i}) converge in the GH-sense to a closed Riemannian manifold NkN^{k} with kNk\leq N. Suppose that all (Xi,di,N)(X_{i},d_{i},\mathcal{H}^{N}) satisfy local (1,v)(1,v)-bounded covering geometry for some v>0v>0. Then for large enough ii, there are fiber bundle maps fi:XiNkf_{i}:X_{i}\to N^{k} which are also ϵi\epsilon_{i}-GHAs with ϵi0\epsilon_{i}\to 0 and any fif_{i}-fiber with the induced metric is bi-Hölder homeomorphic to an (Nk)(N-k)-dimensional infranilmanifold.

3. Limits of almost homogeneous RCD spaces and applications

In this section, we will prove Theorem 1.4 and Theorem 1.5 and derive a series of consequences. The following lemma is obvious but key to our proof of Theorem 1.4 (2).

Lemma 3.1.

Let (Xi,di,mi,qi)(X_{i},d_{i},m_{i},q_{i}) be a sequence of metric measure spaces that converges in the pmGH sense to (X,d,m,q)(X,d,m,q), and GiIso(Xi)G_{i}\leq\operatorname{Iso}(X_{i}) a sequence of closed groups of measure preserving isometries that converges equivariantly to a closed group GIso(X)G\leq\operatorname{Iso}(X). Then GG acts on XX by measure preserving isometries.

Proof.

Let gg be an arbitrary element of the group GG. We need to show that g#m=mg_{\#}m=m.

Notice that by Definition 2.8, there is a sequence of Gromov-Hausdorff approximations fi:GiGf_{i}:G_{i}\to G and giGig_{i}\in G_{i}, such that fi(gi)gf_{i}(g_{i})\to g. Thus, fi(gi)#mg#mf_{i}(g_{i})_{\#}m\to g_{\#}m in Cc(X)C_{c}(X)^{*}. Also, there is a sequence of Gromov-Hausdorff approximations ϕi:XiX\phi_{i}:X_{i}\to X such that (ϕi)#mim(\phi_{i})_{\#}m_{i}\to m in Cc(X)C_{c}(X)^{*}. Then we have (fi(gi)ϕi)#mig#m(f_{i}(g_{i})\circ\phi_{i})_{\#}m_{i}\to g_{\#}m in Cc(X)C_{c}(X)^{*}.

By Definition 2.8, (ϕigi)#mig#m(\phi_{i}\circ g_{i})_{\#}m_{i}\to g_{\#}m in Cc(X)C_{c}(X)^{*}. Since (gi)#mi=mi(g_{i})_{\#}m_{i}=m_{i}, we obtain that (ϕi)#mig#m(\phi_{i})_{\#}m_{i}\to g_{\#}m in Cc(X)C_{c}(X)^{*}. This leads to g#m=mg_{\#}m=m and we complete the proof. ∎

We say that a metric measure space (X,d,m)(X,d,m) is metric measure homogeneous if for all x,yXx,y\in X, there exists a measure preserving isometry hIso(X)h\in\operatorname{Iso}(X) such that h(x)=yh(x)=y.

Lemma 3.2.

Let (X,d,m)(X,d,m) be a metric measure homogeneous RCD(K,N)(K,N) space of essential dimension nn for some KK\in\mathbb{R} and N[1,)N\in[1,\infty). Then XX is isometric to a Riemannian nn-manifold and m=cnm=c\mathcal{H}^{n} for some c>0c>0. In particular, (X,d,n)(X,d,\mathcal{H}^{n}) is a non-collapsed RCD(K,n)(K,n) space.

Proof.

By [57, Proposition 5.14], XX is isometric to a Riemannian nn-manifold. Due to homogeneity, X=nX=\mathcal{R}_{n}^{*} (see Theorem 2.13) and the limit

limr0+m(Br(x))ωnrn\lim\limits_{r\to 0^{+}}\frac{m(B_{r}(x))}{\omega_{n}r^{n}}

is a constant denoted by cc. Then by Theorem 2.13, we have m=cnm=c\mathcal{H}^{n} and hence, (X,d,n)(X,d,\mathcal{H}^{n}) is RCD(K,N)(K,N). Since XX is a Riemannian nn-manifold, (X,d,n)(X,d,\mathcal{H}^{n}) is an RCD(K,n)(K,n) space. This completes the proof. ∎

Remark 3.3.

In a recent work [32], Honda-Nepechiy arrived at the same conclusion as Lemma 3.2 by assuming only that (X,d,m)(X,d,m) is locally metric measure homogeneous. For our purposes, Lemma 3.2 is sufficient, and the proof is considerably simpler.

Now, we can prove Theorem 1.4 and for the convenience of readers, we rewrite it here.

Theorem 3.4 (Theorem 1.4).

Let K,N(1,)K\in\mathbb{R},N\in(1,\infty) and (Xi,di,mi,pi)(X_{i},d_{i},m_{i},p_{i}) be a sequence of (ϵi,Gi)(\epsilon_{i},G_{i})-homogeneous pointed RCD(K,N)(K,N) spaces converging to (X,d,m,p)(X,d,m,p) in the pmGH-sense with ϵi0\epsilon_{i}\to 0. Assume that XX is not a point. Then the followings hold:

  1. (1)

    (X,d)(X,d) is isometric to an nn-dimensional nilpotent Lie group with a left invariant Riemannian metric for some nNn\leq N;

  2. (2)

    if the actions of GiG_{i} are measure-preserving, then m=cnm=c\mathcal{H}^{n} for some c>0c>0 and (X,d,n)(X,d,\mathcal{H}^{n}) is an RCD(K,n)(K,n) space. In particular, RicXK\operatorname{Ric}_{X}\geq K if n2n\geq 2;

  3. (3)

    if XiX_{i} are compact (or equivalently, GiG_{i} are finite), then mm and n\mathcal{H}^{n} are mutually abolutely continuous and (X,d,n)(X,d,\mathcal{H}^{n}) is an RCD(K,n)(K,n) space. In particular, RicXK\operatorname{Ric}_{X}\geq K if n2n\geq 2;

  4. (4)

    if XX is compact, then XX is isometric to a flat torus 𝕋n\mathbb{T}^{n}.

Proof.

(1) By Lemma 2.9 and Lemma 2.10, (X,d,m)(X,d,m) is an homogeneous RCD(K,N)(K,N) space. Then by [57, Proposition 5.14], XX is a Riemannian nn-manifold, where n=dimess(X)Nn=\dim_{ess}(X)\leq N. Combining with Theorem 2.34, (X,d)(X,d) is isometric to an nn-dimensional nilpotent Lie group with a left invariant Riemannian metric.

(2) By Lemma 2.10 and Lemma 3.1, there is a closed group GIso(X)G\leq\operatorname{Iso}(X) acting transitively on XX by measure preserving isometries. Then by Lemma 3.2, m=cnm=c\mathcal{H}^{n} for some c>0c>0 and (X,d,n)(X,d,\mathcal{H}^{n}) is an RCD(K,n)(K,n) space. Since XX is a Riemannian n-manifold, RicXK\operatorname{Ric}_{X}\geq K if n2n\geq 2.

(3) Since GiG_{i} is a finite group, we can apply [57, Theorem A] to obtain a GiG_{i}-invariant measure mGim_{G_{i}} (so GiG_{i} acts on (Xi,di,mGi)(X_{i},d_{i},m_{G_{i}}) by measure preserving isometries) such that (Xi,di,mGi)(X_{i},d_{i},m_{G_{i}}) is also RCD(K,N)(K,N). After normalizing the measure mGim_{G_{i}} and passing to a subsequence, we can assume (Xi,di,mGi,pi)(X_{i},d_{i},m_{G_{i}},p_{i}) converges in the pmGH-sense to (X,d,m,p)(X,d,m^{*},p). By (2), m=cnm^{*}=c\mathcal{H}^{n} for some c>0c>0 and (X,d,n)(X,d,\mathcal{H}^{n}) is an RCD(K,n)(K,n) space. Also, mm and n\mathcal{H}^{n} are mutually absolutely continuous due to [42].

(4) By (1), XX is a compact connected nilpotent Lie group and hence, a torus. Since the metric is invariant and Riemannian, XX is a flat torus. ∎

Remark 3.5.

One may prove (4) without using the nilpotency obtained in (1). In fact, if XX is compact, then GiG_{i} are finite groups and the orbits GipiG_{i}\cdot p_{i} are finite homogeneous metric spaces. Notice that GipiG_{i}\cdot p_{i} converges in the GH-sense to XX. Then by [26, Theorem 1.1] and [6, Theorem 2.2.4], XX is a torus with an invariant metric. Since XX is a Riemannian manifold, XX is a flat torus.

Below, we will present a number of results that follow from Theorem 1.4. The following is readily obtained from Theorem 1.4 and Theorem 2.34 (see also Remark 2.41).

Proposition 3.6.

Let KK\in\mathbb{R}, N(1,)N\in(1,\infty) and (Xi,di,mi,pi)(X_{i},d_{i},m_{i},p_{i}) be a sequence of almost homogeneous pointed RCD(K,N)(K,N) spaces converging to (X,d,m,p)(X,d,m,p) in the pmGH-sense. Assume that π1(Xi)\pi_{1}(X_{i}) are finite groups. Then XX is isometric to a simply connected nilpotent Lie group with a left invariant Riemannian metric. In particular, XX is diffeomorphic to n\mathbb{R}^{n}, where n=dim(X)n=\dim(X).

If we additionally assume that XiX_{i} are compact in Proposition 3.6, then the sequence XiX_{i} must be collapsed.

Proposition 3.7.

Let (Xi,di,mi,pi)(X_{i},d_{i},m_{i},p_{i}) and (X,d,m,p)(X,d,m,p) be as in Proposition 3.6. Additionally, assume XiX_{i} are compact and not single points. Then dim(X)lim infidimess(Xi)1\dim(X)\leq\liminf\limits_{i\to\infty}\dim_{ess}(X_{i})-1.

Proof.

We can assume without loss of generality that dimess(Xi)=k1\dim_{ess}(X_{i})=k\geq 1 for all ii. Then by [44, Theorem 1.5], dim(X)k\dim(X)\leq k. Let (Xi,di)(X_{i},d_{i}) be (ϵi,Gi)(\epsilon_{i},G_{i})-homogeneous with ϵi0\epsilon_{i}\to 0. Since GiG_{i} are finite groups, we can apply [57, Theorem A] to obtain GiG_{i}-invariant measures mGim_{G_{i}} such that (Xi,di,mGi)(X_{i},d_{i},m_{G_{i}}) are also RCD(K,N)(K,N). Recall that the essential dimension is invariant under changes of measure (see [7, Remark 2.12] for instance).

Let us now argue by contradiction. Suppose that dim(X)=k\dim(X)=k. Then it follows from Theorem 2.12 that any sequence of small subgroups WiIso(Xi)W_{i}\leq\operatorname{Iso}(X_{i}) will be trivial for large enough ii. Then by Lemma 2.35 and Lemma 2.37, there exist subgroups GiGiG_{i}^{\prime}\leq G_{i} such that diam(Xi/Gi)0{\text{diam}}(X_{i}/G_{i}^{\prime})\to 0 and GiG_{i}^{\prime} acts freely on XiX_{i} for large ii. So XiXi/GiX_{i}\to X_{i}/G_{i}^{\prime} is a covering. Let X~i\tilde{X}_{i} be the universal cover of XiX_{i} and Xi/GiX_{i}/G_{i}^{\prime}. Note that X~i\tilde{X}_{i} are compact.

By [25, Theorem 7.24], Xi/GiX_{i}/G_{i}^{\prime} endowed with the quotient metric and quotient measure is an RCD(K,N)(K,N) space. Then it follows from [58, Theorem 2] that diam(X~i)0{\text{diam}}(\tilde{X}_{i})\to 0. Hence diam(Xi)0{\text{diam}}(X_{i})\to 0 and this leads to a contradiction. ∎

For an RCD(K,N)(K,N) space with K>0K>0, the Bonnet-Myers theorem on RCD spaces [61] will lead to a uniform diameter upper bound and the finiteness of the fundamental group. So the following corollary follows from Proposition 3.6.

Corollary 3.8.

Let (Xi,di,mi)(X_{i},d_{i},m_{i}) be a sequence of almost homogeneous RCD(K,N)(K,N) spaces for some K>0K>0 and N(1,)N\in(1,\infty). Then diam(Xi)0{\text{diam}}(X_{i})\to 0.

Remark 3.9.

Corollary 3.8 can also be obtained from (3) and (4) in Theorem 1.4.

Recall that if an RCD(0,N)(0,N) space (X,d,m)(X,d,m) admits a discrete cocompact group GIso(X)G\leq\operatorname{Iso}(X), then XX splits as k×Y\mathbb{R}^{k}\times Y where YY is a compact RCD(0,Nk)(0,N-k) space (see Theorem 2.14 and Remark 2.15). We can derive the following corollary directly from Proposition 3.6 and proof by contradiction.

Corollary 3.10.

Let (X,d,m)(X,d,m) be an RCD(0,N)(0,N) space for some N1N\geq 1. Assume that π1(X)\pi_{1}(X) is finite and GG is a discrete subgroup of Iso(X)\operatorname{Iso}(X) with diam(X/G)<{\text{diam}}(X/G)<\infty. Then (X,d,m)(X,d,m) is isomorphic to (k×Y,dk×dY,kν)(\mathbb{R}^{k}\times Y,d_{\mathbb{R}^{k}}\times d_{Y},\mathcal{H}^{k}\otimes\nu), where (Y,dY,ν)(Y,d_{Y},\nu) is a compact RCD(0,Nk)(0,N-k) space with diam(Y)C(N)diam(X/G)\operatorname{diam}(Y)\leq C(N)\cdot\operatorname{diam}(X/G).

Proof.

Without loss of generality, we can assume that diam(X/G)=1{\text{diam}}(X/G)=1. Let us argue by contradiction. Suppose that there is a sequence of (1,Gi)(1,G_{i})-homogeneous RCD(0,N)(0,N) spaces (Xi,di,mi)(X_{i},d_{i},m_{i}) which are isomorphic to ki×Yi\mathbb{R}^{k_{i}}\times Y_{i}, where YiY_{i} are compact RCD(0,Nki)(0,N-k_{i}) spaces with finite fundamental groups and diam(Yi){\text{diam}}(Y_{i})\to\infty. We may also assume kikk_{i}\equiv k.

Let ri=diam(Yi)r_{i}={\text{diam}}(Y_{i}). Then ri1(k×Yi)k×Yr_{i}^{-1}(\mathbb{R}^{k}\times Y_{i})\to\mathbb{R}^{k}\times Y in the pGH-sense for some space YY with diam(Y)=1{\text{diam}}(Y)=1. Due to Proposition 3.6, this is a contradiction. ∎

The above corollary is also obtained by Pan-Rong in [53], where they considered Riemannian manifolds with Ric0\text{Ric}\geq 0.

Theorem 1.5 is a direct consequence of Theorem 1.4 and Theorem 2.14.

Proof of Theorem 1.5.

It follows from Theorem 1.4 and Theorem 2.14 that XX is isometric to k×Nnk\mathbb{R}^{k}\times N^{n-k} for some 0knN0\leq k\leq n\leq N, where NnkN^{n-k} is a nilpotent Lie group with a left invariant Riemannian metric which contains no lines. From the proof of Theorem 1.4, we know that XX is homogeneous which implies that NnkN^{n-k} is homogeneous. Then NnkN^{n-k} is compact and hence, is a torus. Since the metric is Riemannian and invariant, NnkN^{n-k} is a flat torus 𝕋nk\mathbb{T}^{n-k}. ∎

Remark 3.11.

We note that Theorem 1.5 can also be derived from Theorem 1.4 and [37, Theorem 1.1].

Theorem 1.5 and Proposition 3.6 simply implies the following corollary.

Corollary 3.12.

Let (Xi,di,mi,pi)(X_{i},d_{i},m_{i},p_{i}) be a sequence of almost homogeneous RCD(δi,N)(-\delta_{i},N) spaces converging to (X,d,m,p)(X,d,m,p) in the pmGH-sense with δi0\delta_{i}\to 0. Suppose that π1(Xi)\pi_{1}(X_{i}) are finite groups. Then XX is isometric to n\mathbb{R}^{n} for some nNn\leq N.

The proof of Corollary 1.6 and Theorem 1.7 is immediate.

Proof of Corollary 1.6.

We argue by contradiction. Suppose that there is a sequence of almost homogeneous RCD(K,N)(K,N) spaces (Xi,di,N)(X_{i},d_{i},\mathcal{H}^{N}) with N(Xi)v\mathcal{H}^{N}(X_{i})\geq v and diam(Xi)D{\text{diam}}(X_{i})\leq D and all XiX_{i} are not bi-Hölder homeomorphic to flat torus 𝕋N\mathbb{T}^{N}. Up to a subsequence, XiX_{i} converges in the GH-sense to YY and by Theorem 1.4 (4) and Theorem 2.18, YY is isometric to a flat torus 𝕋N\mathbb{T}^{N}. Then by Theorem 2.19, XiX_{i} is bi-Hölder homeomorphic to 𝕋N\mathbb{T}^{N} for sufficiently large ii, which leads to a contradiction.

Moreover, if XX is a Riemannian manifold, then we can show that XX is diffeomorphic to 𝕋N\mathbb{T}^{N} by the same argument applying [13, Theorem A.1.12] instead of Theorem 2.19. ∎

Proof of Theorem 1.7.

Using Theorem 1.4 (4), Theorem 2.44 and proof by contradiction, the proof is easily obtained. For the smooth case, substitute Theorem 2.43 for Theorem 2.44 and the conclusion follows from the same argument. ∎

4. Topological rigidity of almost homogeneous non-collapsed RCD spaces

The main purpose of this section is to prove Theorem 1.8. The proof will be divided into two parts (Theorem 4.4 and Theorem 4.5), based on Zamora-Zhu’s results in [69] and Wang’s arguments in [66] respectively. In addition, we will simultaneously obtain a proof of Theorem 1.9.

Let us first review the following theorem from [69].

Theorem 4.1 (69).

For each KK\in\mathbb{R} and N1N\geq 1, there is ϵ=ϵ(K,N)\epsilon=\epsilon(K,N) such that if (X,d,m)(X,d,m) is an (ϵ,G)(\epsilon,G)-homogeneous RCD(K,N)(K,N) space, then rank(G)N\operatorname{rank}(G)\leq N and in the case of equality, XX is homeomorphic to N\mathbb{R}^{N}.

The contractibility of XX is particular powerful when paired with the following theorem, which is an observation by Kapovitch in [38]. We refer to [38] (see also [69]) for the proof.

Theorem 4.2 (38).

Let MM be a closed aspherical topological manifold with π1(M)\pi_{1}(M) virtually nilpotent. Then MM is homeomorphic to an infranilmanifold.

By the definition of rank(G)\operatorname{rank}(G) (see Definition 2.24), the group GG in Theorem 4.1 is virtually polycyclic. Indeed, due to Breuillard-Green-Tao’s result [8], GG is virtually nilpotent.

Lemma 4.3.

For each KK\in\mathbb{R} and N1N\geq 1, there is ϵ=ϵ(K,N)\epsilon=\epsilon(K,N) such that if (X,d,m)(X,d,m) is an (ϵ,G)(\epsilon,G)-homogeneous RCD(K,N)(K,N) space, then GG is finitely generated virtually nilpotent with rank(G)N\operatorname{rank}(G)\leq N.

Proof.

Fix pXp\in X. By [68, Lemma 2.5], GG is generated by the set

S:={gG:d(gp,p)3diam(X/G)}.S:=\left\{g\in G:\ d(gp,p)\leq 3\cdot{\text{diam}}(X/G)\right\}.

Then by [8, Corollary 1.15], there is small ϵ=ϵ(K,N)\epsilon=\epsilon(K,N) such that GG is finitely generated virtually nilpotent. It follows from Theorem 4.1 that rank(G)N\operatorname{rank}(G)\leq N. ∎

When rank(G){\text{rank}}(G) attains its maximum value NN, XX is homeomorphic to N\mathbb{R}^{N} (Theorem 4.1) and in fact, the converse also holds. In addition, we can prove the first part of Theorem 1.8, which demonstrate a set of conditions equivalent to maximal rank. Also, note that the first statement in Theorem 1.9 is just a corollary.

Theorem 4.4.

For each KK\in\mathbb{R} and N1N\geq 1, there is ϵ=ϵ(K,N),v=v(K,N)\epsilon=\epsilon(K,N),v=v(K,N) such that for any (ϵ,G)(\epsilon,G)-homogeneous RCD(K,NK,N) space (X,d,m)(X,d,m), the followings are equivalent:

  1. (1)

    XX is homeomorphic to N\mathbb{R}^{N};

  2. (2)

    XX is a contractible topological NN-manifold without boundary;

  3. (3)

    rank(G)\operatorname{rank}(G) is equal to NN;

  4. (4)

    XX is simply connected and N(B1(x))v\mathcal{H}^{N}(B_{1}(x))\geq v for some xXx\in X;

  5. (5)

    π1(X)\pi_{1}(X) is finite and N(B1(x))v\mathcal{H}^{N}(B_{1}(x))\geq v for some xXx\in X.

Proof.

We will prove that (3)\Rightarrow(4)\Rightarrow(5)\Rightarrow(3) and (1)\Rightarrow(2)\Rightarrow(3)\Rightarrow(1).

(3)\Rightarrow(4): By Theorem 4.1 and Theorem 2.17, XX is simply connected and m=cNm=c\mathcal{H}^{N} for some c>0c>0. We only need to show N(B1(x))v\mathcal{H}^{N}(B_{1}(x))\geq v for some xXx\in X.

By contradiction, we assume that there is a sequence of (ϵi,Gi)(\epsilon_{i},G_{i})-homogeneous RCD(K,N)(K,N) spaces (Xi,di,N)(X_{i},d_{i},\mathcal{H}^{N}) with ϵi0\epsilon_{i}\to 0, such that rank(Gi)=N{\text{rank}}(G_{i})=N and N(B1(xi))0\mathcal{H}^{N}(B_{1}(x_{i}))\to 0 for some sequence xiXix_{i}\in X_{i}. By compactness, Theorem 1.4 and Theorem 2.18, we can assume (Xi,xi)(X_{i},x_{i}) converges in the pGH-sense to (X,x)(X,x), where XX is a nilpotent Lie group of dimension nN1n\leq N-1. Let GiG_{i}^{\prime} be as in Lemma 2.35. Then by Remark 2.40, rank(Gi)N1{\text{rank}}(G_{i}^{\prime})\leq N-1 which implies rank(Gi)N1{\text{rank}}(G_{i})\leq N-1. This leads to a contradiction.

(4)\Rightarrow(5): This is trivial.

(5)\Rightarrow(3): Notice that dim(X)=N\dim_{\mathcal{H}}(X)=N and by Theorem 2.17, m=cNm=c\mathcal{H}^{N} for some c>0c>0. We then argue by contradiction. Assume that there is a sequence of (ϵi,Gi)(\epsilon_{i},G_{i})-homogeneous RCD(K,N)(K,N) spaces (Xi,di,N)(X_{i},d_{i},\mathcal{H}^{N}) with ϵi0\epsilon_{i}\to 0, such that π1(Xi)\pi_{1}(X_{i}) are finite, rank(Gi)<N{\text{rank}}(G_{i})<N and N(B1(xi))v>0\mathcal{H}^{N}(B_{1}(x_{i}))\geq v>0 for some sequence xiXix_{i}\in X_{i}. Due to compactness, Theorem 1.4 and Theorem 2.18, we can assume (Xi,xi)(X_{i},x_{i}) converges in the pGH-sense to (X,x)(X,x), where XX is a nilpotent Lie group of dimension NN. Then by Lemma 2.39 and Remark 2.41, rank(Gi)=N{\text{rank}}(G_{i})=N which leads to a contradiction.

(1)\Rightarrow(2): This is trivial.

(2)\Rightarrow(3): By Lemma 4.3, GG is a finitely generated virtually nilpotent group. Then GG contains a torsion free nilpotent subgroup of finite index (see [41]), denoted by Γ\Gamma. Notice that ΓIso(X)\Gamma\leq\operatorname{Iso}(X) is a discrete group acting freely on XX. Since XX is a contractible topological NN-manifold, X/ΓX/\Gamma is a closed aspherical topological manifold. Then by Theorem 4.2, X/ΓX/\Gamma is homeomorphic to an NN-dimensional nilmanifold. So Γ=π1(X/Γ)\Gamma=\pi_{1}(X/\Gamma) has rank NN and thus, rank(G)=N{\text{rank}}(G)=N.

(3)\Rightarrow(1): This follows from Theorem 4.1. ∎

We now proceed to prove the last statement in Theorem 1.8 and Theorem 1.9. Notice that we only need to work on non-collapsed RCD(K,N)(K,N) spaces (X,d,N)(X,d,\mathcal{H}^{N}). Also, if GG is isomorphic to an almost-crystallographic group of dimension NN, then rank(G)=N\operatorname{rank}(G)=N. Therefore, we only need to show the following theorem.

Theorem 4.5.

For each KK\in\mathbb{R} and N1N\geq 1, there is ϵ=ϵ(K,N)\epsilon=\epsilon(K,N) such that for any (ϵ,G)(\epsilon,G)-homogeneous RCD(K,NK,N) space (X,d,N)(X,d,\mathcal{H}^{N}), if GG does not contain a non-trivial finite normal subgroup and rank(G)=N\operatorname{rank}(G)=N, then X/GX/G is bi-Hölder homeomorphic to an NN-dimensional infranil orbifold 𝒩/Γ\mathcal{N}/\Gamma, where 𝒩\mathcal{N} is a simply connected nilpotent Lie group endowed with a left invariant metric and GG is isomorphic to Γ\Gamma.

Furthermore, if XX is a smooth Riemannian manifold, then the Riemannian orbifold X/GX/G is diffeomorphic to an NN-dimensional infranil orbifold.

First note that Theorem 2.32 and Lemma 4.3 will imply that for a small ϵ\epsilon, the group GG in Theorem 4.5 is isomorphic to an almost-crystallographic group of dimension NN.

Assume that Theorem 4.5 does not hold. Then after rescaling on metrics, there is a sequence of (ϵi,Gi)(\epsilon_{i},G_{i})-homogeneous RCD(ϵi,N)(-\epsilon_{i},N) spaces (Xi,di,N)(X_{i},d_{i},\mathcal{H}^{N}) with ϵi0\epsilon_{i}\to 0, such that any GiG_{i} is isomorphic to an almost-crystallographic group of dimension NN and Xi/GiX_{i}/G_{i} is not bi-Hölder homeomorphic to any infranil orbifold of the form 𝒩i/Gi\mathcal{N}_{i}/G_{i}.

Let GiG_{i}^{\prime} be the bounded index normal subgroups of GiG_{i} in Lemma 2.35. Then rank(Gi)=N\operatorname{rank}(G_{i}^{\prime})=N and it follows from [68, Lemma 2.6] that diam(Xi/Gi)0{\text{diam}}(X_{i}/G_{i}^{\prime})\to 0. Due to Remark 2.40 and Corollary 3.12, we have the following diagram:

(Xi,pi,Gi)\textstyle{(X_{i},p_{i},G_{i}^{\prime})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}eqGH\scriptstyle{eqGH}(N,0,G)\textstyle{(\mathbb{R}^{N},0,G)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Xi/Gi\textstyle{X_{i}/G_{i}^{\prime}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}GH\scriptstyle{GH} pt .\textstyle{\text{ pt }.}

By Lemma 2.36, GG acts freely and transitively on N\mathbb{R}^{N} and hence, G=NG=\mathbb{R}^{N}.

By Theorem 2.12, the groups GiG_{i} admit on non-trivial small subgroups. Then by Lemma 2.39 and Remark 2.40, GiG_{i}^{\prime} are torsion free nilpotent groups. Let 𝒩i\mathcal{N}_{i} be the Mal’cev completion of GiG_{i}^{\prime}. It follows from Proposition 2.33 that GiG_{i} is an almost-crystallographic group modeled on 𝒩i\mathcal{N}_{i} for each ii.

Our goal is to find a left invariant metric on 𝒩i\mathcal{N}_{i} so that Xi/GiX_{i}/G_{i} is bi-Hölder homeomorphic to 𝒩i/Gi\mathcal{N}_{i}/G_{i}. The proof is essentially the same as in [66]. For the convenience of readers, we give the construction of the left invariant metric on 𝒩i\mathcal{N}_{i} in [66, Lemma 4.5].

Lemma 4.6 (66).

Let (Xi,di,pi,Gi)(X_{i},d_{i},p_{i},G_{i}^{\prime}) and 𝒩i\mathcal{N}_{i} be as above. For any ϵ(0,1)\epsilon\in(0,1) and large ii, 𝒩i\mathcal{N}_{i} admits a left invariant metric g𝒩ig_{\mathcal{N}_{i}} with inj𝒩i1ϵ\mathrm{inj}_{\mathcal{N}_{i}}\geq\frac{1}{\epsilon}. Moreover, there is ϵi0\epsilon_{i}\to 0 so that g𝒩i\forall g\in\mathcal{N}_{i}, B1ϵ(g)𝒩iB_{\frac{1}{\epsilon}}(g)\subset\mathcal{N}_{i} is ϵi\epsilon_{i}-C4C^{4}-close, by expg1\mathrm{exp}_{g}^{-1}, to the 1ϵ\frac{1}{\epsilon}-ball in Tg𝒩iT_{g}\mathcal{N}_{i} with the flat metric.

Proof.

Note that the groups GiG_{i}^{\prime} admit no non-trivial small subgroups and rank(Gi)=N\operatorname{rank}(G_{i}^{\prime})=N. Then by Lemma 2.39 and Remark 2.40, for ii large enough there are generators u1,i,,uN,iGiu_{1,i},\ldots,u_{N,i}\in G_{i}^{\prime}, and C1,i,,CN,i+C_{1,i},\ldots,C_{N,i}\in\mathbb{R}^{+} with the following properties:

  1. (1)

    There are polynomials Qi:N×NNQ_{i}:\mathbb{R}^{N}\times\mathbb{R}^{N}\to\mathbb{R}^{N} of degree d(N)\leq d(N) giving the group structures on N\mathbb{R}^{N} by x1x2=Qi(x1,x2)x_{1}\cdot x_{2}=Q_{i}(x_{1},x_{2}) such that for each ii, GiG_{i}^{\prime} is isomorphic to the group (,NQi|×NN)({}^{N},Q_{i}|_{{}^{N}\times{}^{N}}) and the group (N,Qi)(\mathbb{R}^{N},Q_{i}) is isomorphic to 𝒩i\mathcal{N}_{i}.

  2. (2)

    There is C>0C>0 such that the set

    Pi:=P(u1,i,,uN,i;C1,i,,CN,i)GiP_{i}:=P(u_{1,i},\ldots,u_{N,i};C_{1,i},\ldots,C_{N,i})\subset G_{i}^{\prime}

    is a nilprogression in CC-normal form with thick(Pi)\text{thick}(P_{i})\to\infty.

  3. (3)

    For each ε>0\varepsilon>0 there is δ>0\delta>0 such that

    (4.1) G(δPi){\displaystyle G(\delta P_{i})\subset\{ gGi|di(gpi,pi)ε},\displaystyle g\in G_{i}^{\prime}\,|\,d_{i}(gp_{i},p_{i})\leq\varepsilon\},

    for ii large enough.

By (4.1), there is δ1>0\delta_{1}>0 with G(δ1Pi){gGi|di(gpi,pi)1}G(\delta_{1}P_{i})\subset\{g\in G_{i}^{\prime}|d_{i}(gp_{i},p_{i})\leq 1\}. Hence there is an integer DD\in\mathbb{N} so that

(4.2) G(Pi)G(δ1Pi)D{gGi|di(gpi,pi)D}.G(P_{i})\subset G(\delta_{1}P_{i})^{D}\subset\{g\in G_{i}^{\prime}|d_{i}(gp_{i},p_{i})\leq D\}.

Let gj,i:=uj,iCj,iCg_{j,i}:=u_{j,i}^{\lfloor\frac{C_{j,i}}{C}\rfloor} and vj,i:=log(gj,i)Te𝒩iv_{j,i}:=\log(g_{j,i})\in T_{e}\mathcal{N}_{i}. Then {v1,i,,vN,i}\{v_{1,i},\ldots,v_{N,i}\} is a strong Mal’cev basis of the Lie algebra Te𝒩iT_{e}\mathcal{N}_{i} (see [68]).

Notice that the groups GiG_{i}^{\prime} converge equivariantly to the group of translations in N\mathbb{R}^{N}. By (4.2), after passing to a subsequence, for each j{1,,N}j\in\{1,\ldots,N\} we can assume gj,ig_{j,i} converges equivariantly to some vjNv_{j}\in\mathbb{R}^{N}. We may identify N\mathbb{R}^{N} with its Lie algebra and {v1,,vN}\{v_{1},\ldots,v_{N}\} is a basis of N\mathbb{R}^{N}.

Define the left invariant metric g𝒩ig_{\mathcal{N}_{i}} by the inner product on Te𝒩iT_{e}\mathcal{N}_{i} as following:

g𝒩i(vj1,i,vj2,i)=vj1,vj2,g_{\mathcal{N}_{i}}(v_{j_{1},i},v_{j_{2},i})=\langle v_{j_{1}},v_{j_{2}}\rangle,

where 1j1,j2N1\leq j_{1},j_{2}\leq N and the right-hand side is the inner product in N\mathbb{R}^{N}.

Since {vj,i,1jN}\{v_{j,i},1\leq j\leq N\} is a strong Malcev basis of Te𝒩iT_{e}\mathcal{N}_{i}, for any 1j1<j2N1\leq j_{1}<j_{2}\leq N,

(4.3) [vj1,i,vj2,i]=j=j2+1naj1j2,ijvj,i.\displaystyle[v_{j_{1},i},v_{j_{2},i}]=\sum_{j=j_{2}+1}^{n}a_{j_{1}j_{2},i}^{j}\ v_{j,i}.

It is proven in [68, Lemma 2.64 and Proposition 8.2] that the structure coefficients of Te𝒩iT_{e}\mathcal{N}_{i} with respect to the basis {v1,i,,vN,i}\{v_{1,i},\ldots,v_{N,i}\} converge to the structure coefficients in N\mathbb{R}^{N} with respect to {v1,,vN}\{v_{1},\ldots,v_{N}\} as ii\to\infty. Thus aj1j2,ij0a_{j_{1}j_{2},i}^{j}\to 0 as ii\to\infty, since the limit group is abelian. Define aj1j2,ij=0a_{j_{1}j_{2},i}^{j}=0 if jj1j\leq j_{1} or jj2j\leq j_{2}. Then for any 1j1,j2,j3N1\leq j_{1},j_{2},j_{3}\leq N,

g𝒩i(vj1,ivj2,i,vj3,i)=12(aj1j2,ij3aj2j3,ij1+aj3j1,ij2).g_{\mathcal{N}_{i}}(\nabla_{v_{j_{1},i}}v_{j_{2},i},v_{j_{3},i})=\frac{1}{2}(a_{j_{1}j_{2},i}^{j_{3}}-a_{j_{2}j_{3},i}^{j_{1}}+a_{j_{3}j_{1},i}^{j_{2}}).

Note that all terms on the right-hand side are constant (depending on ii) and converge to 0 as ii\to\infty. In particular, the covariant derivatives of the Riemannian curvature tensor g𝒩ig_{\mathcal{N}_{i}} satisfy

|(g𝒩i)kRmg𝒩i|ϵi, 0k3,|(\nabla^{g_{\mathcal{N}_{i}}})^{k}Rm_{g_{\mathcal{N}_{i}}}|\leq\epsilon_{i},\ 0\leq k\leq 3,

where ϵi0\epsilon_{i}\to 0. Then one can easily verify that this metric fulfills the conditions. ∎

Remark 4.7.

In [66], Wang used the results in [8, 68] to embed GiG_{i}^{\prime} as a lattice in a simply connected nilpotent Lie group 𝒩i\mathcal{N}_{i}. In fact, since GiG_{i}^{\prime} is torsion free nilpotent, one can directly use its Mal’cev completion. One can see Remark 4.11 for the reason why the groups GiG_{i} and GiG_{i}^{\prime} are torsion free in Wang’s theorem.

From now on, 𝒩i\mathcal{N}_{i} is always endowed with the metric g𝒩ig_{\mathcal{N}_{i}} constructed in Lemma 4.6.

Define Gi(pi,D):={gGi|di(gpi,pi)D}G_{i}^{\prime}(p_{i},D):=\{g\in G_{i}^{\prime}|d_{i}(gp_{i},p_{i})\leq D\}. Assume that a pseudo-group GG acts on two metric spaces X1,X2X_{1},X_{2} separately by isometries. Following [66], we say a map h:X1X2h:X_{1}\to X_{2} is ϵ\epsilon-almost GG-equivariant if d(h(gx),gh(x))<ϵd(h(gx),gh(x))<\epsilon for any xX1,gGx\in X_{1},g\in G.

The following two lemmas come from [66, 65].

Lemma 4.8 (66).

For any ϵ>0\epsilon>0, let B1ϵ(pi)XiB_{\frac{1}{\epsilon}}(p_{i})\subset X_{i} and B1ϵ(e)𝒩iB_{\frac{1}{\epsilon}}(e)\subset\mathcal{N}_{i}. Then there exists an ϵi\epsilon_{i}-GHA hi:B1ϵ(pi)B1ϵ(e)h_{i}^{\prime}:B_{\frac{1}{\epsilon}}(p_{i})\to B_{\frac{1}{\epsilon}}(e) which is ϵi\epsilon_{i}-almost Gi(pi,1ϵ)G_{i}^{\prime}(p_{i},{\frac{1}{\epsilon}})-equivariant if it is well-defined, where ϵi0\epsilon_{i}\to 0 as ii\to\infty.

Lemma 4.9 (65, 66).

The map hih_{i}^{\prime} in Lemma 4.8 can be extended to a global map hi:Xi𝒩ih_{i}:X_{i}\to\mathcal{N}_{i}, which is an ϵi\epsilon_{i}-GHA on any 1ϵ\frac{1}{\epsilon}-ball and ϵi\epsilon_{i}-almost GiG_{i}^{\prime}-equivariant with ϵi0\epsilon_{i}\to 0.

Now we can follow the arguments in [66] to complete the proof of Theorem 4.5.

Proof of Theorem 4.5.

Let us argue by contradiction. There is a sequence of (ϵi,Gi)(\epsilon_{i},G_{i})-homogeneous RCD(ϵi,N)(-\epsilon_{i},N) spaces (Xi,di,N)(X_{i},d_{i},\mathcal{H}^{N}) with ϵi0\epsilon_{i}\to 0 and rank(Gi)=N\operatorname{rank}(G_{i})=N. Also, we have already established the following diagram:

(Xi,pi,Gi)\textstyle{(X_{i},p_{i},G_{i}^{\prime})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}eqGH\scriptstyle{eqGH}(N,0,N)\textstyle{(\mathbb{R}^{N},0,\mathbb{R}^{N})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Xi/Gi\textstyle{X_{i}/G_{i}^{\prime}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}GH\scriptstyle{GH} pt ,\textstyle{\text{ pt },}

where for each ii, GiG_{i}^{\prime} is a normal subgroup of bounded index in GiG_{i}, embedded as a lattice in an NN-dimensional simply connected nilpotent Lie group 𝒩i\mathcal{N}_{i} and GiG_{i} is an almost-crystallographic group modeled on 𝒩i\mathcal{N}_{i}. We also assumed that none of Xi/GiX_{i}/G_{i} is bi-Hölder homeomorphic to the infranil orbifold 𝒩i/Gi\mathcal{N}_{i}/G_{i}.

By the construction of the metric g𝒩ig_{\mathcal{N}_{i}} in Lemma 4.6, the lattice GiG_{i}^{\prime} is ϵi\epsilon_{i}-dense in 𝒩i\mathcal{N}_{i}. Since diam(Xi/Gi)0{\text{diam}}(X_{i}/G_{i}^{\prime})\to 0, the map hih_{i} in Lemma 4.9 is also ϵi\epsilon_{i}-almost GiG_{i}-equivariant.

By the same arguments in the proof of [66, Theorem A], we can assume that GiG_{i} acts on 𝒩i\mathcal{N}_{i} by isometries and for any small ϵ>0\epsilon>0, there is a normal subgroup Gi′′G_{i}^{\prime\prime} in GiG_{i}^{\prime} of finite index, which is also normal in GiG_{i}, so that Gi′′B1ϵ(e)={e}G_{i}^{\prime\prime}\cap B_{\frac{1}{\epsilon}}(e)=\{e\}.

Since Gi′′B1ϵ(e)={e}G_{i}^{\prime\prime}\cap B_{\frac{1}{\epsilon}}(e)=\{e\}, we can apply Lemma 4.6 to conclude that the injective radius of 𝒩i/Gi′′\mathcal{N}_{i}/G_{i}^{\prime\prime} is at least 1ϵ{\frac{1}{\epsilon}}. For any y𝒩i/Gi′′y\in\mathcal{N}_{i}/G_{i}^{\prime\prime}, B1ϵ(y)𝒩i/Gi′′B_{\frac{1}{\epsilon}}(y)\subset\mathcal{N}_{i}/G_{i}^{\prime\prime} is ϵi\epsilon_{i}-C4C^{4}-close to the 1ϵ\frac{1}{\epsilon}-ball in the tangent space Ty(𝒩i/Gi′′)T_{y}(\mathcal{N}_{i}/G_{i}^{\prime\prime}) with the flat metric.

Since hih_{i} is ϵi\epsilon_{i}-almost GiG_{i}-equivariant, we can reduce hih_{i} to a map

h¯i:Xi/Gi′′𝒩i/Gi′′,\bar{h}_{i}:X_{i}/G_{i}^{\prime\prime}\to\mathcal{N}_{i}/G_{i}^{\prime\prime},

which is an ϵi\epsilon_{i}-GHA on any 1ϵ{\frac{1}{\epsilon}}-ball and ϵi\epsilon_{i}-almost Gi/Gi′′G_{i}/G_{i}^{\prime\prime}-equivariant.

Since Gi/Gi′′G_{i}/G_{i}^{\prime\prime} is finite, we can apply [66, Theorem 3.5] to

h¯i:(Xi/Gi′′,Gi/Gi′′)(𝒩i/Gi′′,Gi/Gi′′).\bar{h}_{i}:(X_{i}/G_{i}^{\prime\prime},G_{i}/G_{i}^{\prime\prime})\longrightarrow(\mathcal{N}_{i}/G_{i}^{\prime\prime},G_{i}/G_{i}^{\prime\prime}).

Thus there is a (Gi/Gi′′)(G_{i}/G_{i}^{\prime\prime})-equivariant map fGi/Gi′′:Xi/Gi′′𝒩i/Gi′′f_{G_{i}/G_{i}^{\prime\prime}}:X_{i}/G_{i}^{\prime\prime}\to\mathcal{N}_{i}/G_{i}^{\prime\prime}, which is harmonic (N,Φ(ϵ|N))(N,\Phi(\epsilon|N))-splitting on any 15ϵ\frac{1}{5\epsilon}-ball. Then by Theorem 2.20,

(1Φ(ϵ|N))di(x,y)1+Φ(ϵ|N)d(fGi/Gi′′(x),fGi/Gi′′(y))(1+Φ(ϵ|N))di(x,y),(1-\Phi(\epsilon|N))d_{i}(x,y)^{1+\Phi(\epsilon|N)}\leq d(f_{G_{i}/G_{i}^{\prime\prime}}(x),f_{G_{i}/G_{i}^{\prime\prime}}(y))\leq(1+\Phi(\epsilon|N))d_{i}(x,y),

for any x,yXi/Gi′′x,y\in X_{i}/G_{i}^{\prime\prime} with di(x,y)110ϵd_{i}(x,y)\leq\frac{1}{10\epsilon}.

Since fGi/Gi′′f_{G_{i}/G_{i}^{\prime\prime}} is (Gi/Gi′′)(G_{i}/G_{i}^{\prime\prime})-equivariant, it can be reduced to a bi-Hölder map on the quotient space f:Xi/Gi𝒩i/Gif:X_{i}/G_{i}\to\mathcal{N}_{i}/G_{i}. This leads to a contradiction to the assumption.

Furthermore, if XiX_{i} is a Riemannian manifold, then Xi/Gi′′X_{i}/G_{i}^{\prime\prime} is also a Riemannian manifold, since the group Gi′′G_{i}^{\prime\prime} acts freely on XiX_{i}. So by Theorem 2.20, the (Gi/Gi′′)(G_{i}/G_{i}^{\prime\prime})-equivariant map fGi/Gi′′:Xi/Gi′′𝒩i/Gi′′f_{G_{i}/G_{i}^{\prime\prime}}:X_{i}/G_{i}^{\prime\prime}\to\mathcal{N}_{i}/G_{i}^{\prime\prime} restricted on any 110ϵ\frac{1}{10\epsilon}-ball, is a diffeomorphism onto its image. Notice that Xi/Gi′′Xi/GiX_{i}/G_{i}^{\prime\prime}\to X_{i}/G_{i} and 𝒩i/Gi′′𝒩i/Gi\mathcal{N}_{i}/G_{i}^{\prime\prime}\to\mathcal{N}_{i}/G_{i} are orbifold coverings. Hence, the reduced map f:Xi/Gi𝒩i/Gif:X_{i}/G_{i}\to\mathcal{N}_{i}/G_{i} is a diffeomorphism between orbifolds. This completes the proof. ∎

Combining Theorem 4.4 and Theorem 4.5, both Theorem 1.8 and Theorem 1.9 are readily obtained.

Remark 4.10.

In the last statement of Theorem 1.9, it is expected that the assumptions requiring a good orbifold and an orbifold fundamental group without non-trivial finite normal subgroups can be eliminated. Due to [21, Proposition 1.4], any almost flat orbifold is an infranil orbifold. So it might be more natural to seek a nearby almost flat metric under the conditions in Theorem 1.9. This is achieved in the manifold case via Ricci flow smoothing techniques (see [35]).

Remark 4.11.

Although in Theorem 1.8, we assume that the group GG does not contain a non-trivial finite normal subgroup, Theorem 1.2 is still a corollary of Theorem 1.8. This is due to the fact that if YY is a closed aspherical topological manifold, then π1(Y)\pi_{1}(Y) is torsion free. Notice that any closed topological manifold is homotopy equivalent to a CW complex [43] and the fundamental group of an aspherical finite-dimensional CW complex is torsion free [48].

5. Rigidity and regularity of almost homogeneous Einstein metrics

In this section, we mainly focus on the rigidity and ϵ\epsilon-regularity for almost homogeneous Riemannian orbifolds and manifolds with bounded Ricci curvature. We first give the proof of Theorem 1.10, which is an orbifold verion of [59, Theorem 0.2].

Proof of Theorem 1.10.

Since (1)\Rightarrow(2)\Rightarrow(3) is trivial and (3) is equivalent to (4) by Theorem 1.8, it suffices to show that (3) and (4) together imply (1).

Let us argue by contradiction. Suppose that there is a sequence of non-flat Einstein nn-orbifolds (𝒪i,gi)(\mathcal{O}_{i},g_{i}) such that Ricgiλi\operatorname{Ric}_{g_{i}}\equiv\lambda_{i} with λi(n1)\lambda_{i}\geq-(n-1), diam(𝒪i,gi)0\operatorname{diam}(\mathcal{O}_{i},g_{i})\to 0, and satisfying (3) and (4) in Theorem 1.10.

Consider the universal orbifold covers (𝒪~i,g~i)(\tilde{\mathcal{O}}_{i},\tilde{g}_{i}). Due to Corollary 3.8 and Corollary 3.10, we can assume that λi<0\lambda_{i}<0. Then up to a rescaling on metrics, we can further assume that λi=(n1)\lambda_{i}=-(n-1). Note that 𝒪i\mathcal{O}_{i} still converges to a point and rank(π1orb(𝒪i))=n\operatorname{rank}(\pi_{1}^{orb}(\mathcal{O}_{i}))=n. By Theorem 1.8, volg~i(B1(x~i))v(n)>0\operatorname{vol}_{\tilde{g}_{i}}(B_{1}(\tilde{x}_{i}))\geq v^{\prime}(n)>0 for some x~i𝒪~i\tilde{x}_{i}\in\tilde{\mathcal{O}}_{i}. Up to a subsequence, we have the following pmGH-convergence by Theorem 2.18,

(𝒪~i,g~i,volg~i,x~i)pmGH(X~,g~,n,x~),(\tilde{\mathcal{O}}_{i},\tilde{g}_{i},\operatorname{vol}_{\tilde{g}_{i}},\tilde{x}_{i})\xrightarrow{pmGH}(\tilde{X},\tilde{g},\mathcal{H}^{n},\tilde{x}),

where by Proposition 3.6, X~\tilde{X} is isometric to a simply connected nilpotent Lie group with a left invariant Riemannian metric, denoted by g~\tilde{g}.

On the other hand, for any ii, (|𝒪~i|reg,gi,reg)(|\tilde{\mathcal{O}}_{i}|_{reg},g_{i,reg}) is a smooth open Riemannian manifold with Ric(n1)\text{Ric}\equiv-(n-1) and volg~i(Br(y~i)|𝒪~i|reg)=volg~i(Br(y~i))\operatorname{vol}_{\tilde{g}_{i}}(B_{r}(\tilde{y}_{i})\cap|\tilde{\mathcal{O}}_{i}|_{reg})=\operatorname{vol}_{\tilde{g}_{i}}(B_{r}(\tilde{y}_{i})) for any y~i𝒪~i\tilde{y}_{i}\in\tilde{\mathcal{O}}_{i} and r>0r>0. By the standard Schauder estimate, (|𝒪~i|reg,gi,reg)(|\tilde{\mathcal{O}}_{i}|_{reg},g_{i,reg}) converges in the ClocC^{\infty}_{loc}-norm to a full measure subset of (X~,g~)(\tilde{X},\tilde{g}) (see [13]). Since X~\tilde{X} is a Riemannian manifold, RicX~(n1)\text{Ric}_{\tilde{X}}\equiv-(n-1).

By [50, Theorem 2.4], any left invariant Riemannian metric of a nilpotent but not abelian Lie group has both directions of strictly negative and positve Ricci curvature. Thus, (X~,g~)(\tilde{X},\tilde{g}) must be isometric to n\mathbb{R}^{n}, which leads to a contradiction. ∎

Recall that if (M,g)(M,g) is a Riemannian manifold and GG is a discrete subgroup of Iso(M)\operatorname{Iso}(M), then M/GM/G admits a natural orbifold structure. Let M~\tilde{M} be the universal cover of MM. Then M~\tilde{M} is also the universal orbifold cover of the good orbifold M/GM/G. Moreover, if MM is simply connected, then π1orb(M/G)=G\pi_{1}^{orb}(M/G)=G. Therefore, the following corollary is readily derived from Theorem 1.8 and Theorem 1.10.

Corollary 5.1.

There is ϵ=ϵ(n)>0,v=v(n)>0\epsilon=\epsilon(n)>0,v=v(n)>0 such that if an (ϵ,G)(\epsilon,G)-homogeneous Einstein nn-manifold (M,g)(M,g) satisfies Ricg=λg\operatorname{Ric}_{g}=\lambda g with λ(n1)\lambda\geq-(n-1), then the followings are equivalent:

  1. (1)

    vol(B1(x))v\operatorname{vol}(B_{1}(x))\geq v for some xMx\in M, and MM is simply connected;

  2. (2)

    rank(G)=n\operatorname{rank}(G)=n;

  3. (3)

    MM is diffeomorphic to n\mathbb{R}^{n};

  4. (4)

    MM is isometric to n\mathbb{R}^{n}.

In particular, if we only assume vol(B1(x))v\operatorname{vol}(B_{1}(x))\geq v for some xMx\in M, then MM is flat.

The following proposition is a quantitative rigidity version of Corollary 5.1.

Proposition 5.2.

Given v>0v>0 and p(1,)p\in(1,\infty), for any δ>0\delta>0, there is ϵ=ϵ(n,v,p,δ)\epsilon=\epsilon(n,v,p,\delta) such that if an (ϵ,G)(\epsilon,G)-homogeneous nn-manifold (M,g)(M,g) satisfies |Ricλg|ϵ|\operatorname{Ric}-\lambda g|\leq\epsilon with λ(n1)\lambda\geq-(n-1) and vol(B1(x))v\operatorname{vol}(B_{1}(x))\geq v, then B1(x)|Rm|pδ\int_{B_{1}(x)}|Rm|^{p}\leq\delta.

Proof.

Argue by contradiction. Suppose that there exists δ0>0\delta_{0}>0 such that for any ϵi0\epsilon_{i}\to 0, there is a sequence of (ϵi,Gi)(\epsilon_{i},G_{i})-homogeneous pointed nn-manifolds (Mi,gi,xi)(M_{i},g_{i},x_{i}) satisfying |Ricλigi|ϵi|\operatorname{Ric}-\lambda_{i}g_{i}|\leq\epsilon_{i} with λi(n1)\lambda_{i}\geq-(n-1), volgi(B1(xi))v\operatorname{vol_{g_{i}}}(B_{1}(x_{i}))\geq v and B1(xi)|Rm|p>δ0\int_{B_{1}(x_{i})}|Rm|^{p}>\delta_{0}.

By Corollary 3.8, we can assume that λi\lambda_{i} converges to some λ0\lambda_{\infty}\leq 0. Up to a subsequence, we have the following pGH-convergence

(Mi,gi,xi)pGH(X,d,x),(M_{i},g_{i},x_{i})\xrightarrow{pGH}(X,d,x),

where (X,d)(X,d) is isometric to a Riemannian manifold with Ricλ\text{Ric}\geq\lambda_{\infty} by Theorem 1.4. Note that (Mi,gi,xi)(M_{i},g_{i},x_{i}) converges in the pointed C1,αW2,qC^{1,\alpha}\cap W^{2,q}-topology to (X,gX,x)(X,g_{X},x), where the metric gXg_{X} is a weak solution of the Einstein equation

Δg+Q(g,g)=λg,\Delta g+Q(g,\partial g)=\lambda_{\infty}g,

under harmonic coordinate charts (see [4, 13]). Hence, gXg_{X} is a smooth metric and XX is an Einstein manifold with RicgX=λ\text{Ric}_{g_{X}}=\lambda_{\infty}. By the same arguments in the proof of Theorem 1.10, XX is a flat manifold. Since (Mi,gi,xi)(M_{i},g_{i},x_{i}) converges in the pointed C1,αW2,qC^{1,\alpha}\cap W^{2,q}-topology to (X,gX,x)(X,g_{X},x) for any 0<α<10<\alpha<1 and 1<q<1<q<\infty, we have B1(xi)|Rm|p0\int_{B_{1}(x_{i})}|Rm|^{p}\to 0. This leads to a contradiction. ∎

Let (M,g)(M,g) be a Riemannian manifold. Recall that the CkC^{k}-harmonic radius at xMx\in M is defined to be the largest r>0r>0 such that there exists a harmonic coordinate system on Br(x)B_{r}(x) with CkC^{k}-control on the metric tensor. Harmonic coordinates have an abundancce of good properties when it comes to regularity issues. We refer to [54] for a nice introduction. In particular, if the Ricci curvature is uniformly bounded, then in harmonic coordinates, the metric gijg_{ij} has a priori C1,αW2,qC^{1,\alpha}\cap W^{2,q}-bounds for any α(0,1)\alpha\in(0,1) and q(1,)q\in(1,\infty).

Let us now proceed to prove the ϵ\epsilon-regularity theorem (Theorem 1.12).

Proof of Theorem 1.12.

Let us argue by contradiction. Suppose that there exists a sequence of almost homogeneous pointed nn-orbifolds (𝒪i,gi,xi)(\mathcal{O}_{i},g_{i},x_{i}) satisfying |Ricgi|n1|\text{Ric}_{g_{i}}|\leq n-1, vol(B1(xi))v\operatorname{vol}(B_{1}(x_{i}))\geq v and B1(xi)|Rm|p\int_{B_{1}(x_{i})}|Rm|^{p}\to\infty. Then up to a subsequence, (𝒪i,gi,xi)(\mathcal{O}_{i},g_{i},x_{i}) converges in the pGH-sense to (X,d,p)(X,d,p). By Theorem 1.4, XX is a Riemannian manifold. Hence, B1(xi)|𝒪i|regB_{1}(x_{i})\cap|\mathcal{O}_{i}|_{reg} converges to an open subset of (X,d,p)(X,d,p) in the C1,αW2,qC^{1,\alpha}\cap W^{2,q}-norm for any α(0,1)\alpha\in(0,1) and q(1,)q\in(1,\infty), which implies that B1(xi)|Rm|p=B1(xi)|𝒪i|reg|Rm|p\int_{B_{1}(x_{i})}|Rm|^{p}=\int_{B_{1}(x_{i})\cap|\mathcal{O}_{i}|_{reg}}|Rm|^{p} is bounded. This leads to a contradiction.

For the last statement, notice that the C1C^{1}-harmonic radius rhr_{h} is continuous under C1,αC^{1,\alpha}-topology and hence, a similar proof applies. ∎

The proof of Theorem 1.13 is readily obtained.

Proof of Theorem 1.13.

By [17, Corollary 1.2], (1) implies (2). It follows from Theorem 1.1 that (2), (3) and (4) are equivalent. Then Theorem 1.12 shows that (4) implies (1). One can adjust the constants to make (1), (2), (3) and (4) equivalent. ∎

References

  • [1] Luigi Ambrosio, Nicola Gigli, and Giuseppe Savaré. Calculus and heat flow in metric measure spaces and applications to spaces with Ricci bounds from below. Inventiones mathematicae, 195(2):289–391, 2014.
  • [2] Luigi Ambrosio, Nicola Gigli, and Giuseppe Savaré. Metric measure spaces with riemannian Ricci curvature bounded from below. Duke Mathematical Journal, 163(7):1405–1490, 2014.
  • [3] Luigi Ambrosio, Shouhei Honda, and David Tewodrose. Short-time behavior of the heat kernel and weyl’s law on RCD*(K, N) spaces. Annals of Global Analysis and Geometry, 53:97–119, 2018.
  • [4] Michael T Anderson. Convergence and rigidity of manifolds under Ricci curvature bounds. Inventiones mathematicae, 102(1):429–445, 1990.
  • [5] Louis Auslander. Bieberbach’s theorems on space groups and discrete uniform subgroups of lie groups. Annals of Mathematics, 71(3):579–590, 1960.
  • [6] Itai Benjamini, Hilary Finucane, and Romain Tessera. On the scaling limit of finite vertex transitive graphs with large diameter. Combinatorica, 37:333–374, 2017.
  • [7] Camillo Brena, Nicola Gigli, Shouhei Honda, and Xingyu Zhu. Weakly non-collapsed RCD spaces are strongly non-collapsed. Journal für die reine und angewandte Mathematik (Crelles Journal), 2023(794):215–252, 2023.
  • [8] Emmanuel Breuillard, Ben Green, and Terence Tao. The structure of approximate groups. Publications mathématiques de l’IHÉS, 116:115–221, 2012.
  • [9] Elia Bruè, Aaron Naber, and Daniele Semola. Boundary regularity and stability for spaces with Ricci bounded below. Inventiones mathematicae, 228(2):777–891, 2022.
  • [10] Elia Brué and Daniele Semola. Constancy of the dimension for RCD(K, N) spaces via regularity of lagrangian flows. Communications on Pure and Applied Mathematics, 73(6):1141–1204, 2020.
  • [11] P. Buser and H. Karcher. Gromov’s Almost Flat Manifolds. Société mathématique de France, 1981.
  • [12] Fabio Cavalletti and Emanuel Milman. The globalization theorem for the curvature-dimension condition. Inventiones mathematicae, 226(1):1–137, 2021.
  • [13] Jeff Cheeger and Tobias H Colding. On the structure of spaces with Ricci curvature bounded below. I. Journal of Differential Geometry, 46(3):406–480, 1997.
  • [14] Jeff Cheeger, Kenji Fukaya, and Mikhael Gromov. Nilpotent structures and invariant metrics on collapsed manifolds. Journal of the American Mathematical Society, 5(2):327–372, 1992.
  • [15] Jeff Cheeger and Detlef Gromoll. The splitting theorem for manifolds of nonnegative ricci curvature. Journal of Differential Geometry, 6(1):119–128, 1971.
  • [16] Jeff Cheeger, Wenshuai Jiang, and Aaron Naber. Rectifiability of singular sets of noncollapsed limit spaces with Ricci curvature bounded below. Annals of Mathematics, 193(2):407–538, 2021.
  • [17] Xianzhe Dai, Peter Petersen, and Guofang Wei. Integral pinching theorems. manuscripta mathematica, 101:143–152, 2000.
  • [18] Guido De Philippis and Nicola Gigli. Non-collapsed spaces with ricci curvature bounded from below. Journal de l’École polytechnique—Mathématiques, 5:613–650, 2018.
  • [19] Karel Dekimpe. A users’ guide to infra-nilmanifolds and almost-bieberbach groups. arXiv preprint arXiv:1603.07654, 2016.
  • [20] Qin Deng, Jaime Santos-Rodríguez, Sergio Zamora, and Xinrui Zhao. Margulis lemma on RCD(K,N) spaces. arXiv preprint arXiv:2308.15215, 2023.
  • [21] Yu Ding. A restriction for singularities on collapsing orbifolds. International Scholarly Research Notices, 2011(1):502814, 2011.
  • [22] Matthias Erbar, Kazumasa Kuwada, and Karl-Theodor Sturm. On the equivalence of the entropic curvature-dimension condition and Bochner’s inequality on metric measure spaces. Inventiones mathematicae, 201(3):993–1071, 2015.
  • [23] Kenji Fukaya. Collapsing riemannian manifolds to ones of lower dimensions. Journal of differential geometry, 25(1):139–156, 1987.
  • [24] Kenji Fukaya and Takao Yamaguchi. The fundamental groups of almost nonnegatively curved manifolds. Annals of mathematics, 136(2):253–333, 1992.
  • [25] Fernando Galaz-Garcia, Martin Kell, Andrea Mondino, and Gerardo Sosa. On quotients of spaces with Ricci curvature bounded below. Journal of Functional Analysis, 275(6):1368–1446, 2018.
  • [26] Tsachik Gelander. Limits of finite homogeneous metric spaces. L’Enseignement Mathématique, 59(1):195–206, 2013.
  • [27] Nicola Gigli. An overview of the proof of the splitting theorem in spaces with non-negative Ricci curvature. Analysis and Geometry in Metric Spaces, 2(1), 2014.
  • [28] Nicola Gigli. On the differential structure of metric measure spaces and applications, volume 236. American Mathematical Society, 2015.
  • [29] Nicola Gigli, Andrea Mondino, and Giuseppe Savaré. Convergence of pointed non-compact metric measure spaces and stability of Ricci curvature bounds and heat flows. Proceedings of the London Mathematical Society, 111(5):1071–1129, 2015.
  • [30] Mikhail Gromov. Almost flat manifolds. Journal of Differential Geometry, 13(2):231–241, 1978.
  • [31] André Haefliger and Quach Ngoc Du. Appendice: une présentation du groupe fondamental d’une orbifold. Astérisque, 116:98–107, 1984.
  • [32] Shouhei Honda and Artem Nepechiy. Locally homogeneous rcd spaces. arXiv preprint arXiv:2404.02390, 2024.
  • [33] Shouhei Honda and Yuanlin Peng. A note on the topological stability theorem from rcd spaces to riemannian manifolds. manuscripta mathematica, 172(3-4):971–1007, 2023.
  • [34] Hongzhi Huang. Fibrations and stability for compact group actions on manifolds with local bounded ricci covering geometry. Frontiers of Mathematics in China, 15(1):69–89, 2020.
  • [35] Hongzhi Huang, Lingling Kong, Xiaochun Rong, and Shicheng Xu. Collapsed manifolds with ricci bounded covering geometry. Transactions of the American Mathematical Society, 373(11):8039–8057, 2020.
  • [36] Shaosai Huang, Xiaochun Rong, Bing Wang, et al. Collapsing geometry with ricci curvature bounded below and ricci flow smoothing. SIGMA. Symmetry, Integrability and Geometry: Methods and Applications, 16:123, 2020.
  • [37] Yonghong Huang and Shanzhong Sun. Non-embedding theorems of nilpotent lie groups and sub-riemannian manifolds. Frontiers of Mathematics in China, 15:91–114, 2020.
  • [38] Vitali Kapovitch. Mixed curvature almost flat manifolds. Geometry & Topology, 25(4):2017–2059, 2021.
  • [39] Vitali Kapovitch and Andrea Mondino. On the topology and the boundary of N–dimensional RCD (K,N) spaces. Geometry & Topology, 25(1):445–495, 2021.
  • [40] Vitali Kapovitch and Burkhard Wilking. Structure of fundamental groups of manifolds with Ricci curvature bounded below. arXiv preprint arXiv:1105.5955, 2011.
  • [41] Mikhail Ivanovich Kargapolov and Jurij Ivanovič Merzljakov. Fundamentals of the Theory of Groups, volume 62. Springer, 1979.
  • [42] Martin Kell. Transport maps, non-branching sets of geodesics and measure rigidity. Advances in Mathematics, 320:520–573, 2017.
  • [43] Robion C Kirby, Laurence C Siebenmann, et al. On the triangulation of manifolds and the hauptvermutung. Bull. Amer. Math. Soc, 75(4):742–749, 1969.
  • [44] Yu Kitabeppu. A sufficient condition to a regular set being of positive measure on spaces. Potential analysis, 51(2):179–196, 2019.
  • [45] Bruce Kleiner and John Lott. Geometrization of three-dimensional orbifolds via ricci flow. arXiv preprint arXiv:1101.3733, 2011.
  • [46] Zhenhao Li. The globalization theorem for CD(K, N) on locally finite spaces. Annali di Matematica Pura ed Applicata (1923-), 203(1):49–70, 2024.
  • [47] John Lott and Cédric Villani. Ricci curvature for metric-measure spaces via optimal transport. Annals of Mathematics, pages 903–991, 2009.
  • [48] Wolfgang Lück. Aspherical manifolds. Bulletin of the Manifold Atlas, 1(17):5–6, 2012.
  • [49] Anatoliĭ Ivanovich Mal’cev. On a class of homogeneous spaces. Izv. Akad. Nauk SSSR, Ser. Mat., 13:9–32, 1949.
  • [50] John Milnor. Curvatures of left invariant metrics on lie groups. Advances in Mathematics, 21(3):293–329, 1976.
  • [51] Andrea Mondino and Guofang Wei. On the universal cover and the fundamental group of an RCD*(K, N)-space. Journal für die reine und angewandte Mathematik (Crelles Journal), 2019(753):211–237, 2019.
  • [52] Aaron Naber and Ruobing Zhang. Topology and ϵ\epsilon–regularity theorems on collapsed manifolds with ricci curvature bounds. Geometry & Topology, 20(5):2575–2664, 2016.
  • [53] Jiayin Pan and Xiaochun Rong. Ricci curvature and isometric actions with scaling nonvanishing property. arXiv preprint arXiv:1808.02329, 2018.
  • [54] Peter Petersen. Convergence theorems in riemannian geometry. Comparison geometry (Berkeley, CA, 1993–94), 30:167–202, 1997.
  • [55] Xiaochun Rong. Collapsed manifolds with local ricci bounded covering geometry. arXiv preprint arXiv:2211.09998, 2022.
  • [56] Ernst A Ruh. Almost flat manifolds. Journal of Differential Geometry, 17(1):1–14, 1982.
  • [57] Jaime Santos-Rodríguez. Invariant measures and lower Ricci curvature bounds. Potential Analysis, 53(3):871–897, 2020.
  • [58] Jaime Santos-Rodriguez and Sergio Zamora-Barrera. On fundamental groups of rcd spaces. Journal für die reine und angewandte Mathematik (Crelles Journal), 2023(799):249–286, 2023.
  • [59] Cuifang Si and Shicheng Xu. Rigidity for einstein manifolds under bounded covering geometry. arXiv preprint arXiv:2409.15707, 2024.
  • [60] Karl-Theodor Sturm. On the geometry of metric measure spaces. I. Acta Math, 196:65–131, 2006.
  • [61] Karl-Theodor Sturm. On the geometry of metric measure spaces. II. Acta Math, 196:133–177, 2006.
  • [62] William P HG Thurston. Three-Dimensional Geometry and Topology, Volume 1: Volume 1. Princeton university press, 1997.
  • [63] Alan M Turing. Finite approximations to lie groups. Annals of Mathematics, 39(1):105–111, 1938.
  • [64] Cédric Villani. Optimal transport: old and new, volume 338. Springer, 2009.
  • [65] Jikang Wang. On the limit of simply connected manifolds with discrete isometric cocompact group actions. arXiv preprint arXiv:2307.07658, 2023.
  • [66] Jikang Wang. On the non-collapsed RCD spaces with local bounded covering geometry. arXiv preprint arXiv:2412.06131, 2024.
  • [67] Jikang Wang. RCD*(K, N) spaces are semi-locally simply-connected. Journal für die reine und angewandte Mathematik (Crelles Journal), 2024(806):1–7, 2024.
  • [68] Sergio Zamora. Limits of almost homogeneous spaces and their fundamental groups. Groups, Geometry, and Dynamics, 18(3):761–798, 2024.
  • [69] Sergio Zamora and Xingyu Zhu. Topological rigidity of small RCD(K, N) spaces with maximal rank. arXiv preprint arXiv:2406.10189, 2024.