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

On the non-collapsed RCD spaces with local bounded covering geometry

Jikang Wang UC Berkeley, Berkeley, CA, US [email protected]
Abstract.

We consider a RCD((N1),N)(-(N-1),N) space (X,d,N)(X,d,\mathcal{H}^{N}) with local bounded covering geometry. The first result is related to Gromov’s almost flat manifold theorem. Specifically, if for every point p~\tilde{p} in the universal cover X~\widetilde{X}, we have N(B1(p~))v>0\mathcal{H}^{N}(B_{1}(\tilde{p}))\geq v>0 and the diameter of XX is sufficiently small, then XX is biHölder homeomorphic to an infranil-manifold. Moreover, if XX is a smooth Riemannian NN-manifold with Ric(N1)\mathrm{Ric}\geq-(N-1), then XX is biHölder diffeomorphic to an infranil-manifold. An application of our argument is to confirm the conjecture that Gromov’s almost flat manifold theorem holds in the RCD+CBA\mathrm{RCD}+\mathrm{CBA} setting.

The second result concerns a regular fibration theorem. Let (Xi,di,N)(X_{i},d_{i},\mathcal{H}^{N}) be a sequence of RCD((N1),N)(-(N-1),N) spaces converging to a compact smooth kk-dimensional manifold KK in the Gromov-Hausdorff sense. Assume that for any piXip_{i}\in X_{i}, the local universal cover is non-collapsing, i.e., for any pre-image point p~i\tilde{p}_{i} of pip_{i} in the universal cover of the ball B3(pi)B_{3}(p_{i}), we have N(B1(p~i))v\mathcal{H}^{N}(B_{1}(\tilde{p}_{i}))\geq v for some fixed v>0v>0. Then for sufficiently large ii, there exists a fibration map fi:XiKf_{i}:X_{i}\to K, where the fiber is an infra-nilmanifold and the structure group is affine.

Dedicated to Xiaochun Rong’s 70th Birthday

1. Introduction

In this paper, we study the topology of a RCD((N1),N)(-(N-1),N) space (X,d,N)(X,d,\mathcal{H}^{N}) which satisfies (ρ,v)(\rho,v)-bound covering condition, i.e., for any pXp\in X, take the universal cover of ρ\rho-ball Bρ(p)B_{\rho}(p) and p~\tilde{p} is a pre-image of pp, then N(Bρ/3(p~))v\mathcal{H}^{N}(B_{\rho/3}(\tilde{p}))\geq v, where ρ,v>0\rho,v>0.

We first review the topological theory of collapsing manifolds with sectional curvature bound. The one of most important theorems is Gromov’s almost flat theorem.

Theorem 1.1 (Gromov’s almost flat manifold theorem, [20, 34, 36]).

Given nn\in\mathbb{N}, there exists ϵ(n)>0\epsilon(n)>0 and C(n)>0C(n)>0 so that for any compact almost flat nn-manifold MM with

max(|secM|)diam(M)2<ϵ(n),\max(|\sec_{M}|)\mathrm{diam}(M)^{2}<\epsilon(n),

Then MM is diffeomorphic to an infra-nilmanifold, 𝒩/Γ\mathcal{N}/\Gamma, where 𝒩\mathcal{N} is a simply connected nilpotent nn-dim Lie group and Γ\Gamma is a discrete subgroup of 𝒩Aut(𝒩)\mathcal{N}\rtimes\text{Aut}(\mathcal{N}) and [Γ:Γ𝒩]C(n)[\Gamma:\Gamma\cap\mathcal{N}]\leq C(n).

A related result, due to Fukaya, asserts that if a sequence of nn-manifolds MiM_{i} with |sec|1|\sec|\leq 1 converging to a lower dimensional manifold KK, then there is a fibration map fi:MiKf_{i}:M_{i}\to K where the fibers are infra-nilmanifolds with an affine structure group.

Theorem 1.2 (smooth fibration, [16, 9]).

Assume that a sequence of nn-manifolds MiM_{i} with |secMi|1|\sec_{M_{i}}|\leq 1 converges to a compact lower dimensional manifold KK in the Gromov-Hausdorff sense, MiGHKM_{i}\overset{GH}{\longrightarrow}K. Then for ii large enough, there is a smooth fiberation map fi:MiNf_{i}:M_{i}\to N with fiber an infra-nilmanifold and an affine structure group, and fif_{i} is a Gromov-Hausdorff approximation (GHA).

More generally, the theories of singular fibration and nilpotent killing structure for collapsed manifolds with |sec|1|\sec|\leq 1 have been extensively studied in [9, 10, 11, 16, 17, 18]. However, Theorems 1.1 and 1.2 may not hold if we replace the sectional curvature by the Ricci curvature [2].

Around 2016, Rong proposed to investigate the class of nn-dim manifolds MM satisfying (ρ,v)(\rho,v)-bound Ricci covering geometry. Specifically, RicM(n1)\mathrm{Ric}_{M}\geq-(n-1), and for any point pMp\in M, take the universal cover of the ρ\rho-ball Bρ(p)B_{\rho}(p) with p~\tilde{p} a pre-image of pp. Then the Vol(Bρ/3(p~))v\mathrm{Vol}(B_{\rho/3}(\tilde{p}))\geq v. According to [9], any nn-manifold with |sec|1|\mathrm{sec}|\leq 1 satisfies the (ρ,v)(\rho,v)-bound Ricci covering condition for some ρ,v>0\rho,v>0 depending on nn.

Theorems 1.1 and 1.2 can be generalized to (ρ,v)(\rho,v)-bound Ricci covering geometry, see Theorem 1.3 and 1.8. If the diameter of MM is less than ρ\rho, then (ρ,v)(\rho,v)-bound covering condition is exactly that the universal cover M~\widetilde{M} is non-collapsing.

Theorem 1.3 ([24, 35]).

Given n,v>0n,v>0, there exists ϵ(n,v)>0\epsilon(n,v)>0 and C(n)>0C(n)>0, so that if a nn-manifold MM satisfies:

Ric(n1),diam(M)<ϵϵ(n,v),Vol(B1(p~))v,p~M~,\mathrm{Ric}\geq-(n-1),\ \mathrm{diam}(M)<\epsilon\leq\epsilon(n,v),\ \mathrm{Vol}(B_{1}(\tilde{p}))\geq v,\ \forall\tilde{p}\in\widetilde{M},

then MM is diffeomorphic to an infra-nilmanifold.

We summarize the proofs of Theorem 1.3 as follows. In [24], it was proved that the Ricci flow on MM exists for a definite time. After running the flow, we get an almost flat metric on MM. Then we can apply Gromov’s almost flat manifold theorem. The proof in [35] uses successively blowing up technique and the structure of iterated bundles, avoiding reliance on Gromov’s result.

Recently, Zamora and Zhu proved the topology rigidity for a RCD(K,N)(K,N) space with a small diameter, extending previous work in [28]. By the generalized Margulis lemma (Theorem 2.4), if the diameter of a RCD(K,N)(K,N) space (X,p)(X,p) is sufficiently small, then π1(X,p)\pi_{1}(X,p) contains a nilpotent subgroup GG with index C(N)\leq C(N). Then we can find a descending sequence

G=G1G2Gk={e}G=G_{1}\vartriangleright G_{2}\vartriangleright...\vartriangleright G_{k}=\{e\}

such that Gi/Gi+1G_{i}/G_{i+1} is cyclic, i=1,2,k1i=1,2...,k-1. Then we define the rank of π1(X)\pi_{1}(X) is the number of ii such that Gi/Gi+1G_{i}/G_{i+1} is infinite.

Theorem 1.4.

([45]) For any KK\in\mathbb{R} and N1N\geq 1, there exists ϵ>0\epsilon>0 such that for any RCD(K,N)(K,N) space (X,d,𝔪)(X,d,\mathfrak{m}) with diameter less than ϵ\epsilon, then rank(π1(X))N(\pi_{1}(X))\leq N. Moreover, if rank(π1(X))=N(\pi_{1}(X))=N, then XX is homeomorphic to an infranil-manifold of dimension NN.

Zamora and Zhu conjectured that the homeomorphism in Theorem 1.4 can be biHölder.

We call a RCD((N1),N)(-(N-1),N) space non-collapsed if the measure is the Hausdorff measure N\mathcal{H}^{N}; in particular, NN\in\mathbb{N}. We say that a sequence of RCD((N1),N)(-(N-1),N) spaces (Xi,di,N)(X_{i},d_{i},\mathcal{H}^{N}) are non-collapsing if there exists v>0v>0 such that for any pi(Xi,di,N)p_{i}\in(X_{i},d_{i},\mathcal{H}^{N}), N(B1(pi))v\mathcal{H}^{N}(B_{1}(p_{i}))\geq v. Sometimes, we simply say a RCD((N1),N)(-(N-1),N) space (X,d,N)(X,d,\mathcal{H}^{N}) non-collapsing if the Hausdorff measure of any one ball in XX is bounded below by a fixed number.

Remark 1.5.

Consider a RCD((N1),N)(-(N-1),N) space (X,d,N)(X,d,\mathcal{H}^{N}) with a small diameter. By [45], if rank(π1(X))=N(\pi_{1}(X))=N then the universal cover X~\widetilde{X} is non-collapsing. Conversely, if the universal cover X~\widetilde{X} is non-collapsing, then rank(π1(X))=N(\pi_{1}(X))=N by our proof of Theorem A.

At present, the proofs of Theorem 1.3 and the rigidity part of Theorem 1.4 are different. The proofs in [24, 35] for Theorem 1.3 rely on the smooth structure, and hence cannot be directly extended to the non-smooth setting. On the other hand, the proof in [45] for the rigidity part of Theorem 1.4 uses a topological result for aspherical manifolds to find the homeomorphism, but does not prove diffeomorphism in the smooth case. The first main result in this paper is to give a new proof that works for both Theorem 1.3 and the rigidity part of Theorem 1.4. Further, we can show that the homeomorphism is biHölder, thereby confirming the conjecture in [45].

Theorem A.

Given N,v>0N,v>0, there exists ϵ(N,v)>0\epsilon(N,v)>0 and C(N)>0C(N)>0, so that if a RCD((N1),N)(-(N-1),N) space (X,d,N)(X,d,\mathcal{H}^{N}) satisfies:

diam(X)<ϵϵ(N,v),N(B1(p~))v,p~X~,\mathrm{diam}(X)<\epsilon\leq\epsilon(N,v),\ \mathcal{H}^{N}(B_{1}(\tilde{p}))\geq v,\ \forall\tilde{p}\in\widetilde{X},

then XX is biHölder homeomorphic to an infranil-manifold 𝒩/Γ\mathcal{N}/\Gamma where 𝒩\mathcal{N} has a left-invariant metric. More precisely, there exists f:X𝒩/Γf:X\to\mathcal{N}/\Gamma such that for all x,yXx,y\in X,

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

where Φ(ϵ|N,v)0\Phi(\epsilon|N,v)\to 0 as ϵ0\epsilon\to 0. Moreover, if XX is a smooth NN-manifold with Ric(N1)\mathrm{Ric}\geq-(N-1), then ff is a diffeomorphism.

The proof of Theorem A can be extended to metric spaces with mixed curvature. Kapovitch showed that Gromov’s almost flat manifolds theorem holds for weighted closed manifolds with upper sectional and lower Bakry–Emery Ricci curvature bounds.

Theorem 1.6 ([26]).

For any 1<N<1<N<\infty, there exists ϵ>0\epsilon>0 such that the following hold. If (M,g,efn)(M,g,e^{-f}\mathcal{H}^{n}) is a weighted closed Riemannian nn-manifold with nNn\leq N, secϵ\mathrm{sec}\leq\epsilon, diamϵ\mathrm{diam}\leq\epsilon and Ricf,Nϵ\mathrm{Ric}_{f,N}\geq-\epsilon, then MM is diffeomorphic to an infranil-manifold.

We briefly recall the proof of Theorem 1.6. Kapovitch first proved that MM is aspherical by a fibration theorem and an induction argument. By some topological results, MM is homeomorphic to an infranil-manifold. Then applying Ricci flow smoothing techniques and Gromov’s almost flat theorem, MM is diffeomorphic to an infranil-manifold.

It was conjectured in [26] that Gromov’s almost flat manifold theorem also holds under the RCD+CBA\mathrm{RCD}+\mathrm{CBA} conditions. We recall the structure theory of RCD+CBA\mathrm{RCD}+\mathrm{CBA} spaces from [27]. If (X,d,𝔪)(X,d,\mathfrak{m}) is RCD(K,N)\mathrm{RCD}(K,N) and CBA(k)\mathrm{CBA}(k) with N<N<\infty then XX is a topological manifold with boundary of dimension N\leq N. The manifold part of XX is a smooth C1C^{1}-manifold with a C0BVC^{0}\cap\mathrm{BV} Riemannian metric which induces the distance function dd on XX. In particular, if X=\partial X=\emptyset, then XX is a smooth manifold.

Modify the proof Theorem A a little bit, we can prove the following conjecture in [26].

Theorem 1.7.

For any 1<N<1<N<\infty there exists ϵ(N)\epsilon(N) such that for any ϵ<ϵ(N)\epsilon<\epsilon(N) the following holds. If (X,d,𝔪)(X,d,\mathfrak{m}) is an RCD(ϵ,N)\mathrm{RCD}(-\epsilon,N) space such that (X,d)(X,d) is CBA(ϵ)\mathrm{CBA}(\epsilon), X=\partial X=\emptyset and diam(X)ϵ\mathrm{diam}(X)\leq\epsilon, then XX is biLipschitz diffeomorphic to an infranil-manifold of dimension N\leq N. That is, there exist a infranil-manifold 𝒩/Γ\mathcal{N}/\Gamma and a diffeomorphic map f:X𝒩/Γf:X\to\mathcal{N}/\Gamma so that

(1Φ(ϵ|N))d(x,y)d(f(x),f(y))(1+Φ(ϵ|N))d(x,y),(1-\Phi(\epsilon|N))d(x,y)\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.

It was known in [26] that Theorem 1.7 holds up to homeomorphism, while it remains unknown whether we can apply Ricci flow to smooth the metric. In particular, the proof for Theorem 1.7 provides an alternative approach to proving Theorem 1.6, without relying on Ricci flow smoothing techniques or Gromov’s almost flat manifold theorem.

Our next result is to prove a regular fibraion theorem in the RCD setting. The following fibration theorem in the smooth case was proved by Huang and Rong.

Theorem 1.8 ([23, 35]).

Given n,v>0n,v>0, there exists ϵ(n,v)>0\epsilon(n,v)>0 and C(n)>0C(n)>0 so that for any compact nn-manifold MM and kk-manifold KK satisfying:

RicM(n1)\text{Ric}_{M}\geq-(n-1), (1,v)(1,v)-bound covering geometry holds on MM,

|secK|1|\text{sec}_{K}|\leq 1, injK1\text{inj}_{K}\geq 1, dGH(M,K)<ϵ<ϵ(n,v)d_{GH}(M,K)<\epsilon<\epsilon(n,v),

then there is a smooth fiber bundle map, f:MKf:M\to K that is a Φ(ϵ|n,v)\Phi(\epsilon|n,v)-Gromov-Hausdorff approximation (GHA), where Φ(ϵ|n,v)0\Phi(\epsilon|n,v)\to 0 as ϵ0\epsilon\to 0. The fiber is an infranil-manifold and the structure group is affine

Huang first constructed the fibration in [23], and Rong proved that the fiber is an infranil-manifold and that the structure group be can be reduced to be affine in [35].

We generalize Theorem 1.8 to RCD((N1),N)(-(N-1),N) spaces (X,d,N)(X,d,\mathcal{H}^{N}) with (ρ,v)(\rho,v)-bound covering geometry.

Theorem B.

Given N,v>0N,v>0, suppose that a sequence of compact RCD((N1),N)(-(N-1),N) spaces (Xi,di,N)(X_{i},d_{i},\mathcal{H}^{N}), with (1,v)(1,v)-bound covering geometry, converges to a compact smooth kk-manifold KK in the Gromov-Hausdorff sense. Then for sufficiently large ii, there exists a fiber bundle map fi:XiKf_{i}:X_{i}\to K that is an ϵi\epsilon_{i}-Gromov-Hausdorff approximation(GHA) where ϵi0\epsilon_{i}\to 0 as ii\to\infty. Moreover, the fiber is homeomorphic to an infranil-manifold and the structure group is affine.

Next, we study the limit of RCD((N1),N)(-(N-1),N) spaces (Xi,di,N)(X_{i},d_{i},\mathcal{H}^{N}) with (ρ,v)(\rho,v)-bound covering geometry. By [25, 46], for any k3k\geq 3, there exists a collapsing Ricci limit space with the rectifiable dimension equal to kk, which contains no manifold points. We shall prove that, manifold points in the limit of (Xi,di,N)(X_{i},d_{i},\mathcal{H}^{N}) with (ρ,v)(\rho,v)-bound covering geometry have full measure.

Theorem 1.9.

Assume that (Xi,di,N,pi)(X_{i},d_{i},\mathcal{H}^{N},p_{i}) is a sequence of pointed RCD((N1),N)(-(N-1),N) spaces with (1,v)(1,v)-bound covering geometry, and suppose that

(Xi,di,N,pi)pmGH(X,d,𝔪,p)(X_{i},d_{i},\mathcal{H}^{N},p_{i})\overset{\mathrm{pmGH}}{\longrightarrow}(X,d,\mathfrak{m},p)

with rectifiable dim(X)=k<N(X)=k<N. Then
(1) (kk-regular points are manifold points) For any qk(X)q\in\mathcal{R}^{k}(X), a neighborhood of qq is biHölder to an open set in k\mathbb{R}^{k};
(2) (fibration near kk-regular point) If qk(X)q\in\mathcal{R}^{k}(X) and qiXiq_{i}\in X_{i} converging to qq, for sufficiently large ii, a neighborhood UiU_{i} is of qiq_{i} is biHölder to V×𝒩i/ΓiV\times\mathcal{N}_{i}/\Gamma_{i} where VV is a neighborhood of qq and 𝒩i/Γi\mathcal{N}_{i}/\Gamma_{i} is an infranil-manifold.

Remark 1.10.

It was shown by Rong that in the setting of Theorem 1.9, the Hausdorff dimension of (X,d)(X,d) is equal to kk and any tangent cone of XX is a metric cone.

Now we sketch our proof of Theorem A. We proceed by contradiction. Suppose that there exists a sequence of RCD((N1),N)(-(N-1),N) spaces (Xi,di,N,pi)(X_{i},d_{i},\mathcal{H}^{N},p_{i}) with non-collapsing universal covers and diam(Xi)0\mathrm{diam}(X_{i})\to 0, while none of XiX_{i} is biHölder to an infranil-manifold. By the generalized Margulis lemma, there is a nilpotent subgroup GiG_{i} of Gi=π1(Xi,pi)G_{i}^{\prime}=\pi_{1}(X_{i},p_{i}) with index C(n)\leq C(n). We may blow up the metric slowly at a regular point if necessary, and assume that the following diagram holds:

(X~i,p~i,Gi,Gi)eGH(N,p~,G,G)ππX~i/GiGHpt.\begin{CD}(\widetilde{X}_{i},\tilde{p}_{i},G_{i},G_{i}^{\prime})@>{eGH}>{}>(\mathbb{R}^{N},\tilde{p},G,G^{\prime})\\ @V{}V{\pi}V@V{}V{\pi}V\\ \widetilde{X}_{i}/G_{i}@>{GH}>{}>\mathrm{pt}.\end{CD}

We can show that GG actions are free translation actions, which can be identified as N\mathbb{R}^{N}. Therefore by the structure theorem of approximate groups [3, 44], a neighborhood of the identity in GiG_{i} forms a nilprogression. Roughly speaking, a nilprogression is a subset of a lattice in a simply connected nilpotent Lie group, and the nilprogression contains all generators and relations of the lattice.

Since the diameter of X~i/Gi\widetilde{X}_{i}/G_{i} converges to 0, GiG_{i} is determined by the neighborhood of the identity. This small nrighborhood contains all generators and relations of GiG_{i}. Thus GiG_{i} must be isomorphic to a lattice (the groupfication of the nilprogression) in a simply connected nilpotent NN-dim Lie group 𝒩i\mathcal{N}_{i}, where the Lie algebra structure of 𝒩i\mathcal{N}_{i} converges to the one of N\mathbb{R}^{N}. Hence we can endow 𝒩i\mathcal{N}_{i} with a left-invariant metric, which pointed converges to the flat metric on N\mathbb{R}^{N} in the C4C^{4}-sense.

Next we identify GiG_{i}^{\prime} as a subgroup in 𝒩iAut(𝒩i)\mathcal{N}_{i}\rtimes\text{Aut}(\mathcal{N}_{i}) by the rigidity. (X~i,p~i,Gi)(\widetilde{X}_{i},\tilde{p}_{i},G_{i}^{\prime}) is eGH-close to (𝒩i,e,Gi)(\mathcal{N}_{i},e,G_{i}^{\prime}) on the 1ϵ\frac{1}{\epsilon}-ball of the base point for some fixed small ϵ>0\epsilon>0. Thus by an extension lemma, we can construct a global map h:X~i𝒩i,h:\widetilde{X}_{i}\to\mathcal{N}_{i}, which is almost GiG_{i}^{\prime}-equivariant and a GHA on any 1ϵ\frac{1}{\epsilon}-ball in X~i\widetilde{X}_{i}. Specifically, for any x~X~i\tilde{x}\in\widetilde{X}_{i}, h:Br(x~)𝒩ih:B_{r}(\tilde{x})\to\mathcal{N}_{i} is a ϵ\epsilon-GHA to its image; d(h(gx~),gh(x~))ϵd(h(g\tilde{x}),gh(\tilde{x}))\leq\epsilon for any gGig\in G_{i}^{\prime} and x~X~i\tilde{x}\in\widetilde{X}_{i}.

Then we can find a normal subgroup Gi′′G_{i}^{\prime\prime} of GiG_{i}^{\prime} with finite index, so that Gi′′GiG_{i}^{\prime\prime}\subset G_{i} and Gi′′B1ϵ(e)=G_{i}^{\prime\prime}\cap B_{\frac{1}{\epsilon}}(e)=\emptyset where e𝒩ie\in\mathcal{N}_{i}. Thus (X~i/Gi′′,Gi/Gi′′)(\widetilde{X}_{i}/G_{i}^{\prime\prime},G_{i}^{\prime}/G_{i}^{\prime\prime}) is eGH close to (𝒩i/Gi′′,Gi/Gi′′)(\mathcal{N}_{i}/G_{i}^{\prime\prime},G_{i}^{\prime}/G_{i}^{\prime\prime}) on any 1ϵ\frac{1}{\epsilon}-ball. Then we apply an averaging technique (see Theorem 3.5) to obtain a Gi/Gi′′G_{i}^{\prime}/G_{i}^{\prime\prime}-equivariant map fGi/Gi′′f_{G_{i}^{\prime}/G_{i}^{\prime\prime}} from X~i/Gi′′\widetilde{X}_{i}/G_{i}^{\prime\prime} to 𝒩i/Gi′′\mathcal{N}_{i}/G_{i}^{\prime\prime}, which is locally almost NN-splitting. By the canonical Reifenberg method from [12] and [22], fGi/Gi′′f_{G_{i}^{\prime}/G_{i}^{\prime\prime}} must be a biHöder homeomorphism. Since fGi/Gi′′f_{G_{i}^{\prime}/G_{i}^{\prime\prime}} is Gi/Gi′′G_{i}^{\prime}/G_{i}^{\prime\prime}-equivariant, it follows that Xi=X~i/GiX_{i}=\widetilde{X}_{i}/G_{i}^{\prime} is biHölder homeomorphic to the infranil-manifold 𝒩i/Gi\mathcal{N}_{i}/G_{i}^{\prime}. Then we finish the proof of Theorem A.

The smooth fibration map ff in Theorem 1.8 is constructed by averaging and gluing some local almost kk-splitting maps. Thus ff is a smooth GHA. Then Huang compared ff to a linear average and showed that dfdf is non-degenerate using the local bounded covering geometry. Therefore ff is a fibration map by the implicit function theorem.

In the non-smooth setting in Theorem B, we can construct a GHA fi:XiKf_{i}:X_{i}\to K which is a locally almost kk-splitting map. Although there is no implicit function theorem in the non-smooth setting, we can prove that fif_{i} is a fibration with infranil-manifold fiber by applying the proof in Theorem A to the collapsing directions of XiX_{i}. The structure group can be affine because the nilpotent structure of the fiber is independent of the choice of the base point; see also [35].

Remark 1.11.

Pointed Gromov-Hausdorff approximations typically cannot provide information about the global topology of non-compact spaces, as they do not capture the geometry outside of a large ball. This limitation is why we consider a global map hh which is almost equivariant and acts as a GHA on any 1ϵ\frac{1}{\epsilon}-ball.

Acknowledgments. The author would like to thank Xiaochun Rong, Vitali Kapovitch, Jiayin Pan, Shicheng Xu, Xingyu Zhu and Sergio Zamora for helpful discussions.

2. Preliminaries

2.1. Equivariant Gromov-Hausdorff convergence and isometry group on a Ricci limit space

We review the notion of equivariant Gromov-Hausdorff convergence introduced by Fukaya and Yamaguchi [15, 19].

Let (X,p)(X,p) and (Y,q)(Y,q) be two pointed metric spaces. Let HH and KK be closed subgroups of Isom(X)\mathrm{Isom}(X) and Isom(Y)\mathrm{Isom}(Y), respectively. For any r>0r>0, define the sets

H(p,r)={hH|d(hp,p)r},K(q,r)={kK|d(kq,q)r}.H(p,r)=\{h\in H|d(hp,p)\leq r\},\ K(q,r)=\{k\in K|d(kq,q)\leq r\}.

For ϵ>0\epsilon>0, a pointed ϵ\epsilon-equivariant Gromov-Hausdorff approximation (or simply an ϵ\epsilon-eGHA) is a triple of maps (f,ϕ,ψ)(f,\phi,\psi) where:

f:B1ϵ(p)B1ϵ+ϵ(q),ϕ:H(p,1ϵ)K(q,1ϵ),ψ:K(q,1ϵ)H(p,1ϵ)f:B_{\frac{1}{\epsilon}}(p)\to B_{\frac{1}{\epsilon}+\epsilon}(q),\quad\phi:H(p,\frac{1}{\epsilon})\to K(q,\frac{1}{\epsilon}),\quad\psi:K(q,\frac{1}{\epsilon})\to H(p,\frac{1}{\epsilon})

satisfying the following conditions:
(1) f(p)=qf(p)=q, f(B1ϵ(p))f(B_{\frac{1}{\epsilon}}(p)) is 2ϵ2\epsilon-dense in B1ϵ+ϵ(q)B_{\frac{1}{\epsilon}+\epsilon}(q) and |d(f(x1),f(x2))d(x1,x2)|ϵ|d(f(x_{1}),f(x_{2}))-d(x_{1},x_{2})|\leq\epsilon for all x1,x2B1ϵ(p)x_{1},x_{2}\in B_{\frac{1}{\epsilon}}(p);
(2) d(ϕ(h)f(x),f(hx))<ϵd(\phi(h)f(x),f(hx))<\epsilon for all hH(1ϵ)h\in H(\frac{1}{\epsilon}) and xB1ϵ(p)x\in B_{\frac{1}{\epsilon}}(p);
(3) d(kf(x),f(ψ(k)x))<ϵd(kf(x),f(\psi(k)x))<\epsilon for all kK(1ϵ)k\in K(\frac{1}{\epsilon}) and xB1ϵ(p)x\in B_{\frac{1}{\epsilon}}(p).

The equivariant Gromov-Hausdorff(eGH) distance deGH((Xi,pi,Gi),(X,p,G))d_{eGH}((X_{i},p_{i},G_{i}),(X,p,G)) is defined as the infimum of ϵ\epsilon so that there exists a ϵ\epsilon-eGHA. A sequence of metric space with isometric actions (Xi,pi,Gi)(X_{i},p_{i},G_{i}) converges to a limit space (X,p,G)(X,p,G), if deGH((Xi,pi,Gi),(X,p,G))0d_{eGH}((X_{i},p_{i},G_{i}),(X,p,G))\to 0.

Given a Gromov-Hausdorff approximation(GHA) ff as in condition (1) above, we can construct an admissible metric on the disjoint union B1ϵ(p)B1ϵ(q)B_{\frac{1}{\epsilon}}(p)\sqcup B_{\frac{1}{\epsilon}}(q) so that

B1ϵ(p)B1ϵ(p)B1ϵ(q),B1ϵ(q)B1ϵ(p)B1ϵ(q)B_{\frac{1}{\epsilon}}(p)\hookrightarrow B_{\frac{1}{\epsilon}}(p)\sqcup B_{\frac{1}{\epsilon}}(q),B_{\frac{1}{\epsilon}}(q)\hookrightarrow B_{\frac{1}{\epsilon}}(p)\sqcup B_{\frac{1}{\epsilon}}(q)

are isometric embedding and for any xB1ϵ(p)x\in B_{\frac{1}{\epsilon}}(p),d(x,f(x))2ϵd(x,f(x))\leq 2\epsilon. We always assume such an admissible metric whenever Gromov-Hausdorff distance between two metric spaces are small.

We have the following pre-compactness theorem for equivariant Gromov-Hausdorff convergence [15, 19].

Theorem 2.1 ([15, 19]).

Let (Xi,pi)(X_{i},p_{i}) be a sequence of metric spaces converging to a limit space (X,p)(X,p) in the pointed Gromov-Hausdorff sense. For each ii, let GiG_{i} be a closed subgroup of Isom(Xi)\mathrm{Isom}(X_{i}), the isometry group of XiX_{i}. Then passing to a subsequence if necessary,

(Xi,pi,Gi)eGH(X,p,G),(X_{i},p_{i},G_{i})\overset{eGH}{\longrightarrow}(X,p,G),

where GG is a closed subgroup of Isom(X)\mathrm{Isom}(X). Moreover, the quotient spaces (Xi/Gi,p¯i)(X_{i}/G_{i},\bar{p}_{i}) pointed Gromov-Hausdorff converge to (X/G,p)(X/G,p).

2.2. Geometric Theory of RCD(K,N)(K,N) spaces

In this subsection, we review the structure theory of RCD(K,N)(K,N) spaces. We assume that the reader is familiar with the basic notions of RCD spaces. A measured metric RCD(K,N)(K,N) space (X,d,𝔪)(X,d,\mathfrak{m}) is non-collapsed if NN\in\mathbb{N} and 𝔪=N\mathfrak{m}=\mathcal{H}^{N}.

We begin by defining regular points on an RCD space. We use the notation 0k0^{k} to refer the origin in k\mathbb{R}^{k}.

Definition 2.2.

Let (X,d,𝔪)(X,d,\mathfrak{m}) be an RCD(K,N)(K,N) space. Given ϵ>0\epsilon>0, r>0r>0, and kk\in\mathbb{N}, we define

ϵ,rk(X)={xX:dGH(Bs(x),Bs(0k))<ϵs, 0<s<2r},\mathcal{R}_{\epsilon,r}^{k}(X)=\{x\in X:d_{GH}(B_{s}(x),B_{s}(0^{k}))<\epsilon s,\forall\ 0<s<2r\},

where 0kk0^{k}\in\mathbb{R}^{k}, and

ϵk(X)=r>0ϵ,rk(X),k(X)=ϵ>0ϵk(X).\mathcal{R}_{\epsilon}^{k}(X)=\bigcup_{r>0}\mathcal{R}_{\epsilon,r}^{k}(X),\ \mathcal{R}^{k}(X)=\bigcap_{\epsilon>0}\mathcal{R}_{\epsilon}^{k}(X).

For any RCD(K,N)(K,N) space (X,d,𝔪)(X,d,\mathfrak{m}), there exists kNk\leq N s.t. 𝔪(Xk(X))=0\mathfrak{m}(X-\mathcal{R}^{k}(X))=0 [5]. We call this kk the rectifiable dimension of (X,d,𝔪)(X,d,\mathfrak{m}).

Then we review the theory of almost splitting for RCD spaces, following the notations in [4].

Definition 2.3.

Let (X,d,𝔪)(X,d,\mathfrak{m}) be a RCD(K,N)(K,N) space for some KK\in\mathbb{R} and 1N<1\leq N<\infty. Let pXp\in X and s>0s>0. A map u:B2s(p)ku:B_{2s}(p)\to\mathbb{R}^{k} is a (k,δ)(k,\delta)-splitting map if it belongs to the domain of the local Laplacian on B2s(p)B_{2s}(p), and

|u||Bs(p)C(N),|\nabla u|_{|B_{s}(p)}\leq C(N),
a,b=1kBs(p))|ua,ubδab|δ2,\sum_{a,b=1}^{k}\fint_{B_{s}(p))}|\langle\nabla u^{a},\nabla u^{b}\rangle-\delta_{a}^{b}|\leq\delta^{2},
a=1kBs(p))s2|Hess(ua)|2δ2.\sum_{a=1}^{k}\fint_{B_{s}(p))}s^{2}|\mathrm{Hess}(u^{a})|^{2}\leq\delta^{2}.

Such a map uu is sometimes referred as almost kk-splitting if it is (k,δ)(k,\delta)-splitting for some sufficiently small δ>0\delta>0. In the literature, it is often assumed that an almost kk-splitting map is harmonic. However, it is convenient to drop the harmonicity assumption in this paper. By Lemma 4.4 in [22], if we further assume that B2s(p)|Δu|δ\fint_{B_{2s}(p)}|\nabla\Delta u|\leq\delta, then |u||Bs(p)1+Φ(δ|N,k)|\nabla u|_{|B_{s}(p)}\leq 1+\Phi(\delta|N,k) where Φ(δ|N,k)0\Phi(\delta|N,k)\to 0 as δ0\delta\to 0.

It is a classical result that the existence of an almost splitting function is equivalent to pmGH closeness to a space that splits off a Euclidean factor, see [4, 7, 8, 13].

We will use the fact that isometry group of a RCD(K,N)\mathrm{RCD}(K,N) space is a Lie group [39, 21].

By [14], the generalized Margulis lemma holds for any RCD(K,N)(K,N) space, see also [28] for the manifolds with Ricci curvature bounded from below.

Theorem 2.4 (Generalized Margulis lemma, [14]).

For any KK\in\mathbb{R} and N1N\geq 1, there exists CC and ϵ0>0\epsilon_{0}>0 such that for any RCD(K,N)(K,N) space (X,d,𝔪,p)(X,d,\mathfrak{m},p) with rectifiable dimension kk, the image of the natural homomorphism

π1(Bϵ0(p),p)π1(B1(p),p)\pi_{1}(B_{\epsilon_{0}}(p),p)\to\pi_{1}(B_{1}(p),p)

contains a normal nilpotent subgroup of index C\leq C. Moreover, this nilpotent subgroup has a nilpotent basis of length at most kk.

We now recall the volume convergence theorem for non-collapsed RCD(K,N)(K,N) spaces.

Theorem 2.5 (Volume convergence, [33]).

Assume that (Xi,di,N,pi)(X_{i},d_{i},\mathcal{H}^{N},p_{i}) is a sequence of non-collapsed RCD(K,N)(K,N) spaces which pointed measured GH converge to (X,d,𝔪,p)(X,d,\mathfrak{m},p). If N(B1(pi))v>0\mathcal{H}^{N}(B_{1}(p_{i}))\geq v>0, then 𝔪=N\mathfrak{m}=\mathcal{H}^{N} and r>0\forall r>0, N(Br(pi))\mathcal{H}^{N}(B_{r}(p_{i})) converges to N(Br(p))\mathcal{H}^{N}(B_{r}(p)).

We next state the transformation theorem in [4], see also [12, 13, 22].

Theorem 2.6 (Transformation, [4]).

Let 1N<1\leq N<\infty. For any δ>0\delta>0 there exists ϵ(N,δ)>0\epsilon(N,\delta)>0 such that for any ϵ<ϵ(N,δ)\epsilon<\epsilon(N,\delta) and any xx in a RCD(ϵ2(N1),N)(-\epsilon^{2}(N-1),N) space (X,d,N)(X,d,\mathcal{H}^{N}), the following holds. If Bs(x)B_{s}(x) is an (N,ϵ2)(N,\epsilon^{2})-symmetric ball for any r0s1r_{0}\leq s\leq 1 and u:B2(x)Nu:B_{2}(x)\to\mathbb{R}^{N} is a (N,ϵ)(N,\epsilon)-splitting map, then for each scale r0s1r_{0}\leq s\leq 1 there exists an N×NN\times N lower triangular matrix TsT_{s} such that
(1) Tsu:Bs(x)NT_{s}u:B_{s}(x)\to\mathbb{R}^{N} is an (N,δ)(N,\delta)-splitting map on Bs(x)B_{s}(x);
(2) Bs(x)(Tsu)a(Tsu)bdN=δab\fint_{B_{s}(x)}\nabla(T_{s}u)^{a}\cdot\nabla(T_{s}u)^{b}d\mathcal{H}^{N}=\delta_{a}^{b};
(3) |TsT2s1Id|δ|T_{s}\circ T_{2s}^{-1}-\mathrm{Id}|\leq\delta.

An application of the transformation theorem is to construct a biHöder map [12, 22].

Theorem 2.7 (Canonical Reifenberg method [12, 22]).

Assume that (X,d,n,p)(X,d,\mathcal{H}^{n},p) is a RCD(ϵ2(N1),N1)(-\epsilon^{2}(N-1),N-1) 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,v))d(x,y)1+Φ(ϵ|n,v)d(f(x),f(y))(1+Φ(ϵ|n,v))d(x,y).(1-\Phi(\epsilon|n,v))d(x,y)^{1+\Phi(\epsilon|n,v)}\leq d(f(x),f(y))\leq(1+\Phi(\epsilon|n,v))d(x,y).

Moreover, if XX is a smooth NN-manifold with Ricϵ2(N1)\geq-\epsilon^{2}(N-1), then for any xB1(p)x\in B_{1}(p), du:TxNdu:T_{x}\to\mathbb{R}^{N} is nondegenerate.

In Theorem 2.7, the harmonicity condition on uu can be replaced by the condition Bs(p)|Δu|ϵ\fint_{B_{s}(p)}|\nabla\Delta u|\leq\epsilon [22].

2.3. The universal covers and local relatives covers of RCD(K,N)(K,N) spaces

For any RCD(K,N)(K,N) space (X,d,𝔪)(X,d,\mathfrak{m}), the universal cover X~\widetilde{X} exists and is a RCD(K,N)(K,N) space with induced metric and measure [31]. XX is semi-locally simply connected [41, 42], thus the fundamental group is isomorphic to the deck transformation group of X~\widetilde{X}.

We first consider a compact RCD(K,N)(K,N) space (X,d,𝔪,p)(X,d,\mathfrak{m},p) with diameter less than DD. Let (X~,p~)(\widetilde{X},\tilde{p}) be the universal cover and GG be the fundamental group of XX. Define the set

G(p~,20D)={gG|d(p~,gp~)20D}.G(\tilde{p},20D)=\{g\in G|d(\tilde{p},g\tilde{p})\leq 20D\}.

We shall review that B¯10D(p~)\bar{B}_{10D}(\tilde{p}) and G(p~,20D)G(\tilde{p},20D) determine X~\widetilde{X} and GG [19, 37, 40].

G(p~,20D)G(\tilde{p},20D) is a pseudo-group, i.e. for some g1,g2G(p~,20D)g_{1},g_{2}\in G(\tilde{p},20D), g1g2g_{1}g_{2} is not defined within G(p~,20D)G(\tilde{p},20D). To handle this, we define the groupfication G^\hat{G} of G(p~,20D)G(\tilde{p},20D) as follows. Let FF be the free group generated by elements ege_{g} for each gG(p~,20D)g\in G(\tilde{p},20D). We may quotient FF by the normal subgroup generated by all elements of the form eg1eg2eg1g21e_{g_{1}}e_{g_{2}}e_{g_{1}g_{2}}^{-1}, where g1,g2G(p~,20D)g_{1},g_{2}\in G(\tilde{p},20D) with g1g2G(p~,20D)g_{1}g_{2}\in G(\tilde{p},20D). The quotient group is denoted G^\hat{G}.

There is a natural (pseudo-group) homomorphism

i:G(p~,20D)G^,i(g)=[eg],i:G(\tilde{p},20D)\to\hat{G},\ i(g)=[e_{g}],

where [eg][e_{g}] is the quotient image of ege_{g}. Define

π:G^G,π([eg1eg2egk])=g1g2gk.\pi:\hat{G}\to G,\ \pi([e_{g_{1}}e_{g_{2}}...e_{g_{k}}])=g_{1}g_{2}...g_{k}.

Then πi\pi\circ i is the identity map on G(p~,20D)G(\tilde{p},20D), thus ii is injective. Since G(p~,20D)G(\tilde{p},20D) generates GG, π\pi is surjective.

Now we can glue a space by equivalence relation,

G^×G(p~,20D)B¯10D(p~)=G^×B¯10D(p~)/,\hat{G}\times_{G(\tilde{p},20D)}\bar{B}_{10D}(\tilde{p})=\hat{G}\times\bar{B}_{10D}(\tilde{p})/\sim,

where the equivalence relation is given by (g^i(g),x)(g^,gx)(\hat{g}i(g),x)\sim(\hat{g},gx) for all gG(p~,20D),g^G^,xB¯10D(p~)g\in G(\tilde{p},20D),\hat{g}\in\hat{G},x\in\bar{B}_{10D}(\tilde{p}) with gxB¯10D(p~)gx\in\bar{B}_{10D}(\tilde{p}). Endow G^×G(p~,20D)B¯10D(p~)\hat{G}\times_{G(\tilde{p},20D)}\bar{B}_{10D}(\tilde{p}) with the induced length metric. Then G^×G(p~,20D)B¯10D(p~)\hat{G}\times_{G(\tilde{p},20D)}\bar{B}_{10D}(\tilde{p}) is a covering space of X~\tilde{X} with deck transformation group ker(π)\ker(\pi). X~\tilde{X} is simply connected, thus we obtain the following result.

Theorem 2.8.

([19, 37, 40]) G^\hat{G} is isomorphic to GG and G^×G(p~,20D)B¯10D(p~)\hat{G}\times_{G(\tilde{p},20D)}\bar{B}_{10D}(\tilde{p}) with induced length metric is isometric to X~\widetilde{X}.

We next consider the local relative covers of a RCD(K,N)(K,N) space (X,d,𝔪)(X,d,\mathfrak{m}). By [38], there exists a sequence of compact nn-manifolds (Mi,pi)(M_{i},p_{i}) with a uniform lower sectional curvature bound, so that the local universal cover of B1(pi)~\widetilde{B_{1}(p_{i})} admits no converging subsequence in the pointed Gromov-Hausdorff sense. To address this, we refer to a precompactness theorem for relative covers of open balls by Xu [43].

For any pXp\in X, r2>r1>0r_{2}>r_{1}>0, define local relative over (B~(p,r1,r2),p~)(\widetilde{B}(p,r_{1},r_{2}),\tilde{p}) as the a connected component of the pre-image of Br1(p)B_{r_{1}}(p) in the universal cover of Br2(p)B_{r_{2}}(p), where p~\tilde{p} is a pre-image point of pp.

Theorem 2.9 (Precompactness of relative covers, [43]).

(B~(p,r1,r2),p~)(\widetilde{B}(p,r_{1},r_{2}),\tilde{p}) equipped with its length metric and measure, is globally AK,N,r1,r2(r)A_{K,N,r_{1},r_{2}}(r)-doubling, that is, there exists a positive non-decreasing function AK,N,r1,r2(r)A_{K,N,r_{1},r_{2}}(r) such that

0<𝔪(Br(x))AK,N,r1,r2(r)𝔪(Br2(x)), for any x(B~(p,r1,r2),p~).0<\mathfrak{m}(B_{r}(x))\leq A_{K,N,r_{1},r_{2}}(r)\mathfrak{m}(B_{\frac{r}{2}}(x)),\text{ \ for any \ }x\in(\widetilde{B}(p,r_{1},r_{2}),\tilde{p}).

In particular, let GG be the image of π1(Br1(p),p)π1(Br2(p),p)\pi_{1}(B_{r_{1}}(p),p)\to\pi_{1}(B_{r_{2}}(p),p), then the family consisting of all such triples (B~(p,r1,r2),p~,G)(\widetilde{B}(p,r_{1},r_{2}),\tilde{p},G) is precompact in the pointed equivariant Gromov-Hausdorff topology.

Remark 2.10.

In Theorem 2.9, the quotient space B~(p,r1,r2)/G\widetilde{B}(p,r_{1},r_{2})/G is isometric to Br1(p)B_{r_{1}}(p) with the length metric on itself. The length metric on Br1(p)B_{r_{1}}(p) may differ from the original metric dd. However, they coincide on Br13(p)B_{\frac{r_{1}}{3}}(p). To simplify the notation, we always assume that the length metric on Br1(p)B_{r_{1}}(p) is the original metric, otherwise we consider Br13(p)B_{\frac{r_{1}}{3}}(p) with length metric on Br1(p)B_{r_{1}}(p).

We state some technical lemmas.

Lemma 2.11.

Let GG be a nilpotent Lie group and G0G_{0} be the identity component. Then for any compact subgroup KGK\subset G, the commutator [K,G0][K,G_{0}] is trivial.

Lemma 2.12 (Covering lemma, [28]).

There exists C(N)C(N) such that the following holds. Let (X,d,𝔪,p)(X,d,\mathfrak{m},p) be a RCD((N1),N)(-(N-1),N) space and f:Xf:X\to\mathbb{R} is a non-negative function. Let π:X~X\pi:\widetilde{X}\to X be the universal cover of XX and p~X~\tilde{p}\in\widetilde{X} is a lift of pp. Let f~=fπ\tilde{f}=f\circ\pi, then

C1(N)B13(p)fB1(p~)f~C(N)B1(p)f.C^{-1}(N)\fint_{B_{\frac{1}{3}}(p)}f\leq\fint_{B_{1}(\tilde{p})}\tilde{f}\leq C(N)\fint_{B_{1}(p)}f.
Lemma 2.13 (Gap lemma, [28]).

Assume that (Xi,pi,Gi)(X_{i},p_{i},G_{i}) is a sequence of length metric space and

(Xi,pi,Gi)eGH(X,p,G).(X_{i},p_{i},G_{i})\overset{eGH}{\longrightarrow}(X,p,G).

Assume that there exists b>a>0b>a>0 such that G(p,r)\langle G(p,r)\rangle is the same group for any r(a,b)r\in(a,b). Then there exists ϵi0\epsilon_{i}\to 0, so that for any sufficiently large ii, Gi(pi,r)\langle G_{i}(p_{i},r)\rangle is the same group for any r(a+ϵi,bϵi)r\in(a+\epsilon_{i},b-\epsilon_{i}).

2.4. Approximate groups and almost homogeneous spaces

The references of this subsection are [3, 44].

Definition 2.14.

A (symmetric) local group GG is a topological space with the identity element eGe\in G, together with a global inverse map ()1:GG()^{-1}:G\to G and a partially defined product map :ΩG\cdot:\Omega\to G, satisfying the following axioms:
(1) Ω\Omega is an open neighborhood of (G×{e}({e}×G)(G\times\{e\}\cup(\{e\}\times G) in G×GG\times G.
(2) The map ()1:xx1()^{-1}:x\to x^{-1} and :(x,y)xy\cdot:(x,y)\to xy are continuous.
(3) If g,h,kGg,h,k\in G s.t. that (gh)k(gh)k and g(hk)g(hk) are well-defined, then (gh)k=g(hk)(gh)k=g(hk).
(4) For any gGg\in G, eg=ge=geg=ge=g.
(5) For any gGg\in G, gg1gg^{-1} and g1gg^{-1}g are well-defined and equal to ee.

In particular, if Ω=G×G\Omega=G\times G, we call GG a global group or a topological group.

Definition 2.15.

Let GG be a local group and g1,g2,,gmGg_{1},g_{2},...,g_{m}\in G. We say that the product g1g2gmg_{1}g_{2}...g_{m} is well-defined, if for each 1ijm1\leq i\leq j\leq m we can find g[i,j]Gg_{[i,j]}\in G s.t. g[k,k]=gkg_{[k,k]}=g_{k} for any 1km1\leq k\leq m and g[i,j]g[j+1,k]g_{[i,j]}g_{[j+1,k]} is well-defined and equal to g[i,k]g_{[i,k]} for any 1ij<km1\leq i\leq j<k\leq m.

For sets A1,A2,,AmGA_{1},A_{2},...,A_{m}\subset G, we say the product A1A2AmA_{1}A_{2}...A_{m} is well-defined if for any choices of gjAjg_{j}\in A_{j}, g1g2gmg_{1}g_{2}...g_{m} is well-defined.

Definition 2.16.

AGA\subset G is called a multiplicative set if it is symmetric A=A1A=A^{-1}, and A200A^{200} is well-defined.

Definition 2.17.

Let AA be a finite symmetric subset of a multiplicative set and CC\in\mathbb{N}, AA is called a CC-approximate group if A2A^{2} can be covered by CC left translate of AA.

Definition 2.18.

Let AA be a CC-approximate group. We call AA a strong CC-approximate group if there is a symmetric set SAS\subset A so that
(1) ({asa1|sS,aA})103C3A(\{asa^{-1}|s\in S,a\in A\})^{10^{3}C^{3}}\subset A;
(2) if g,g2,,g1000A100g,g^{2},...,g^{1000}\in A^{100}, then gAg\in A;
(3) if g,g2,,g103C3A100g,g^{2},...,g^{10^{3}C^{3}}\in A^{100}, then gSg\in S.

Consider a multiplicative set AGA\subset G. For any gGg\in G, define the escape norm as

||g||A=:inf{1m+1|e,g,g2,,gmA}.||g||_{A}=:\inf\{\frac{1}{m+1}|e,g,g^{2},...,g^{m}\in A\}.
Theorem 2.19 (escape norm estimate, [3]).

For each C>0C>0, there is M>0M>0 s.t. if AA is a strong CC-approximate group and g1,g2,,gkA10g_{1},g_{2},...,g_{k}\in A^{10}, then
(1) g1g2gkAMj=1kgjA||g_{1}g_{2}...g_{k}||_{A}\leq M\sum_{j=1}^{k}||g_{j}||_{A}.
(2) g2g1g21A103g1A||g_{2}g_{1}g_{2}^{-1}||_{A}\leq 10^{3}||g_{1}||_{A}.
(3) [g1,g2]AMg1Ag2A||[g_{1},g_{2}]||_{A}\leq M||g_{1}||_{A}||g_{2}||_{A}.

Remark 2.20.

Due to (1) and (2), the set w={gA|gA=0}w=\{g\in A|||g||_{A}=0\} is a normal subgroup of AA. We call that AA contains no small subgroup if ww is trivial. For any general strong approximate group AA, A/wA/w is a strong approximate group with no small subgroup.

Readers may compare (3) in Theorem 2.19 with the holonomy estimates in Gromov’s argument for almost flat manifolds, see [6] Chapter 2.

Definition 2.21.

(nilprogression) Let GG be a local group, u1,u2,,urGu_{1},u_{2},...,u_{r}\in G and C1,C2,,Cr+C_{1},C_{2},...,C_{r}\in\mathbb{R}^{+}. The set P(u1,,ur;C1,,Cr)P(u_{1},...,u_{r};C_{1},...,C_{r}) is defined as the the set of 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 CiC_{i}. We call P(u1,,ur;C1,,Cr)P(u_{1},...,u_{r};C_{1},...,C_{r}) a nilprogession of rank rr if every word in it is well defined in GG. We say it a nilprogession in CC-normal form for some C>0C>0 if it satisfies the following properties:
(1) For all 1ijr1\leq i\leq j\leq r, we have

[ui±1,uj±1]P(uj+1,,ur;CCj+1CiCj,,CCrCiCj).[u_{i}^{\pm 1},u_{j}^{\pm 1}]\in P(u_{j+1},...,u_{r};\frac{CC_{j+1}}{C_{i}C_{j}},...,\frac{CC_{r}}{C_{i}C_{j}}).

(2) The expression u1n1urnru_{1}^{n_{1}}...u_{r}^{n_{r}} represent different elements in GG for |n1|C1C,,|nr|CrC|n_{1}|\leq\frac{C_{1}}{C},...,|n_{r}|\leq\frac{C_{r}}{C}.
(3) 1C(2C1+1)(2Cr+1)|P|C(2C1+1)(2Cr+1)\frac{1}{C}(2\lfloor C_{1}\rfloor+1)...(2\lfloor C_{r}\rfloor+1)\leq|P|\leq C(2\lfloor C_{1}\rfloor+1)...(2\lfloor C_{r}\rfloor+1).

For a nilprogression P(u1,,ur;C1,,Cr)P(u_{1},...,u_{r};C_{1},...,C_{r}) in CC-normal form and ϵ(0,1)\epsilon\in(0,1), define ϵP=P(u1,,ur;ϵC1,,ϵCr)\epsilon P=P(u_{1},...,u_{r};\epsilon C_{1},...,\epsilon C_{r}). Define the thickness of PP as the minimum of C1,,CrC_{1},...,C_{r} and we denote it by thick(P)(P). The set {u1n1urnr||n1|C1/C,,|nr|Cr/C}\{u_{1}^{n_{1}}...u_{r}^{n_{r}}||n_{1}|\leq C_{1}/C,...,|n_{r}|\leq C_{r}/C\} is called the grid part of PP and is denoted by G(P)G(P).

Definition 2.22.

Let P(u1,,ur;C1,,Cr)P(u_{1},...,u_{r};C_{1},...,C_{r}) be a nilprogression in CC-normal form with thick(P)>C(P)>C. Set ΓP\Gamma_{P} to be the abstract group generated by γ1,,γr\gamma_{1},...,\gamma_{r} with the relation [γj,γk]=γk+1βj,kk+1γrβj,kr[\gamma_{j},\gamma_{k}]=\gamma_{k+1}^{\beta_{j,k}^{k+1}}...\gamma_{r}^{\beta_{j,k}^{r}} whenever j<kj<k, where [uj,uk]=uk+1βj,kk+1urβj,kr[u_{j},u_{k}]=u_{k+1}^{\beta_{j,k}^{k+1}}...u_{r}^{\beta_{j,k}^{r}} and |βj,kl|CNlNjNk|\beta_{j,k}^{l}|\leq\frac{CN_{l}}{N_{j}N_{k}}. We say that PP is good if each element of ΓP\Gamma_{P} has a unique expression of the form γ1n1γrnr\gamma_{1}^{n_{1}}...\gamma_{r}^{n_{r}} with n1,,nrn_{1},...,n_{r}\in\mathbb{Z}.

Theorem 2.23 ([3, 44]).

For each r,C>0r\in\mathbb{N},C>0, there is ϵ>0\epsilon>0 so that the following holds. Let P(u1,,ur;C1,,Cr)P(u_{1},...,u_{r};C_{1},...,C_{r}) be a nilprogression in CC-normal form. if thick(P)(P) is large enough depending on rr and CC, then PP is good and the map ujγju_{j}\to\gamma_{j} extends to a product preserving embedding from G(ϵP)G(\epsilon P) to ΓP\Gamma_{P}. And ΓP\Gamma_{P} is isomorphic to the lattice in a rr-dim simply connected nilpotent Lie group 𝒩\mathcal{N}.

Remark 2.24.

During the proof of Theorem A and B, we shall use nilprogressions PP satisfying Theorem 2.23. We may simply identify the grid part G(ϵP)G(\epsilon P) and PP. Then we simply call ΓP\Gamma_{P} in Theorem 2.23 the groupfication of PP.

In [44], Zamora used the structure of approximates groups to study the limit of almost homogeneous spaces. A sequence of geodesic metric spaces ZiZ_{i} is called almost homogeneous if there are discrete isometric group actions GiG_{i} on ZiZ_{i} with diam(Zi/Gi)0(Z_{i}/G_{i})\to 0. Now we assume that (Zi,pi,Gi)eGH(Z,p,G)(Z_{i},p_{i},G_{i})\overset{eGH}{\longrightarrow}(Z,p,G). If we further assume that ZZ is semi-locally simply connected, Zamora proved that GG is a Lie group.

(2.1) (Zi,pi,Gi)eGH(Z,p,G)ππMiGHpt\displaystyle\begin{CD}(Z_{i},p_{i},G_{i})@>{eGH}>{}>(Z,p,G)\\ @V{}V{\pi}V@V{}V{\pi}V\\ M_{i}^{\prime}@>{GH}>{}>\mathrm{pt}\end{CD}
Remark 2.25.

If we further assume that ZiZ_{i} is a RCD((N1),N)(-(N-1),N) space, then the limit ZZ is also a RCD((N1),N)(-(N-1),N) space with a limit measure. Therefore GG is a Lie group and ZZ must be semi-locally simply connected. We may always assume that ZiZ_{i} and ZZ in 2.1 are RCD spaces.

Assume dim(G)=r(G)=r. A small neighborhood of the identity eGe\in G is a strong CC approximate group for some C>0C>0. Thus for any small δ\delta,

Gi(δ)={gGi|d(gx,x)δ,xB¯1δ(pi)}G_{i}(\delta)=\{g\in G_{i}|d(gx,x)\leq\delta,\forall x\in\bar{B}_{\frac{1}{\delta}}(p_{i})\}

is a strong CC approximate group.

We say that GiG_{i} has no small subgroup if there is no non-trivial subgroup of GiG_{i} converging to the identity eGe\in G as ii\to\infty. The next Theorem states that if GiG_{i} has no small subgroup and equivariantly converges to a Lie group, then GiG_{i} contains a large nilprogression which includes a neighborhood of the identity element eGie\in G_{i}.

Theorem 2.26 ([3, 44]).

Assume that GiG_{i} in 2.1 contains no small small subgroup, then for any ϵ>0\epsilon>0 sufficiently small, there exists δ>0\delta>0 independent of the choice of ii, such that Gi(ϵ)G_{i}(\epsilon) contains a nilprogession Pi(u1,,ur;C1,,Cr)P_{i}(u_{1},...,u_{r};C_{1},...,C_{r}) in CC-normal form for some constant C>0C>0 and Gi(δ)G(Pi)G_{i}(\delta)\subset G(P_{i}).

Remark 2.27.

The original statement of Theorem 2.26 in [3, 44] is purely algebraic using ultralimits. It is convenient in this paper to state it geometrically using equivariant convergence.

We first clarify the notations of exponential maps. If GG is a Lie group with Lie algebra 𝔤\mathfrak{g} and a left-invariant metric, we denote by

expG:𝔤G\mathrm{exp}_{G}:\mathfrak{g}\to G

as the Lie group exponential map. For the identity element ee, define the Riemannian exponential map at ee:

expe:TeGG,\mathrm{exp}_{e}:T_{e}G\to G,

where TeGT_{e}G is the tangent space to GG at the identity.

We briefly recall how to construct the nilprogression in Theorem 2.26. Choose ϵ\epsilon small enough so that the set

G(10ϵ)={gG|d(gx,x)10ϵ,xB¯110ϵ(p)}G(10\epsilon)=\{g\in G|d(gx,x)\leq 10\epsilon,\forall x\in\bar{B}_{\frac{1}{10\epsilon}}(p)\}

is connected and expG1\exp^{-1}_{G} from G(10ϵ)G(10\epsilon) to the Lie algebra of GG is diffeomorphic.

Since Gi(ϵ)G_{i}(\epsilon) is a strong approximate group with no small subgroup, in particular, the escape norm is always non-zero. we find the element u1u_{1} with the smallest escape norm. We may assume the diameter of ZiZ_{i} is small enough so that there are generators of GiG_{i} with norm <1/M<1/M, where MM is a constant obtained from Theorem 2.19. Then by (3) in Theorem 2.19, the commutator [u1,g][u_{1},g] is trivial for any gg in the chosen generators; otherwise we get a non-trivial element whose escape norm is strictly less than u1u_{1}’s, a contradiction. In particular, u1u_{1} must lie in the center of Gi(δ)G_{i}(\delta). The group generated by u1u_{1}, u1\langle u_{1}\rangle, converges to an one-parameter subgroup in the center of GG. By taking quotient groups and applying an induction argument, we can construct the nilprogression PiP_{i}.

An important observation from the above construction is that the nilpotent structure of PiP_{i} is determined by the escape norm of Gi(ϵ)G_{i}(\epsilon). We shall use this observation to prove that structure group is affine in Theorem B.

Next recall the structure theory for nilpotent Lie groups and their Lie algebras.

Definition 2.28.

Let 𝔤\mathfrak{g} be a nilpotent Lie algebra. We say that an ordered basis {v1,,vr}\{v_{1},...,v_{r}\} of 𝔤\mathfrak{g} is a strong Malcev basis if for any 1kr1\leq k\leq r, the vector subspace JkJ_{k} generated by {vk+1,,vr}\{v_{k+1},...,v_{r}\} is an ideal, and vk+Jkv_{k}+J_{k} is in the center of 𝔤/Jk\mathfrak{g}/J_{k}.

Theorem 2.29.

Let 𝒩\mathcal{N} be a rr-dim simply connected nilpotent Lie group and 𝔤\mathfrak{g} be its Lie algebra with a strong Malcev basis {v1,,vr}\{v_{1},...,v_{r}\}. Then:
(1) exp𝒩:𝔤𝒩\mathrm{exp}_{\mathcal{N}}:\mathfrak{g}\to\mathcal{N} is a diffeomorphism;
(2) ϕ:r𝒩\phi:\mathbb{R}^{r}\to\mathcal{N} given by ϕ(x1,,xr)=exp𝒩(x1v1)exp𝒩(xrvr)\phi(x_{1},...,x_{r})=\mathrm{exp}_{\mathcal{N}}(x_{1}v_{1})...\mathrm{exp}_{\mathcal{N}}(x_{r}v_{r}) is a diffeomorphism;
(3) if we identify 𝔤\mathfrak{g} with r\mathbb{R}^{r} by the given basis, then exp𝒩1ϕ:𝔤𝔤\mathrm{exp}_{\mathcal{N}}^{-1}\circ\phi:\mathfrak{g}\to\mathfrak{g} and ϕ1exp𝒩i:𝔤𝔤\phi^{-1}\circ\mathrm{exp}_{\mathcal{N}_{i}}:\mathfrak{g}\to\mathfrak{g} are polynomials of degree bounded by a number depending only on rr.

In the diagram 2.1, for sufficiently large ii, by Theorem 2.23 and 2.26, the (grid part of) nilprogression PiP_{i} can be identified as a generating set of a lattice in simply connected nilpotent group 𝒩i\mathcal{N}_{i}. Let gi,j=ujCj/Cg_{i,j}=u_{j}^{\lfloor C_{j}/C\rfloor} and vi,jv_{i,j} in the Lie algebra 𝔤i\mathfrak{g}_{i} of 𝒩i\mathcal{N}_{i} such that

exp𝒩i(vi,j)=gi,j, 1jr.\mathrm{exp}_{\mathcal{N}_{i}}(v_{i,j})=g_{i,j},\ 1\leq j\leq r.

Then {vi,j,1jr}\{v_{i,j},1\leq j\leq r\} is a strong Malcev basis of 𝔤i\mathfrak{g}_{i}. For any jj, passing to a subseqeunce if necessary, assume gi,jgjGg_{i,j}\to g_{j}\in G and choose vjv_{j} in the Lie algebra 𝔤\mathfrak{g} of GG such that expG(vj)=gj\mathrm{exp}_{G}(v_{j})=g_{j}.

For any fixed ii, since {vi,j,1jr}\{v_{i,j},1\leq j\leq r\} is a strong Malcev basis of 𝔤i\mathfrak{g}_{i}, then ϕi:r𝒩i\phi_{i}:\mathbb{R}^{r}\to\mathcal{N}_{i}, as in Theorem 2.29, is a diffeomorphism. Now we identity (x1,x2,,xr)r(x_{1},x_{2},...,x_{r})\in\mathbb{R}^{r} as j=1rxjvi,j𝔤i\sum_{j=1}^{r}x_{j}v_{i,j}\in\mathfrak{g}_{i}, and define

Qi:r×rr,Qi(x,y)=ϕi1(ϕi(x)ϕi(y)).Q_{i}:\mathbb{R}^{r}\times\mathbb{R}^{r}\to\mathbb{R}^{r},\ Q_{i}(x,y)=\phi_{i}^{-1}(\phi_{i}(x)\phi_{i}(y)).

Similarly define QQ for GG and {vj,1jr}\{v_{j},1\leq j\leq r\}. Roughly speaking, QiQ_{i} and QQ decide Lie algebra structure of 𝒩i\mathcal{N}_{i} and GG respectively.

Theorem 2.30 (Lie algebra structure convergence, [44]).

For sufficiently large ii, QiQ_{i} and QQ are all polynomials of degree d(r)\leq d(r) and coefficients of QiQ_{i} converge to corresponding ones of QQ.

3. Constructing a GHA map which is locally almost splitting

In this section, we want to generalize main results in [23] from the smooth case to a weaker version in the RCD case.

Theorem 3.1 (smooth fibration, [23]).

Given n,v>0n,v>0, there exists ϵ(n,v)>0\epsilon(n,v)>0 and C(n)>0C(n)>0 so that consider a compact nn-manifold MM and a kk-manifold KK satisfying:
RicM(n1)\text{Ric}_{M}\geq-(n-1), (1,v)(1,v)-bound covering geometry holds on MM, |secK|1|\text{sec}_{K}|\leq 1, injK1\text{inj}_{K}\geq 1, dGH(M,K)<ϵ<ϵ(n,v)d_{GH}(M,K)<\epsilon<\epsilon(n,v).
Then there is a smooth fiber bundle map f:MKf:M\to K which is a Φ(ϵ|n,v)\Phi(\epsilon|n,v)-GHA, where Φ(ϵ|n,v)0\Phi(\epsilon|n,v)\to 0 as ϵ0\epsilon\to 0.

Assume a group GG isometrically acts two metric spaces X1X_{1} and X2X_{2} separately, we call a map h:X1X2h:X_{1}\to X_{2} 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.

Theorem 3.2 (stability for compact group actions, [23]).

There exists ϵ(n)>0\epsilon(n)>0 so that the following holds for any ϵ<ϵ(n)\epsilon<\epsilon(n). Assume that MM and KK are compact nn-manifolds so that |secK|1|\mathrm{sec}_{K}|\leq 1, injK1\mathrm{inj}_{K}\geq 1, RicM(n1)\mathrm{Ric}_{M}\geq-(n-1). The group GG acts isometrically on MM and KK separately and there is ϵ\epsilon-GHA h:MKh:M\to K which is ϵ\epsilon-almost GG-equivariant. Then there exists a GG-equivariant diffeomorphism f:MKf:M\to K, that is f(gx)=gf(x)f(gx)=gf(x) for any xMx\in M and gGg\in G, which is a Φ(ϵ|n)\Phi(\epsilon|n)-GHA.

We briefly recall the proof of Theorem 3.1 and 3.2. In Theorem 3.1, the injective radius of KK is at least 11. Then locally we can identify a small ball in KK as an open subset in the tangent space of KK at some point. By our assumption that MM is GH close to a manifold KK, locally we can construct almost kk-splitting maps from a small open neighborhood in MM to the tangent space of KK at some point.

To construct a globally-defined map f:MKf:M\to K, we can glue and average these local almost splitting maps using some cut-off functions and the center of mass technique. Then we have a smooth GHA f:MKf:M\to K. Then Huang showed that, at any point pMp\in M, dfdf is the same as the differential of a local almost kk-splitting function, which is constructed by the linear average. Then Huang proved that the differential of any almost kk-splitting map is non-degenerate under (1,v)(1,v)-bound covering condition. Thus dfdf is non-degenerate and ff is a fibration map by the implicit function theorem.

The proof of Theorem 3.2 follows a similar approach. Huang construct a GG-equivariant map using the center of mass and applies the canonical Reifenberg method to show that the map is a diffeomorphism.

In the non-smooth RCD case, we have no implicit function theorem. However, we can prove the following two theorems using ideas from [23]. For any metric space (X,d)(X,d) and pXp\in X, r>0r>0, we use rXrX or (X,rd)(X,rd) for the rescaled metric on XX. Then rB1r(p)rB_{\frac{1}{r}}(p) is actually a unit ball in rXrX.

Theorem 3.3 (Almost kk-splitting).

Assume that a sequence of compact RCD((N1),N)(-(N-1),N) spaces (Xi,di,𝔪)(X_{i},d_{i},\mathfrak{m}) converges to a smooth compact kk-manifold (K,g)(K,g) in the Gromov-Hausdorff sense. Then for sufficiently large ii, there is a continuous GHA fi:XiKf_{i}:X_{i}\to K which is local almost kk-splitting, i.e., for any δ>0\delta>0, piXip_{i}\in X_{i} close to pKp\in K, and ii large enough,

1δBδ(pi)fi1δBδ(p)expp1TpK=k\frac{1}{\delta}B_{\sqrt{\delta}}(p_{i})\overset{f_{i}}{\longrightarrow}\frac{1}{\delta}B_{\sqrt{\delta}}(p)\overset{exp_{p}^{-1}}{\longrightarrow}T_{p}K=\mathbb{R}^{k}

is a (k,δ)(k,\delta)-splitting δ\delta-GHA.

Proof.

Take a small number ϵ<1\epsilon<1. We may assume that the injective radius of KK is at least 10ϵ\frac{10}{\epsilon} and XiX_{i} is RCD((N1)ϵ2,N)(-(N-1)\epsilon^{2},N). We also assume that for any pKp\in K, B10ϵ(p)B_{\frac{10}{\epsilon}}(p) is ϵ2\epsilon^{2}-C4C^{4}-close, by expp1\mathrm{exp}_{p}^{-1}, to its pre-image in TpKT_{p}K with the flat metric. Otherwise we can consider (K,rg)(K,rg) for sufficiently large rr, then (Xi,rdi)(X_{i},rd_{i}) will still converges to (K,rg)(K,rg).

Let Φ(ϵ|k,N)\Phi(\epsilon|k,N) denote a function which converges to 0 as ϵ0\epsilon\to 0, for fixed k,Nk,N. The value of Φ(ϵ|k,N)\Phi(\epsilon|k,N) may vary depending on the specific case. Take any large ii such that dGH(Xi,K)<ϵd_{GH}(X_{i},K)<\epsilon. Our goal is to construct a Φ(ϵ|k,N)\Phi(\epsilon|k,N)-GHA fi:XiKf_{i}:X_{i}\to K which is (k,Φ(ϵ|k,N))(k,\Phi(\epsilon|k,N))-splitting on any 1ϵ\frac{1}{\epsilon} ball. Once this is established, we just take ϵ\epsilon small enough so that Φ(ϵ|k,N)<δ\Phi(\epsilon|k,N)<\delta, thereby completing the proof.

Let {pj}j=1,2,J\{p^{j}\}_{j=1,2,...J} be a 1ϵ\frac{1}{\epsilon}-net in KK and find pijXip_{i}^{j}\in X_{i} ϵ\epsilon-close to pjKp^{j}\in K for each jj. Λj\Lambda_{j} is a cut-off function on KK such that Λj(B1ϵ(pj))=1\Lambda_{j}(B_{\frac{1}{\epsilon}}(p^{j}))=1 and supp(Λj)B2ϵ(pj)\mathrm{supp}(\Lambda_{j})\subset B_{\frac{2}{\epsilon}}(p^{j}). We may assume |lΛj|Φ(ϵ|k,N)|\nabla^{l}\Lambda_{j}|\leq\Phi(\epsilon|k,N), l=1,2,3l=1,2,3. Let Bij=B2ϵ(pij)B_{ij}=B_{\frac{2}{\epsilon}}(p_{i}^{j}), Bj=B2ϵ(pj)B_{j}=B_{\frac{2}{\epsilon}}(p^{j}), Bj1B_{j}^{-1} be the pre-image of BjB_{j} in TpjKT_{p^{j}}K with the flat metric.

BijB_{ij} is ϵ\epsilon-GH close to BjB_{j}, BjB_{j} is ϵ2\epsilon^{2}-C4C^{4}-close to the Bj1B^{-1}_{j} with the flat metric. Take a smaller radius if necessary, we can construct a Φ(ϵ|k,N)\Phi(\epsilon|k,N)-GHA

hj:BijBj,h_{j}:B_{ij}\longrightarrow B_{j},

such that exppj1hj:BijBj1k\mathrm{exp}_{p^{j}}^{-1}\circ h_{j}:B_{ij}\to B_{j}^{-1}\subset\mathbb{R}^{k} is a harmonic (k,Φ(ϵ|k,N))(k,\Phi(\epsilon|k,N))-splitting map.

Take the energy function E:Xi×KE:X_{i}\times K\to\mathbb{R} as follows,

E(x,y)=j=1JΛj(hj(x))d(hj(x),y)2j=1JΛj(hj(x)).E(x,y)=\frac{\sum_{j=1}^{J}\Lambda_{j}(h_{j}(x))d(h_{j}(x),y)^{2}}{\sum_{j=1}^{J}\Lambda_{j}(h_{j}(x))}.

Since supp(Λj)\mathrm{supp}(\Lambda_{j}) is contained in the image of hjh_{j}, Λj(hj(x))\Lambda_{j}(h_{j}(x)) is well-defined for any xXix\in X_{i} by a 0 extension outside of the support. Since hjh_{j} is a GHA, the image of all {hj(x)}j=1,2,,J\{h_{j}(x)\}_{j=1,2,...,J} (if defined) is contained in a Φ(ϵ|k,N)\Phi(\epsilon|k,N)-ball for any fixed xXix\in X_{i}. Let yKy^{\prime}\in K be a point close to xx, then E(x,)E(x,\cdot) is strictly convex in B1ϵ(y)B_{\frac{1}{\epsilon}}(y^{\prime}) and achieve a global minimum at some yB1ϵ(y)y\in B_{\frac{1}{\epsilon}}(y^{\prime}), which is the center of mass with respect to EE. Define fi(x)=yf_{i}(x)=y, then fif_{i} is a Φ(ϵ|k,N)\Phi(\epsilon|k,N)-GHA.

We next show that fif_{i} is (k,Φ(ϵ|k,N))(k,\Phi(\epsilon|k,N))-splitting map on any 1ϵ\frac{1}{\epsilon}-ball. For any x0Xix_{0}\in X_{i}, take y0=f(x0)y_{0}=f(x_{0}). There exists at most C(N)C(N) many points pjp^{j} in the net contained in B4ϵ(y0)B_{\frac{4}{\epsilon}}(y_{0}), saying j1,j2,j3,,jCj_{1},j_{2},j_{3},...,j_{C}. Then the value of fif_{i} on B1ϵ(x0)B_{\frac{1}{\epsilon}}(x_{0}) only depends on hjh_{j} and Λj\Lambda_{j} for j=j1,j2,,jCj=j_{1},j_{2},...,j_{C}.

Consider the energy function on the product space, E~:l=1CBjl×K\tilde{E}:\prod_{l=1}^{C}B_{j_{l}}\times K\to\mathbb{R},

E~(l=1Cyl,y)=l=1CΛjl(yl)d(yl,y)2l=1CΛjl(yl),yK,ylBjl,l=1,2,,C.\tilde{E}(\prod_{l=1}^{C}y_{l},y)=\frac{\sum_{l=1}^{C}\Lambda_{j_{l}}(y_{l})d(y_{l},y)^{2}}{\sum_{l=1}^{C}\Lambda_{j_{l}}(y_{l})},\forall y\in K,y_{l}\in B_{j_{l}},l=1,2,...,C.

For any l=1Cyll=1CBjl\prod_{l=1}^{C}y_{l}\in\prod_{l=1}^{C}B_{j_{l}}, define h~(l=1Cyl)\tilde{h}(\prod_{l=1}^{C}y_{l}) to be the center of mass with respect to E~\tilde{E}. Then by the definition,

fi(x)=h~(hj1(x),hj2(x),,hjC(x)),xB1ϵ(x0).f_{i}(x)=\tilde{h}(h_{j_{1}}(x),h_{j_{2}}(x),...,h_{j_{C}}(x)),\ \forall x\in B_{\frac{1}{\epsilon}}(x_{0}).

Now consider the center of mass on the Euclidean space, which is a linear average. Define

h¯:l=1CBjlTy0K,h¯(l=1Cyl)=l=1CΛjl(yl)expy01(yl)l=1CΛjl(yl).\bar{h}:\prod_{l=1}^{C}B_{j_{l}}\to T_{y_{0}}K,\ \bar{h}(\prod_{l=1}^{C}y_{l})=\frac{\sum_{l=1}^{C}\Lambda_{j_{l}}(y_{l})\mathrm{exp}_{y_{0}}^{-1}(y_{l})}{\sum_{l=1}^{C}\Lambda_{j_{l}}(y_{l})}.

Then expy01h~\mathrm{exp}_{y_{0}}^{-1}\circ\tilde{h} is Φ(ϵ|k,N)\Phi(\epsilon|k,N)-C3C^{3} close to h¯\bar{h}, since the metric on B10ϵ(y0)B_{\frac{10}{\epsilon}}(y_{0}) is ϵ2\epsilon^{2}-C4C^{4} close the a 10ϵ\frac{10}{\epsilon}-ball in Ty0KT_{y_{0}}K with the flat metric; the center of mass in flat k\mathbb{R}^{k} is the linear average.

For any 1lC1\leq l\leq C, exppjl1hjl\mathrm{exp}_{p^{j_{l}}}^{-1}\circ h_{j_{l}} is a harmonic Φ(ϵ|k,N)\Phi(\epsilon|k,N)-GHA and expy01exppjl\mathrm{exp}_{y_{0}}^{-1}\circ\mathrm{exp}_{p^{j_{l}}} is Φ(ϵ|k,N)\Phi(\epsilon|k,N)-C3C^{3}-close to the a linear isometric action on B3ϵ(0k)B_{\frac{3}{\epsilon}}(0^{k}) by our assumption, thus

expy01hjl=expy01exppjlexppjl1hjl\mathrm{exp}_{y_{0}}^{-1}\circ h_{j_{l}}=\mathrm{exp}_{y_{0}}^{-1}\circ\mathrm{exp}_{p^{j_{l}}}\circ\mathrm{exp}_{p^{j_{l}}}^{-1}\circ h_{j_{l}}

is a Φ(ϵ|k,N)\Phi(\epsilon|k,N)-GHA with

|Δ(expy01hjl)|Φ(ϵ|k,N),|Δ(expy01hjl)|Φ(ϵ|k,N)|\Delta(\mathrm{exp}_{y_{0}}^{-1}\circ h_{j_{l}})|\leq\Phi(\epsilon|k,N),|\nabla\Delta(\mathrm{exp}_{y_{0}}^{-1}\circ h_{j_{l}})|\leq\Phi(\epsilon|k,N)

on B1ϵ(y0)B_{\frac{1}{\epsilon}}(y_{0}).

Since |lΛj|Φ(ϵ|k,N)|\nabla^{l}\Lambda_{j}|\leq\Phi(\epsilon|k,N), l=1,2,3l=1,2,3, the linear average

h¯(hj1,hj2,,hjC)=l=1CΛjl(hjl(x))expy01(hjl(x))l=1CΛjl(hjl(x)):B1ϵ(x0)Ty0K\bar{h}\circ(h_{j_{1}},h_{j_{2}},...,h_{j_{C}})=\frac{\sum_{l=1}^{C}\Lambda_{j_{l}}(h_{j_{l}}(x))\mathrm{exp}_{y_{0}}^{-1}(h_{j_{l}}(x))}{\sum_{l=1}^{C}\Lambda_{j_{l}}(h_{j_{l}}(x))}:B_{\frac{1}{\epsilon}}(x_{0})\to T_{y_{0}}K

is a Φ(ϵ|k,N)\Phi(\epsilon|k,N)-GHA with

|Δ(h¯(hj1,hj2,,hjC))|Φ(ϵ|k,N),|Δ(h¯(hj1,hj2,,hjC)|Φ(ϵ|k,N)|\Delta(\bar{h}(h_{j_{1}},h_{j_{2}},...,h_{j_{C}}))|\leq\Phi(\epsilon|k,N),|\nabla\Delta(\bar{h}(h_{j_{1}},h_{j_{2}},...,h_{j_{C}})|\leq\Phi(\epsilon|k,N)

on B1ϵ(y0)B_{\frac{1}{\epsilon}}(y_{0}). In particular, h¯(hj1,hj2,,hjC)\bar{h}(h_{j_{1}},h_{j_{2}},...,h_{j_{C}}) is a (k,Φ(ϵ|k,N))(k,\Phi(\epsilon|k,N))-splitting map on B1ϵ(x0)B_{\frac{1}{\epsilon}}(x_{0}).

Since expy01h~\mathrm{exp}_{y_{0}}^{-1}\circ\tilde{h} is Φ(ϵ|k,N)\Phi(\epsilon|k,N)-C3C^{3}-close to h¯\bar{h}, expy01fi=expy01h~(hj1,hj2,,hjC)\mathrm{exp}_{y_{0}}^{-1}\circ f_{i}=\mathrm{exp}_{y_{0}}^{-1}\circ\tilde{h}(h_{j_{1}},h_{j_{2}},...,h_{j_{C}}) is also a (k,Φ(ϵ|k,N))(k,\Phi(\epsilon|k,N))-splitting map on B1ϵ(x0)B_{\frac{1}{\epsilon}}(x_{0}). ∎

Remark 3.4.

We can also use the embedding argument in [22] to prove Theorem 3.1. By the Nash embedding theorem, we can isometrically embed KK into some n\mathbb{R}^{n} where nn only depends on kk. The embedding map is Φ=(ϕ1,ϕ2,,ϕn):Kn\Phi=(\phi_{1},\phi_{2},...,\phi_{n}):K\hookrightarrow\mathbb{R}^{n}. Let π\pi be the projection map from a neighborhood of KnK\subset\mathbb{R}^{n} to KK. By [22], if ii is large enough, we can construct Ψi=(ψ1i,ψ2i,,ψni):Xin\Psi^{i}=(\psi_{1}^{i},\psi_{2}^{i},...,\psi_{n}^{i}):X_{i}\to\mathbb{R}^{n} so that f=Φ1πΨif=\Phi^{-1}\circ\pi\circ\Psi^{i} is a GHA and Δψji\Delta\psi^{i}_{j} is H1,2H^{1,2} close to Δϕi\Delta\phi_{i} for any 1jn1\leq j\leq n. Then ff is locally almost kk-splitting if we sufficiently blow up the metric.

We next state a GG-stability result in the RCD setting.

Theorem 3.5 (GG-equivariant).

There exists ϵ(N)\epsilon(N) such that for any 0<ϵϵ(N)0<\epsilon\leq\epsilon(N), the following holds. For any compact RCD((N1)ϵ2,N)(-(N-1)\epsilon^{2},N) space (X,d,𝔪)(X,d,\mathfrak{m}) and kk-manifold KK satisfying that injK10ϵ\mathrm{inj}_{K}\geq\frac{10}{\epsilon} and for any pKp\in K, B10ϵ(p)B_{\frac{10}{\epsilon}}(p) is ϵ2\epsilon^{2}-C4C^{4}-close, by expp1\mathrm{exp}_{p}^{-1}, to its preimage in the tangent space TpKT_{p}K with the flat metric. Assume that a map h:XKh:X\to K is an ϵ\epsilon-GHA map on any 10ϵ\frac{10}{\epsilon}-ball in XX, that is for any xXx\in X, h:B10ϵ(x)Kh:B_{\frac{10}{\epsilon}}(x)\to K is an ϵ\epsilon-GHA to its image. A finite group GG acts isometrically on XX and KK separately and hh is ϵ\epsilon-almost GG-equivariant. Then there is a GG-equivariant map fG:XKf_{G}:X\to K, which is also a Φ(ϵ|k,N)\Phi(\epsilon|k,N)-GHA and locally (k,Φ(ϵ|k,N))(k,\Phi(\epsilon|k,N))-splitting on any 1ϵ\frac{1}{\epsilon}-ball.

Remark 3.6.

In Theorem 3.5, there is no control of the diameter of XX or the order |G||G|. Instead we only need a global map hh and the gluing is local. If we further assume k=Nk=N and ϵ\epsilon small enough, then fGf_{G} is biHölder on any 1ϵ\frac{1}{\epsilon}-ball due to Theorem 2.7.

Proof.

We can use the same proof of Theorem 3.3 to construct a map f:XKf:X\to K which is (k,Φ(ϵ|k,N))(k,\Phi(\epsilon|k,N))-splitting and a Φ(ϵ|k,N)\Phi(\epsilon|k,N)-GHA on any 1ϵ\frac{1}{\epsilon}-ball. Notice that the condition that h:XKh:X\to K is an ϵ\epsilon-GHA map on any 10ϵ\frac{10}{\epsilon}-ball is enough for the construction in Theorem 3.3 as the gluing and averaging procedure is local.

Since f(x)f(x) is close to h(x)h(x) for any xXx\in X, we have d(f(gx),gf(x))Φ(ϵ|k,N)d(f(gx),gf(x))\leq\Phi(\epsilon|k,N) for any xXx\in X and gGg\in G. In particular, for any gGg\in G,

gf(g1x):XKgf(g^{-1}x):X\to K

is a Φ(ϵ|k,N)\Phi(\epsilon|k,N)-GHA.

Since ff is (k,Φ(ϵ|k,N))(k,\Phi(\epsilon|k,N))-splitting on any 1ϵ\frac{1}{\epsilon}-ball and GG-actions are isometric on XX and KK, thus for any gGg\in G, the map gf(g1x):XKgf(g^{-1}x):X\to K is also (k,Φ(ϵ|k,N))(k,\Phi(\epsilon|k,N))-splitting on any 1ϵ\frac{1}{\epsilon}-ball.

Now we average GG actions by the center of mass. Take the energy function

E(x,y):X×K,E(x,y)=gGd(gf(g1x),y)2|G|.E(x,y):X\times K\to\mathbb{R},\ E(x,y)=\frac{\sum_{g\in G}d(gf(g^{-1}x),y)^{2}}{|G|}.

For any fixed xXx\in X, E(x,)E(x,\cdot) in strictly convex in B1ϵ(f(x))B_{\frac{1}{\epsilon}}(f(x)) thus there is a global minimum point yy. Define fG(x)=yf_{G}(x)=y, then fGf_{G} is GG-equivariant due to the uniqueness of the minimal point. fGf_{G} is also a Φ(ϵ|k,N)\Phi(\epsilon|k,N)-GHA.

We next show that fGf_{G} is (k,Φ(ϵ|k,N))(k,\Phi(\epsilon|k,N))-splitting on every 1ϵ\frac{1}{\epsilon}-ball by a similar argument in the proof of Theorem 3.3. For any x0Xx_{0}\in X and y0=fG(x0)y_{0}=f_{G}(x_{0}). Define the energy function on the product space, E~:l=1|G|B1ϵ(y0)×K\tilde{E}:\prod_{l=1}^{|G|}B_{\frac{1}{\epsilon}}(y_{0})\times K\to\mathbb{R} by

E~(l=1|G|yl,y)=gGd(yl,y)2|G|,yK,ylB1ϵ(y0),l=1,,|G|.\tilde{E}(\prod_{l=1}^{|G|}y_{l},y)=\frac{\sum_{g\in G}d(y_{l},y)^{2}}{|G|},y\in K,y_{l}\in B_{\frac{1}{\epsilon}}(y_{0}),l=1,...,|G|.

Then define h~(l=1|G|yl)\tilde{h}(\prod_{l=1}^{|G|}y_{l}) to be the center of mass with respect to E~\tilde{E}. Then fG(x)=h~(gGgf(g1x))f_{G}(x)=\tilde{h}(\prod_{g\in G}gf(g^{-1}x)) by the definition.

Now consider the center of mass on the Euclidean space, which is a linear average. Define h¯:l=1|G|B1ϵ(y0)Ty0K\bar{h}:\prod_{l=1}^{|G|}B_{\frac{1}{\epsilon}}(y_{0})\to T_{y_{0}}K by h¯(l=1|G|yl)=l=1|G|expy01(yl)|G|\bar{h}(\prod_{l=1}^{|G|}y_{l})=\frac{\sum_{l=1}^{|G|}\mathrm{exp}_{y_{0}}^{-1}(y_{l})}{|G|}. Then expy01h~\mathrm{exp}_{y_{0}}^{-1}\circ\tilde{h} is Φ(ϵ|k,N)\Phi(\epsilon|k,N) -C3C^{3}-close to h¯\bar{h} as the metric on B10ϵ(y0)B_{\frac{10}{\epsilon}}(y_{0}) is ϵ2\epsilon^{2}-C4C^{4}-close the a B10ϵB_{\frac{10}{\epsilon}} ball in Ty0KT_{y_{0}}K with the flat metric.

The linear average h¯(gGgf(g1x)):B1ϵ(x0)Ty0K\bar{h}(\prod_{g\in G}gf(g^{-1}x)):B_{\frac{1}{\epsilon}}(x_{0})\to T_{y_{0}}K is (k,Φ(ϵ|k,N))(k,\Phi(\epsilon|k,N))-splitting by a similar argument as in Theorem 3.3, therefore

expy01fG=expy01h~(gGgf(g1x))\mathrm{exp}_{y_{0}}^{-1}\circ f_{G}=\mathrm{exp}_{y_{0}}^{-1}\circ\tilde{h}(\prod_{g\in G}gf(g^{-1}x))

is also (k,Φ(ϵ|k,N))(k,\Phi(\epsilon|k,N))-splitting on B1ϵ(x0)B_{\frac{1}{\epsilon}}(x_{0}). ∎

4. Proof of Theorem A: construct an infranil-manifold

We prove Theorem A in this section.

Theorem A.

Given N,v>0N,v>0, there exists ϵ(N,v)>0\epsilon(N,v)>0 and C(N)>0C(N)>0, so that if a RCD((N1),N)(-(N-1),N) space (X,d,N)(X,d,\mathcal{H}^{N}) satisfies:

diam(X)<ϵϵ(N,v),N(B1(p~))v,p~X~,\mathrm{diam}(X)<\epsilon\leq\epsilon(N,v),\ \mathcal{H}^{N}(B_{1}(\tilde{p}))\geq v,\ \forall\tilde{p}\in\widetilde{X},

then XX is biHölder homeomorphic to an infranil-manifold 𝒩/Γ\mathcal{N}/\Gamma where 𝒩\mathcal{N} has a left-invariant metric, i.e., there exists f:X𝒩/Γf:X\to\mathcal{N}/\Gamma with

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

where Φ(ϵ|N,v)0\Phi(\epsilon|N,v)\to 0 as ϵ0\epsilon\to 0. Moreover, if XX is a smooth NN-manifold with Ric(N1)\mathrm{Ric}\geq-(N-1), then ff is a diffeomorphism.

Assume that Theorem A does not hold. Then we can find a sequence of RCD((N1),N)(-(N-1),N) space (Xi,di,N)(X_{i},d_{i},\mathcal{H}^{N}) with non-collapsing universal covers and diam(Xi)0\mathrm{diam}(X_{i})\to 0, while these spaces are not biHölder to any infranil-manifold.

Blow up the sequence slowly and pass to a subsequence if necessary, we may assume that the universal covers X~i\widetilde{X}_{i} of XiX_{i} converge to N\mathbb{R}^{N}. By Theorem 2.4, for all sufficiently large ii, Gi=π1(Xi,pi)G_{i}^{\prime}=\pi_{1}(X_{i},p_{i}) contains a normal nilpotent subgroup GiG_{i} of index C(N)\leq C(N). Let Xi=X~i/GiX_{i}^{\prime}=\widetilde{X}_{i}/G_{i}.

(X~i,p~i,Gi,Gi)eGH(N,p~,G,G)ππ(Xi,pi,Gi/Gi)GHptππ(Xi,pi)GHpt\begin{CD}(\widetilde{X}_{i},\tilde{p}_{i},G_{i},G_{i}^{\prime})@>{eGH}>{}>(\mathbb{R}^{N},\tilde{p},G,G^{\prime})\\ @V{}V{\pi}V@V{}V{\pi}V\\ (X_{i}^{\prime},p_{i}^{\prime},G_{i}^{\prime}/G_{i})@>{GH}>{}>\mathrm{pt}\\ @V{}V{\pi}V@V{}V{\pi}V\\ (X_{i},p_{i})@>{GH}>{}>\mathrm{pt}\end{CD}

XiX_{i}^{\prime} converges to a point as it is a finite cover of XiX_{i} with order C\leq C, thus diam(Xi)0\mathrm{diam}(X_{i}^{\prime})\to 0.

Lemma 4.1.

GiG_{i} contains no small subgroup, i.e., there exists δ>0\delta>0 such that the set

Gi(δ)={gGi|d(x,gx)<δ,xB1/δ(p~i)}G_{i}(\delta)=\{g\in G_{i}^{\prime}|d(x,gx)<\delta,x\in B_{1/\delta}(\tilde{p}_{i})\}

contains no non-trivial subgroup for all large ii.

Proof.

Otherwise assume that HiH_{i} is a non-trivial subgroup in Gi(1/i)G_{i}(1/i). Then HiH_{i} converges to the identity, therefore X~i/Hi\widetilde{X}_{i}/H_{i} converges to N\mathbb{R}^{N} as well. By volume convergence theorem, the volume of 11-ball at π(p~)\pi(\tilde{p}) in X~i/Hi\widetilde{X}_{i}/H_{i} is close to the volume of 11-ball at p~\tilde{p} in X~i\widetilde{X}_{i}, a contradiction since HiH_{i} is a small non-trivial subgroup. ∎

Lemma 4.2.

GG is free. In particular, GG can be identified as N\mathbb{R}^{N}.

Proof.

Consider any isotropy group

Gx={gG|gx=x}G_{x}=\{g\in G|gx=x\}

where xNx\in\mathbb{R}^{N}. Then GxG_{x} is compact. Since GG is a nilpotent Lie group, thus [Gx,G0][G_{x},G_{0}] is trivial, where G0G_{0} is the identity component of GG. Since GG is transitive, then G0G_{0} is also transitive. Therefore GxG_{x} fixes all points, thus is a trivial group. ∎

GG has no isotropy. Thus if any gGig\in G_{i} which moves p~i\tilde{p}_{i} small, gg must be close to the identity action.

Corollary 4.3.

For any δ>0\delta>0, there exists ϵ>0\epsilon>0 such that Gi(p~i,ϵ)Gi(δ)G_{i}(\tilde{p}_{i},\epsilon)\subset G_{i}(\delta) for sufficiently large ii.

Lemma 4.4.

For sufficiently large ii, GiG_{i} is isomorphic to a lattice in a NN-dim simply connected nilpotent Lie group 𝒩i\mathcal{N}_{i}.

Proof.

GiG_{i} contains no small subgroup and converges to N\mathbb{R}^{N}. By Theorem 2.26, GiG_{i} contains a nilprogession PiP_{i} of dimension NN, which contains Gi(δ)G_{i}(\delta) for some δ>0\delta>0. By Corollary 4.3, the nilprogession PiP_{i} contains Gi(p~i,ϵ)G_{i}(\tilde{p}_{i},\epsilon) for some ϵ>0\epsilon>0. We choose ii large enough such that diam(Xi)<ϵ20(X_{i})<\frac{\epsilon}{20}, then the groupfication of the PiP_{i} is isomorphic to GiG_{i} by Theorem 2.8. On the other hand, for ii large enough, thick(Pi)(P_{i}) is large enough so that Theorem 2.23 holds, thus the groupfication of the PiP_{i} is a lattice in a simply connected NN-dim nilpotent group 𝒩i\mathcal{N}_{i}. Thus GiG_{i} is isomorphic to a lattice in 𝒩i\mathcal{N}_{i}. ∎

We shall use Theorem 2.30 to find a left-invariant metric on 𝒩i\mathcal{N}_{i} so that it is locally C4C^{4}-close to flat N\mathbb{R}^{N}.

Lemma 4.5.

For any ϵ1\epsilon\leq 1 and ii large enough, 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 exists 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 the tangent space Tg𝒩iT_{g}\mathcal{N}_{i} with the flat metric.

Proof.

We always assume ϵi\epsilon_{i} to be a sequence of numbers converging to 0 while the value of ϵi\epsilon_{i} depends on the specific setting. Consider the Lie algebra structure on 𝒩i\mathcal{N}_{i} using the same notations in the argument after Theorem 2.26. Take {vi,j,1jN}\{v_{i,j},1\leq j\leq N\} as a strong Malcev basis of the Lie algebra of 𝒩i\mathcal{N}_{i}, then exp𝒩i(vi,j)=gi,j𝒩iGi\mathrm{exp}_{\mathcal{N}_{i}}(v_{i,j})=g_{i,j}\in\mathcal{N}_{i}\cap G_{i}. By our assumption, gi,jg_{i,j} action on X~i\widetilde{X}_{i} pointed equivariantly converges to gjG=Ng_{j}\in G=\mathbb{R}^{N} for any 1jN1\leq j\leq N and ii\to\infty. {gj}j=1,2,N\{g_{j}\}_{j=1,2...,N} is a basis of N\mathbb{R}^{N}. We can take the corresponding Lie algebra vj=gjv_{j}=g_{j} since G=NG=\mathbb{R}^{N} is abelian. Define the left-invariant metric g𝒩ig_{\mathcal{N}_{i}} by

g𝒩i(vi,j1,vi,j2)=vj1,vj2,g_{\mathcal{N}_{i}}(v_{i,j_{1}},v_{i,j_{2}})=\langle v_{j_{1}},v_{j_{2}}\rangle,

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

Since {vi,j,1jn}\{v_{i,j},1\leq j\leq n\} is a strong Malcev basis, for any 1j1<j2N1\leq j_{1}<j_{2}\leq N,

[vi,j1,vi,j2]=j=j2+1nai,j1j2jvi,j.[v_{i,j_{1}},v_{i,j_{2}}]=\sum_{j=j_{2}+1}^{n}a_{i,j_{1}j_{2}}^{j}v_{i,j}.

ai,j1j2j0a_{i,j_{1}j_{2}}^{j}\to 0 as i0i\to 0 by Theorem 2.30 and the fact that the limit group is abelian. Define ai,j1j2j=0a_{i,j_{1}j_{2}}^{j}=0 if jj1j\leq j_{1} or jj2j\leq j_{2}. Then by Koszul’s formula, for any 1j1,j2,j3N1\leq j_{1},j_{2},j_{3}\leq N,

g𝒩i(vi,j1vi,j2,vi,j3)=12(ai,j1j2j3ai,j2j3j1+ai,j3j1j2).g_{\mathcal{N}_{i}}(\nabla_{v_{i,j_{1}}}v_{i,j_{2}},v_{i,j_{3}})=\frac{1}{2}(a_{i,j_{1}j_{2}}^{j_{3}}-a_{i,j_{2}j_{3}}^{j_{1}}+a_{i,j_{3}j_{1}}^{j_{2}}).

Since 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)kRg𝒩i|ϵi, 0k3|(\nabla^{g_{\mathcal{N}_{i}}})^{k}R_{g_{\mathcal{N}_{i}}}|\leq\epsilon_{i},\ 0\leq k\leq 3

where ϵi0\epsilon_{i}\to 0.

The sectional curvature of 𝒩i\mathcal{N}_{i} is bounded by ϵi0\epsilon_{i}\to 0. By Theorem 2.29, 𝒩i\mathcal{N}_{i} is diffeomorphiic to N\mathbb{R}^{N}. By our construction of the metric, B4ϵ(e)𝒩iB_{\frac{4}{\epsilon}}(e)\subset\mathcal{N}_{i} must be biLipschitz on B4ϵ(0N)B_{\frac{4}{\epsilon}}(0^{N}). Therefore the injective radius of 𝒩i\mathcal{N}_{i} is at least 1ϵ\frac{1}{\epsilon} for sufficiently large ii. The C4C^{4}-closeness follows from the fact that |(g𝒩i)kRg𝒩i|ϵi|(\nabla^{g_{\mathcal{N}_{i}}})^{k}R_{g_{\mathcal{N}_{i}}}|\leq\epsilon_{i}, 0k30\leq k\leq 3. ∎

From now on we always assume that 𝒩i\mathcal{N}_{i} has the metric g𝒩ig_{\mathcal{N}_{i}} constructed in Lemma 4.5.

Lemma 4.6 (Local eGH closeness).

For any ϵ>0\epsilon>0, let B1ϵ(p~i)X~iB_{\frac{1}{\epsilon}}(\tilde{p}_{i})\subset\widetilde{X}_{i} and B1ϵ(e)𝒩iB_{\frac{1}{\epsilon}}(e)\subset\mathcal{N}_{i}. Then there exists an ϵi\epsilon_{i}-GHA hi:B1ϵ(p~i)B1ϵ(e)h_{i}^{\prime}:B_{\frac{1}{\epsilon}}(\tilde{p}_{i})\to B_{\frac{1}{\epsilon}}(e) which is ϵi\epsilon_{i}-almost Gi(p~i,1ϵ)G_{i}(\tilde{p}_{i},{\frac{1}{\epsilon}})-equivariant if it is well-defined, where ϵi0\epsilon_{i}\to 0 as i0i\to 0.

Proof.

Take a linear map

ψi:NN,ϕ(x1,x2,,xN)=j=1Nxjvi,j.\psi_{i}:\mathbb{R}^{N}\to\mathbb{R}^{N},\ \phi(x_{1},x_{2},...,x_{N})=\sum_{j=1}^{N}x_{j}v_{i,j}.

Take ϕi:N𝒩i\phi_{i}:\mathbb{R}^{N}\to\mathcal{N}_{i} as in Theorem 2.29 (2). By Theorem 2.30 and the definition of the metric in Lemma 4.5,

ϕiψi1:N𝒩i\phi_{i}\circ\psi_{i}^{-1}:\mathbb{R}^{N}\to\mathcal{N}_{i}

is a ϵ\epsilon-GHA on B1ϵ(0N)B_{\frac{1}{\epsilon}}(0^{N}) for sufficiently large ii.

(X~i,p~i)(\widetilde{X}_{i},\tilde{p}_{i}) is pGH-close to N\mathbb{R}^{N} by our assumption, and ϕiψi1:N𝒩i\phi_{i}\circ\psi_{i}^{-1}:\mathbb{R}^{N}\to\mathcal{N}_{i} is a GHA on the 1ϵ\frac{1}{\epsilon}-ball for all large ii. Combine two GHAs, we get hi:B1ϵ(p~i)B1ϵ(e)h_{i}^{\prime}:B_{\frac{1}{\epsilon}}(\tilde{p}_{i})\to B_{\frac{1}{\epsilon}}(e) which is a ϵi\epsilon_{i}-GHA. We need to check hih_{i}^{\prime} is almost Gi(p~i,1ϵ)G_{i}(\tilde{p}_{i},{\frac{1}{\epsilon}})-equivariant.

By our construction of gi,jg_{i,j} (see the argument after Theorem 2.26), exp(vi,j)=gi,j\mathrm{exp}(v_{i,j})=g_{i,j} action on X~i\tilde{X}_{i} is ϵi\epsilon_{i}-close to gjG=Ng_{j}\in G=\mathbb{R}^{N}, 1jN1\leq j\leq N. On the other hand, under the map ϕiψi1:N𝒩i\phi_{i}\circ\psi_{i}^{-1}:\mathbb{R}^{N}\to\mathcal{N}_{i}, gi,jg_{i,j} action on 𝒩i\mathcal{N}_{i} is ϵi\epsilon_{i}-close to the gjG=Ng_{j}\in G=\mathbb{R}^{N} action. Thus gi,jg_{i,j} actions on B1ϵ(p~i)B_{\frac{1}{\epsilon}}(\tilde{p}_{i}) and B1ϵ(e)B_{\frac{1}{\epsilon}}(e) are ϵi\epsilon_{i}-close to each other, that is, hi(gi,jx)h_{i}^{\prime}(g_{i,j}x) is ϵi\epsilon_{i}-close to gi,jhi(x)g_{i,j}h_{i}^{\prime}(x) for all xB1ϵ(p~i)x\in B_{\frac{1}{\epsilon}}(\tilde{p}_{i}) such that gi,jxB1ϵ(p~i)g_{i,j}x\in B_{\frac{1}{\epsilon}}(\tilde{p}_{i}), 1jN1\leq j\leq N.

Take C0C_{0} such that |ψi1(B1ϵ(0N))|C02|\psi_{i}^{-1}(B_{\frac{1}{\epsilon}}(0^{N}))|\leq\frac{C_{0}}{2}. Fix jj, recall gi,j=ui,jljg_{i,j}=u_{i,j}^{l_{j}} where ui,ju_{i,j} is one of the generators of the nilprogression and lj=Cj/Cl_{j}=\lfloor C_{j}/C\rfloor. Since the group actions ui,j\langle u_{i,j}\rangle on both spaces X~i\widetilde{X}_{i} or 𝒩i\mathcal{N}_{i} converge to a line in G=NG=\mathbb{R}^{N}. Then for any kC0k\leq C_{0}, ui,jku_{i,j}^{k} action on B1ϵ(p~i)B_{\frac{1}{\epsilon}}(\tilde{p}_{i}) and B1ϵ(e)B_{\frac{1}{\epsilon}}(e) are both ϵi\epsilon_{i}-close to gjk/ljNg_{j}^{k/l_{j}}\in\mathbb{R}^{N} action.

For a general g=ui,1k1ui,NkNGi(p~i,1ϵ)g=u_{i,1}^{k_{1}}...u_{i,N}^{k_{N}}\in G_{i}(\tilde{p}_{i},\frac{1}{\epsilon}), we have kjC0k_{j}\leq C_{0}, 1jN1\leq j\leq N. Then any component of gg actions, saying ui,1k1u_{i,1}^{k_{1}}, on B1ϵ(p~i)B_{\frac{1}{\epsilon}}(\tilde{p}_{i}) and B1ϵ(e)B_{\frac{1}{\epsilon}}(e) are ϵi\epsilon_{i}-close to each other. Therefore any gGi(p~i,1ϵ)g\in G_{i}(\tilde{p}_{i},\frac{1}{\epsilon}) actions on B1ϵ(p~i)B_{\frac{1}{\epsilon}}(\tilde{p}_{i}) and B1ϵ(e)B_{\frac{1}{\epsilon}}(e) are ϵi\epsilon_{i}-close to each other. ∎

By Lemma 4.6, we have a local GHA hih_{i}^{\prime} defined from B1ϵ(p~i)B_{\frac{1}{\epsilon}}(\tilde{p}_{i}) to B1ϵ(e)𝒩iB_{\frac{1}{\epsilon}}(e)\subset\mathcal{N}_{i}, which is almost Gi(pi~,1ϵ)G_{i}(\tilde{p_{i}},\frac{1}{\epsilon})-equivariant. We can extend hih_{i}^{\prime} to a global map by the follows. For any xX~ix\in\widetilde{X}_{i}, choose gGig\in G_{i} so that gxB1(p~i)gx\in B_{1}(\tilde{p}_{i}). This choice is valid since X~/Gi\widetilde{X}/G_{i} has a small diameter. Define hih_{i} by hi(x)=g1h(gx)𝒩ih_{i}(x)=g^{-1}h^{\prime}(gx)\in\mathcal{N}_{i}. Note that different choices of gg yield only minor differences in hi(x)h_{i}(x), so we can select one for our construction without loss of generality.

Lemma 4.7 (Extend a local approximation to the global map, [40]).

The map hi:X~i𝒩ih_{i}:\widetilde{X}_{i}\to\mathcal{N}_{i} is a global map which is an ϵi\epsilon_{i}-GHA on any 1ϵ\frac{1}{\epsilon}-ball and ϵi\epsilon_{i}-almost GiG_{i}-equivariant, for some ϵi0\epsilon_{i}\to 0.

It is well known that a subgroup of finite index contains a normal subgroup of finite index. We provide a proof for readers’ convenience.

Lemma 4.8.

Assume that group HH contains a subgroup H0H_{0} of index |H:H0|C|H:H_{0}|\leq C, then there is normal subgroup HH^{\prime} of HH such that |H/H|C!|H/H^{\prime}|\leq C! and HH0H^{\prime}\subset H_{0}.

Proof.

Define H/H0H/H_{0} be the set of left cosets of H0H_{0} in HH, the index is C\leq C. Define a homomorphism ϕ:Hsym(H/H0)\phi:H\to\mathrm{sym}(H/H_{0}) by ϕ(g,gH0)=(gg)H0\phi(g,g^{\prime}H_{0})=(gg^{\prime})H_{0}. Then define H=Ker(ϕ)H^{\prime}=\mathrm{Ker}(\phi), which is a normal subgroup of HH with index less than or equal to the order of the symmetric group sym(H/H0)\mathrm{sym}(H/H_{0}), thus |H/H|C!|H/H^{\prime}|\leq C!.

For any gHg\in H^{\prime}, since ϕ(g,H0)=H0\phi(g,H_{0})=H_{0}, it follows that gH0g\in H_{0}. Then HH0H^{\prime}\subset H_{0}. ∎

Now we can complete the proof of Theorem A.

Proof of Theorem A.

Consider a contradiction sequence,

(X~i,p~i,Gi,Gi)eGH(N,p~,G,G)ππ(Xi,pi,Gi/Gi)GHptππ(Xi,pi)GHpt.\begin{CD}(\widetilde{X}_{i},\tilde{p}_{i},G_{i},G_{i}^{\prime})@>{eGH}>{}>(\mathbb{R}^{N},\tilde{p},G,G^{\prime})\\ @V{}V{\pi}V@V{}V{\pi}V\\ (X_{i}^{\prime},p_{i}^{\prime},G_{i}^{\prime}/G_{i})@>{GH}>{}>\mathrm{pt}\\ @V{}V{\pi}V@V{}V{\pi}V\\ (X_{i},p_{i})@>{GH}>{}>\mathrm{pt}.\end{CD}

We assumed that none of XiX_{i} is biHölder to an infranil-manifold.

We have established that GiG_{i} acts as lattice in a simply connected nilpotent Lie group 𝒩i\mathcal{N}_{i}. Since GiG_{i} is a normal subgroup of GiG_{i}^{\prime} with bounded index, by the rigidity result from [30], GiG_{i}^{\prime} can be identified as a discrete subset of 𝒩iAut(𝒩i)\mathcal{N}_{i}\rtimes\text{Aut}(\mathcal{N}_{i}). Since the lattice GiG_{i} is ϵi\epsilon_{i}-dense in 𝒩i\mathcal{N}_{i}, hih_{i} in Lemma 4.7 is also ϵi\epsilon_{i}-almost GiG_{i}^{\prime}-equivariant by the rigidity.

Since g𝒩ig_{\mathcal{N}_{i}} is left-invariant metric on 𝒩i\mathcal{N}_{i}, GiG_{i} actions on 𝒩i\mathcal{N}_{i} are isometric. For any g=(g,ϕ)=Gi𝒩iAut(𝒩i)g^{\prime}=(g,\phi)=G_{i}^{\prime}\subset\mathcal{N}_{i}\rtimes\text{Aut}(\mathcal{N}_{i}) where g𝒩ig\in\mathcal{N}_{i} and ϕAut(𝒩i)\phi\in\text{Aut}(\mathcal{N}_{i}), gg^{\prime} action on either (X~i,p~i)(\widetilde{X}_{i},\tilde{p}_{i}) or (𝒩i,e)(\mathcal{N}_{i},e) is eGH close to an isometric action on N\mathbb{R}^{N}. Since (𝒩i,e)(\mathcal{N}_{i},e) pointed C5C^{5}-close to flat N\mathbb{R}^{N}, ϕ(g𝒩i)\phi^{*}(g_{\mathcal{N}_{i}}) must be C4C^{4} close to g𝒩ig_{\mathcal{N}_{i}}. We have |Gi/Gi|C|G_{i}^{\prime}/G_{i}|\leq C, thus the number of the choices of ϕ\phi is at most CC. Averaging all such ϕ(g𝒩i)\phi^{*}(g_{\mathcal{N}_{i}}) if necessary, we may simply assume that ϕ(g𝒩i)=g𝒩i\phi^{*}(g_{\mathcal{N}_{i}})=g_{\mathcal{N}_{i}}, then GiG_{i}^{\prime} actions on 𝒩i\mathcal{N}_{i} are isometric.

Take any small ϵ>0\epsilon>0. The finitely generated nilpotent group GiG_{i} is residually finite, i.e., for any gGig\in G_{i}, there exists a finite index normal subgroup of GiG_{i} which does not contain gg. Therefore there is a normal subgroup Gi′′G_{i}^{\prime\prime} of GiG_{i} with finite index so that Gi′′B1ϵ(e)=G_{i}^{\prime\prime}\cap B_{\frac{1}{\epsilon}}(e)=\emptyset. We may also assume that Gi′′G_{i}^{\prime\prime} is normal in GiG_{i}^{\prime} as well due to Lemma 4.8.

Since Gi′′B1ϵ(e)=G_{i}^{\prime\prime}\cap B_{\frac{1}{\epsilon}}(e)=\emptyset, we can apply Lemma 4.5 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}^{\prime}-equivariant, we can reduce hih_{i} to a map

h¯i:X~i/Gi′′𝒩i/Gi′′,\bar{h}_{i}:\widetilde{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}^{\prime}/G_{i}^{\prime\prime}-equivariant. For any xX~i/Gi′′x\in\widetilde{X}_{i}/G_{i}^{\prime\prime}, define h¯i(x)=π(hi(x~))𝒩i/Gi′′\bar{h}_{i}(x)=\pi(h_{i}(\tilde{x}))\in\mathcal{N}_{i}/G_{i}^{\prime\prime}, where x~\tilde{x} is a pre-image of xx in X~i\widetilde{X}_{i}. Different choices of x~\tilde{x} lead to only minor differences in h¯i\bar{h}_{i} since hih_{i} is almost GiG_{i}^{\prime}-equivariant.

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

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

We can find a (Gi/Gi′′)(G_{i}^{\prime}/G_{i}^{\prime\prime})-equivariant map fGi/Gi′′f_{G_{i}^{\prime}/G_{i}^{\prime\prime}} from X~i/Gi′′\widetilde{X}_{i}/G_{i}^{\prime\prime} to 𝒩i/Gi′′\mathcal{N}_{i}/G_{i}^{\prime\prime} which is (N,Φ(ϵ|N))(N,\Phi(\epsilon|N))-splitting on any 110ϵ\frac{1}{10\epsilon}-ball. Then we have biHölder estimates from [22],

(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}^{\prime}/G_{i}^{\prime\prime}}(x),f_{G_{i}^{\prime}/G_{i}^{\prime\prime}}(y))\leq(1+\Phi(\epsilon|N))d_{i}(x,y),

for any x,yX~i/Gi′′x,y\in\widetilde{X}_{i}/G_{i}^{\prime\prime} with di(x,y)120ϵd_{i}(x,y)\leq\frac{1}{20\epsilon}.

Since fGi/Gi′′f_{G_{i}^{\prime}/G_{i}^{\prime\prime}} is (Gi/Gi′′)(G_{i}^{\prime}/G_{i}^{\prime\prime})-equivariant, it can be naturally reduced to a biHölder map on the quotient space f:Xi=X~i/Gi𝒩i/Gif:X_{i}=\widetilde{X}_{i}/G_{i}^{\prime}\to\mathcal{N}_{i}/G_{i}^{\prime}. Thus XiX_{i} is biHölder to an infranil-manifold, a contradiction to the assumption. Moreover, if XiX_{i} is a smooth manifold with Ric(N1)\mathrm{Ric}\geq-(N-1), then ff is smooth and dfdf is non-degenerate, thus XiX_{i} is biHölder diffeomorphic to 𝒩i/Gi\mathcal{N}_{i}/G_{i}^{\prime}. ∎

Remark 4.9.

At the beginning of the proof of Theorem A, we slowly blow up the metric to get a limit space N\mathbb{R}^{N}. Therefore, for a contradiction sequence (Xi,di,N)(X_{i},d_{i},\mathcal{H}^{N}), we actually proved the biHölder estimate for (Xi,ridi,N)(X_{i},r_{i}d_{i},\mathcal{H}^{N}), where rir_{i}\to\infty slowly,

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

where dd is the distance function on 𝒩i\mathcal{N}_{i}. Fix a large ii, we can take d=drid^{\prime}=\frac{d}{r_{i}} on 𝒩i\mathcal{N}_{i}. Since

(1Φ(ϵ|N))(ridi(x,y))1+Φ(ϵ|N)ri(1Φ(ϵ|N))(di(x,y))1+Φ(ϵ|N),(1-\Phi(\epsilon|N))(r_{i}d_{i}(x,y))^{1+\Phi(\epsilon|N)}\geq r_{i}(1-\Phi(\epsilon|N))(d_{i}(x,y))^{1+\Phi(\epsilon|N)},

we have

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

Then f:(Xi,di)(𝒩i,d)f:(X_{i},d_{i})\to(\mathcal{N}_{i},d^{\prime}) is also biHölder. Thus the biHölder estimate on the blowing up metric implies the biHölder estimate on the original metric. For this reason, in the next sections we shall omit some blowing up arguments and directly apply Theorem 3.5 and canonical Reifenberg method; see the proofs of Lemma 6.3 and Theorem B.

5. Mixed curvature and almost flat manifolds theorem

In this section we prove Theorem 1.7 using a similar construction in the proof of Theorem A. The main difference is that we need to glue strainer maps instead of almost splitting maps.

Take small ϵ>0\epsilon>0. Assume that (X,d,𝔪,p)(X,d,\mathfrak{m},p) is an RCD(ϵ,N)\mathrm{RCD}(-\epsilon,N) space such that (X,d)(X,d) is CBA(ϵ)\mathrm{CBA}(\epsilon), X=\partial X=\emptyset. Suppose that (K,q)(K,q) is a smooth Riemannian nn-manifold with injK10ϵ\mathrm{inj}_{K}\geq\frac{10}{\epsilon}. B10ϵ(p)B_{\frac{10}{\epsilon}}(p) is ϵ2\epsilon^{2}-C4C^{4}-close, by expp1\mathrm{exp}_{p}^{-1}, to its preimage in the tangent space TpKT_{p}K with the flat metric. Suppose that h:B2(p)B2(q)h:B_{2}(p)\to B_{2}(q) is an ϵ\epsilon-GHA.

By [26], we can use a strainer to construct a differentiable Φ(ϵ|n)\Phi(\epsilon|n)-GHA u:B1(p)B2(q)u:B_{1}(p)\to B_{2}(q) with 1Φ(ϵ|n)|u|1+Φ(ϵ|n)1-\Phi(\epsilon|n)\leq|\nabla u|\leq 1+\Phi(\epsilon|n). In particular, uu is diffeomorphic onto its image. Moreover, if we use another strainer to construct another differentiable Φ(ϵ|n)\Phi(\epsilon|n)-GHA, saying u:B1(p)B2(q)u^{\prime}:B_{1}(p)\to B_{2}(q). Then uu^{\prime} is Φ(ϵ|n)\Phi(\epsilon|n)-C1C^{1}-close to uu.

By the same construction in Theorem 3.5, we can glue local strainer maps and obtain the following theorem; see also the proof of the fibration theorem for mixed curvature spaces in [26].

Theorem 5.1.

Given NN and nNn\leq N, there exists ϵ(N,n)\epsilon(N,n) such that for any ϵϵ(N,n)\epsilon\leq\epsilon(N,n), the following holds. Assume that a smooth nn-manifold KK satisfies injK10ϵ\mathrm{inj}_{K}\geq\frac{10}{\epsilon} and for any pKp\in K, B10ϵ(p)B_{\frac{10}{\epsilon}}(p) is ϵ2\epsilon^{2}-C4C^{4}-close, by expp1\mathrm{exp}_{p}^{-1}, to its preimage in the tangent space TpKT_{p}K with the flat metric. Suppose that (X,d,𝔪)(X,d,\mathfrak{m}) is an RCD(ϵ,N)\mathrm{RCD}(-\epsilon,N) space such that (X,d)(X,d) is CBA(ϵ)\mathrm{CBA}(\epsilon), X=\partial X=\emptyset, dim(X)=n\mathrm{dim}(X)=n. Assume that there is a global map h:XKh:X\to K which is an ϵ\epsilon-GHA map on any 10ϵ\frac{10}{\epsilon}-ball in XX; a finite group GG acts isometrically on XX and KK separately and hh is ϵ\epsilon-almost GG-equivariant.

Then there is a GG-equivariant map fG:XKf_{G}:X\to K, which is also a Φ(ϵ|n,N)\Phi(\epsilon|n,N)-GHA and biLipschitz diffeomorphic on any 1ϵ\frac{1}{\epsilon}-ball, that is, for any x,yXx,y\in X with d(x,y)<1ϵd(x,y)<\frac{1}{\epsilon},

(1Φ(ϵ|n,N))d(x,y)d(fG(x),fG(y))(1+Φ(ϵ|n,N))d(x,y),(1-\Phi(\epsilon|n,N))d(x,y)\leq d(f_{G}(x),f_{G}(y))\leq(1+\Phi(\epsilon|n,N))d(x,y),

where Φ(ϵ|n,N)0\Phi(\epsilon|n,N)\to 0 as ϵ0\epsilon\to 0.

We now proceed to prove Theorem 1.7 using Theorem 5.1 and the construction in the proof of Theorem A.

Proof of Theorem 1.7.

We apply a contradiction argument. Assume that there exists ϵi0\epsilon_{i}\to 0 so that there exists (Xi,di,𝔪i)(X_{i},d_{i},\mathfrak{m}_{i}) which is RCD(ϵi,N)\mathrm{RCD}(-\epsilon_{i},N) and CBA(ϵi)\mathrm{CBA}(\epsilon_{i}); Xi=\partial X_{i}=\emptyset and diam(Xi)ϵi\mathrm{diam}(X_{i})\leq\epsilon_{i}. And we assume that none of XiX_{i} is biLipschitz diffeomorphic to an infranil-manifold of dimension N\leq N. We need to prove that XiX_{i} is biLipschitz diffeomorphic to an infranil-manifold of dimension N\leq N for some large ii.

Passing to a subsequence if necessary, we may assume that all XiX_{i} are manifolds of dimension nn where nNn\leq N. We recall the argument from [26] that the universal cover X~i\widetilde{X}_{i} of XiX_{i} pGH converges to n\mathbb{R}^{n}. Specifically, take piXip_{i}\in X_{i} and fix a large r>0r>0. We consider X^i=Br(0n)TpiXi\hat{X}_{i}=B_{r}(0^{n})\subset T_{p_{i}}X_{i}, the rr-ball centered at the origin in the tangent space TpiXiT_{p_{i}}X_{i} with the pull back metric. Since the curvature of XiX_{i} is bounded above by ϵi0\epsilon_{i}\to 0, the exponential map exppi:X^iXi\mathrm{exp}_{p_{i}}:\hat{X}_{i}\to X_{i} is non-degenerate for sufficiently large ii. Thus X^i\hat{X}_{i} a pseudo-cover of XiX_{i} with pseudo-actions. By Lemma 3.7 in [26] X^i\hat{X}_{i} pGH converges to a rr-ball in n\mathbb{R}^{n}.

Then by Lemma 2.5 in [26], the groupfication of pseudo-actions is exactly π1(Xi)\pi_{1}(X_{i}). Therefore the gluing space using the pseudo-cover and pseudo-actions (see section 2.3) is exactly the universal cover X~i\widetilde{X}_{i} which also pGH converges to n\mathbb{R}^{n}.

Let p~i\tilde{p}_{i} be a lift of pip_{i} in X~i\widetilde{X}_{i}. By the generalized Margulis lemma, for sufficiently large ii, Gi=π1(Xi,pi)G_{i}^{\prime}=\pi_{1}(X_{i},p_{i}) contains a nilpotent subgroup GiG_{i} with index C\leq C. Properly rescale the measure on X~i\widetilde{X}_{i} and let Xi=X~i/GiX_{i}^{\prime}=\widetilde{X}_{i}/G_{i}. By the splitting theorem, we have the following diagram.

(X~i,p~i,Gi,Gi,𝔪i)epmGH(n,p~,G,G,n)ππ(Xi,pi,Gi/Gi)GHptππ(Xi,pi)GHpt\begin{CD}(\widetilde{X}_{i},\tilde{p}_{i},G_{i},G_{i}^{\prime},\mathfrak{m}_{i})@>{epmGH}>{}>(\mathbb{R}^{n},\tilde{p},G,G^{\prime},\mathcal{H}^{n})\\ @V{}V{\pi}V@V{}V{\pi}V\\ (X_{i}^{\prime},p_{i}^{\prime},G_{i}^{\prime}/G_{i})@>{GH}>{}>\mathrm{pt}\\ @V{}V{\pi}V@V{}V{\pi}V\\ (X_{i},p_{i})@>{GH}>{}>\mathrm{pt}\end{CD}

By Lemma 4.2, GG is free, thus can be identified as n\mathbb{R}^{n}.

GiG_{i} has no small subgroup due to the measure convergence. Then by Lemma 4.4, for sufficiently large ii, GiG_{i} is isomorphic to a lattice in a nn-dim simply connected nilpotent Lie group 𝒩i\mathcal{N}_{i}. For any ϵ>0\epsilon>0 and sufficiently large ii, by Lemma 4.5, we can endow 𝒩i\mathcal{N}_{i} a left-invariant metric so that 𝒩i\mathcal{N}_{i} is ϵ\epsilon-C4C^{4}-close to n\mathbb{R}^{n}. Next we apply Lemma 4.6 and 4.7, we can find a map hi:X~i𝒩ih_{i}:\widetilde{X}_{i}\to\mathcal{N}_{i} so that , hih_{i} is an ϵi\epsilon_{i}-GHA on any 1ϵ\frac{1}{\epsilon}-ball in X~i\widetilde{X}_{i} and ϵi\epsilon_{i}-almost GiG_{i}-equivariant. By the rigidity, GiG_{i}^{\prime} can be identified as a discrete subset of 𝒩iAut(𝒩i)\mathcal{N}_{i}\rtimes\text{Aut}(\mathcal{N}_{i}). Then hih_{i} is also ϵi\epsilon_{i}-almost GiG_{i}^{\prime}-equivariant.

We can find a normal subgroup Gi′′G_{i}^{\prime\prime} of GiG_{i}^{\prime} with finite index so that Gi′′GiG_{i}^{\prime\prime}\subset G_{i} and Gi′′B1ϵ(e)=G_{i}^{\prime\prime}\cap B_{\frac{1}{\epsilon}}(e)=\emptyset where e𝒩ie\in\mathcal{N}_{i}. In particular, 𝒩i/Gi′′\mathcal{N}_{i}/G_{i}^{\prime\prime} is compact and ϵ\epsilon-C4C^{4}-close to n\mathbb{R}^{n} on any 1ϵ\frac{1}{\epsilon}-ball. Then we can find a map h¯i:X~i/Gi′′𝒩i/Gi′′\bar{h}_{i}:\widetilde{X}_{i}/G_{i}^{\prime\prime}\to\mathcal{N}_{i}/G_{i}^{\prime\prime}, which is ϵi\epsilon_{i}-GHA on any 1ϵ\frac{1}{\epsilon}-ball and ϵi\epsilon_{i}-almost Gi/Gi′′G_{i}^{\prime}/G_{i}^{\prime\prime}-equivariant. Then applying Theorem 5.1 to h¯i\bar{h}_{i}, we can obtain an Gi/Gi′′G_{i}^{\prime}/G_{i}^{\prime\prime}-equivariant GHA map fGi/Gi′′:X~i/Gi′′𝒩i/Gi′′f_{G_{i}^{\prime}/G_{i}^{\prime\prime}}:\widetilde{X}_{i}/G_{i}^{\prime\prime}\to\mathcal{N}_{i}/G_{i}^{\prime\prime}, which is biLipschitz diffeomorphic on any 1ϵ\frac{1}{\epsilon}-ball.

Since fGi/Gi′′f_{G_{i}^{\prime}/G_{i}^{\prime\prime}} is Gi/Gi′′G_{i}^{\prime}/G_{i}^{\prime\prime}-equivariant and the diameter of XiX_{i} converges to 0, fGi/Gi′′f_{G_{i}^{\prime}/G_{i}^{\prime\prime}} induces a biLipschitz diffeomorphic map on the quotient space

fi:Xi=X~i/Gi𝒩i/Gi.f_{i}:X_{i}=\widetilde{X}_{i}/G_{i}^{\prime}\to\mathcal{N}_{i}/G_{i}^{\prime}.

Therefore XiX_{i} is biLipschitz diffeomorphic to an infranil-manifold 𝒩i/Gi\mathcal{N}_{i}/G_{i}^{\prime}. ∎

6. Proof of Theorem B: construct fibers along the collapsing direction

We restate Theorem B.

Theorem B.

Given N,v>0N,v>0, assume that a sequence of compact RCD((N1),N)(-(N-1),N) spaces (Xi,di,N)(X_{i},d_{i},\mathcal{H}^{N}) with (1,v)(1,v)-bound covering geometry converges to a kk-manifold KK in the Gromov-Hausdorff sense. Then for all large ii, there is a fiber bundle map fi:XiKf_{i}:X_{i}\to K which is a GHA. Moreover, the fiber is homeomorphic to an infranil-manifold and the structure group is affine.

6.1. Basic constructions

To prove Theorem B, we first employ the construction from Theorem 3.3 to obtain an ϵi\epsilon_{i}-GHA fi:XiKf_{i}:X_{i}\to K which is locally almost kk-splitting. We want to prove that fif_{i} is actually a fibration map for sufficiently large ii. Assume that for some piXip_{i}\in X_{i} converging to pKp\in K, fif_{i} is not a fibration map on any small neighborhood of pip_{i}. We will show the existence of fiberation which leads to a contradiction.

By the generalized Margulis lemma, we can find ϵ0>0\epsilon_{0}>0 such that GiG_{i}^{\prime}, the image of π1(Bϵ0(pi))π1(B1(pi))\pi_{1}(B_{\epsilon_{0}}(p_{i}))\to\pi_{1}(B_{1}(p_{i})), contains a nilpotent subgroup GiG_{i} of finite index C\leq C. Let B~(pi,ϵ0,1)\widetilde{B}(p_{i},\epsilon_{0},1) be a connected component of the pre-image of Bϵ0(pi)B_{\epsilon_{0}}(p_{i}) in the universal cover of B1(pi)B_{1}(p_{i}). Passing to a subsequence if necessary,

(B~(pi,ϵ0,1),p~i)pGH(Y,p~)(\widetilde{B}(p_{i},\epsilon_{0},1),\tilde{p}_{i})\overset{pGH}{\longrightarrow}(Y,\tilde{p})

by the precomactness Theorem 2.9. Moreover, by [23], the tangent cone at x~Y\tilde{x}\in Y must be N\mathbb{R}^{N} since any tangent cone at KK is k\mathbb{R}^{k}.

We use the notation riBϵ0(pi)r_{i}B_{\epsilon_{0}}(p_{i}) for the set Bϵ0(pi)XiB_{\epsilon_{0}}(p_{i})\subset X_{i} with the rescaled metric ridir_{i}d_{i}. Take rir_{i}\to\infty slowly to blow up the metric and pass to a subsequence if necessary,

(6.1) (riB~(pi,ϵ0,1),p~i,Gi,Gi)eGH(N,0N,G,G)ππ(riB~(pi,ϵ0,1)/Gi,pi,Gi/Gi)GH(Y,p,G¯)ππ(riBϵ0(pi),pi)Fi(k,0k)\displaystyle\begin{CD}(r_{i}\widetilde{B}(p_{i},\epsilon_{0},1),\tilde{p}_{i},G_{i},G_{i}^{\prime})@>{eGH}>{}>(\mathbb{R}^{N},0^{N},G,G^{\prime})\\ @V{}V{\pi}V@V{}V{\pi}V\\ (r_{i}\widetilde{B}(p_{i},\epsilon_{0},1)/G_{i},p_{i}^{\prime},G_{i}^{\prime}/G_{i})@>{GH}>{}>(Y^{\prime},p^{\prime},\bar{G})\\ @V{}V{\pi}V@V{}V{\pi}V\\ (r_{i}B_{\epsilon_{0}}(p_{i}),p_{i})@>{F_{i}}>{}>(\mathbb{R}^{k},0^{k})\end{CD}

where Fi=expp1fiF_{i}=\exp_{p}^{-1}\circ f_{i} is almost kk-splitting by Theorem 3.3. Since GiG_{i} is a normal subgroup of GiG_{i}^{\prime} with index C\leq C, B~(pi,ϵ0,1)/Gi\widetilde{B}(p_{i},\epsilon_{0},1)/G_{i} is a finite cover of Bϵ0(pi)B_{\epsilon_{0}}(p_{i}).

By the covering lemma 2.12,

F~i:riB~(pi,ϵ0,1)𝜋riBϵi(pi)Fi(k,0k),\tilde{F}_{i}:r_{i}\widetilde{B}(p_{i},\epsilon_{0},1)\overset{\pi}{\longrightarrow}r_{i}B_{\epsilon_{i}}(p_{i})\overset{F_{i}}{\longrightarrow}(\mathbb{R}^{k},0^{k}),
F~i:riB~(pi,ϵ0,1)/Gi𝜋riBϵi(pi)Fi(k,0k)\tilde{F}_{i}^{\prime}:r_{i}\widetilde{B}(p_{i},\epsilon_{0},1)/G_{i}\overset{\pi}{\longrightarrow}r_{i}B_{\epsilon_{i}}(p_{i})\overset{F_{i}}{\longrightarrow}(\mathbb{R}^{k},0^{k})

are also almost kk-splitting. In particular, Y=k×Y′′Y^{\prime}=\mathbb{R}^{k}\times Y^{\prime\prime} and N=k×Nk\mathbb{R}^{N}=\mathbb{R}^{k}\times\mathbb{R}^{N-k}.

Thus G¯,G,G\bar{G},G,G^{\prime} actions are trivial on the first k\mathbb{R}^{k} component of 𝕐\mathbb{Y}^{\prime} and N\mathbb{R}^{N} respectively. In particular, since the order of |Gi/Gi|C|G_{i}^{\prime}/G_{i}|\leq C, Y′′Y^{\prime\prime} is a finite set. Y′′Y^{\prime\prime} is connected, thus Y′′Y^{\prime\prime} is a single point and Y=kY^{\prime}=\mathbb{R}^{k}. Then G¯\bar{G} is trivial; GG and GG^{\prime} act transitively on the Nk\mathbb{R}^{N-k} component of N=k×Nk\mathbb{R}^{N}=\mathbb{R}^{k}\times\mathbb{R}^{N-k}. Thus by the same proof of Lemma 4.2, we conclude that:

Lemma 6.1.

GG is free. In particular, GG can be identified as Nk\mathbb{R}^{N-k}.

By Lemma 4.3, GiG_{i} contains no small subgroup. Now we can simplify diagram 6.1 as:

(6.2) (riB~(pi,ϵ0,1),p~i,Gi,Gi)eGH(N,0N,G=Nk,G)ππ(riB~(pi,ϵ0,1)/Gi,pi,Gi/Gi)eGH(k,0k,id)ππ(riBϵ0(pi),pi)Fi(k,0k)\displaystyle\begin{CD}(r_{i}\widetilde{B}(p_{i},\epsilon_{0},1),\tilde{p}_{i},G_{i},G_{i}^{\prime})@>{eGH}>{}>(\mathbb{R}^{N},0^{N},G=\mathbb{R}^{N-k},G^{\prime})\\ @V{}V{\pi}V@V{}V{\pi}V\\ (r_{i}\widetilde{B}(p_{i},\epsilon_{0},1)/G_{i},p_{i}^{\prime},G_{i}^{\prime}/G_{i})@>{eGH}>{}>(\mathbb{R}^{k},0^{k},\mathrm{id})\\ @V{}V{\pi}V@V{}V{\pi}V\\ (r_{i}B_{\epsilon_{0}}(p_{i}),p_{i})@>{F_{i}}>{}>(\mathbb{R}^{k},0^{k})\end{CD}

The strategy for proving Theorem B follows a similar approach to the proof of Theorem A. Specifically, we use a nilpotent subgroup of the (local relative) fundamental group to construct a simply connected nilpotent Lie group. Then we show that the (local relative) covering space locally admits a fibration structure. However, in the case of Theorem B, there are two additional challenges compared to the proof of Theorem A.

The first challenge is that the generators of GiG_{i} may be large in the rescaled metric spaces riB~(pi,ϵ0,1)r_{i}\widetilde{B}(p_{i},\epsilon_{0},1). The second challenge is that we cannot directly apply Theorem 2.8 to local relative fundamental groups.

From now on we always consider rescaled metrics ridir_{i}d_{i}. We solve the first issue using the gap lemma and choosing a subgroup of GiG_{i}.

For any r>0r>0, define

Gi(p~i,r)={gGi|ridi(gp~i,p^i)r}.G_{i}(\tilde{p}_{i},r)=\{g\in G_{i}|r_{i}d_{i}(g\tilde{p}_{i},\hat{p}_{i})\leq r\}.

By the gap lemma 2.13 and Lemma 6.1, there exists ϵi0\epsilon_{i}\to 0 with ϵiri\epsilon_{i}r_{i}\to\infty and for all r[ϵi,1ϵi]r\in[\epsilon_{i},\frac{1}{\epsilon_{i}}], Gi(p^i,r)\langle G_{i}(\hat{p}_{i},r)\rangle is the same group, saying HiH_{i}. Then the equivariant limit of HiH_{i} contains a neighborhood of G=NkG=\mathbb{R}^{N-k} thus must equal GG itself.

We next prove that

Gi(p~i,r)={gGi|ridi(gp~i,p^i)r}G_{i}^{\prime}(\tilde{p}_{i},r)=\{g^{\prime}\in G_{i}^{\prime}|r_{i}d_{i}(g^{\prime}\tilde{p}_{i},\hat{p}_{i})\leq r\}

generates the same group in GiG_{i}^{\prime} for all r[2ϵi,1ϵiϵi]r\in[2\epsilon_{i},\frac{1}{\epsilon_{i}}-\epsilon_{i}] as well, saying HiH_{i}^{\prime}. For any gGi(p~,1ϵiϵi)g^{\prime}\in G_{i}^{\prime}(\tilde{p},\frac{1}{\epsilon_{i}}-\epsilon_{i}), since Gi/GiG_{i}/G_{i}^{\prime} converges to the trivial group, we can find gGi(p~i,1ϵi)g\in G_{i}(\tilde{p}_{i},\frac{1}{\epsilon_{i}}) so that ridi(gg1p~,p~)2ϵir_{i}d_{i}(g^{\prime}g^{-1}\tilde{p},\tilde{p})\leq 2\epsilon_{i}. Thus gg1Gi(p~,2ϵi)g^{\prime}g^{-1}\in G_{i}^{\prime}(\tilde{p},2\epsilon_{i}). On the other hand gHiGi(p~i,2ϵi)g\in H_{i}\subset\langle G_{i}^{\prime}(\tilde{p}_{i},2\epsilon_{i})\rangle thus g=gg1gGi(p~i,2ϵi)g^{\prime}=g^{\prime}g^{-1}g\in\langle G_{i}^{\prime}(\tilde{p}_{i},2\epsilon_{i})\rangle.

Lemma 6.2.

For sufficiently large ii, HiH_{i} is a normal subgroup of HiH_{i}^{\prime} with index C\leq C, where CC is the constant in the generalized Margulis lemma.

Proof.

Hi=Gi(p~i,2ϵi)H_{i}^{\prime}=\langle G_{i}^{\prime}(\tilde{p}_{i},2\epsilon_{i})\rangle and Hi=Gi(p~i,ϵi)H_{i}=\langle G_{i}(\tilde{p}_{i},\epsilon_{i})\rangle. To show that HiH_{i} is normal in HiH_{i}^{\prime}, we only need to show that h1ghHih^{-1}gh\in H_{i} for any hGi(p~i,2ϵi)h\in G_{i}^{\prime}(\tilde{p}_{i},2\epsilon_{i}) and gGi(p~i,ϵi)g\in G_{i}(\tilde{p}_{i},\epsilon_{i}). Since GiG_{i} is normal in GiG_{i}^{\prime}, h1ghGih^{-1}gh\in G_{i}. Then

ridi(p~i,h1ghp~i)\displaystyle r_{i}d_{i}(\tilde{p}_{i},h^{-1}gh\tilde{p}_{i}) =ridi(hp~i,ghp~i)\displaystyle=r_{i}d_{i}(h\tilde{p}_{i},gh\tilde{p}_{i})
ridi(hp~i,p~i)+ridi(p~i,gp~i)+ridi(gp~i,ghp~i)\displaystyle\leq r_{i}d_{i}(h\tilde{p}_{i},\tilde{p}_{i})+r_{i}d_{i}(\tilde{p}_{i},g\tilde{p}_{i})+r_{i}d_{i}(g\tilde{p}_{i},gh\tilde{p}_{i})
5ϵi.\displaystyle\leq 5\epsilon_{i}.

Since Gi(p~i,5ϵi)HiG_{i}(\tilde{p}_{i},5\epsilon_{i})\subset H_{i}, h1ghHih^{-1}gh\in H_{i}. Thus HiH_{i} is a normal subgroup of HiH_{i}^{\prime}.

Next we prove that |Hi/Hi|C|H_{i}^{\prime}/H_{i}|\leq C. Assume |Hi/Hi|>C|Gi/Gi||H_{i}^{\prime}/H_{i}|>C\geq|G_{i}^{\prime}/G_{i}|. Let TT denote the image of Hi(p~i,2ϵi)H_{i}^{\prime}(\tilde{p}_{i},2\epsilon_{i}) in Hi/HiH_{i}^{\prime}/H_{i}. Define T2={g1g2|g1,g2T}T^{2}=\{g_{1}g_{2}|g_{1},g_{2}\in T\} and similarly TlT^{l} for l>0l>0. If Tl+1=TlT^{l+1}=T^{l} for some l>0l>0, then Tl=Hi/HiT^{l}=H_{i}^{\prime}/H_{i}. Therefore either |Tl+1||Tl|+1|T^{l+1}|\geq|T^{l}|+1 or |Tl|>C|T^{l}|>C. In either case, we always have the order |TC+1|C+1|T^{C+1}|\geq C+1. However, since ϵi0\epsilon_{i}\to 0, (Hi(p~i,2ϵi))C+1Hi(p~i,1)(H_{i}^{\prime}(\tilde{p}_{i},2\epsilon_{i}))^{C+1}\subset H_{i}^{\prime}(\tilde{p}_{i},1) for sufficiently large ii. Then we can find g1,g2Hi(p~i,1)g_{1},g_{2}\in H_{i}^{\prime}(\tilde{p}_{i},1) so that they have different images in Hi/HiH_{i}^{\prime}/H_{i} but the same image in Gi/GiG_{i}^{\prime}/G_{i}. Then g1g21Gi(p~i,3)Hig_{1}g_{2}^{-1}\in G_{i}(\tilde{p}_{i},3)\subset H_{i}, a contradiction. ∎

Since any giGig_{i}\in G_{i} but giHig_{i}\notin H_{i} satisfies ridi(p~,gp~)>1ϵir_{i}d_{i}(\tilde{p},g\tilde{p})>\frac{1}{\epsilon_{i}}, it diverges in the pointed Gromov-Hausdorff sense. Thus a large ball in riB~(pi,ϵ0,1)/Gir_{i}\widetilde{B}(p_{i},\epsilon_{0},1)/G_{i} is isometric to a large ball in riB~(pi,ϵ0,1)/Hir_{i}\widetilde{B}(p_{i},\epsilon_{0},1)/H_{i}. To simplify the notation, we can identify:

riB~(pi,ϵ0,1)/Gi=riB~(pi,ϵ0,1)/Hi.r_{i}\widetilde{B}(p_{i},\epsilon_{0},1)/G_{i}=r_{i}\widetilde{B}(p_{i},\epsilon_{0},1)/H_{i}.

Similar we can identify:

riBϵ0(pi)=riB~(pi,ϵ0,1)/Gi=riB~(pi,ϵ0,1)/Hi.r_{i}B_{\epsilon_{0}}(p_{i})=r_{i}\widetilde{B}(p_{i},\epsilon_{0},1)/G_{i}^{\prime}=r_{i}\widetilde{B}(p_{i},\epsilon_{0},1)/H_{i}^{\prime}.

From 6.2, we have the following commutative diagram:

(6.3) (riB~(pi,ϵ0,1),p~i,Hi,Hi)eGH(N,0N,H=Nk,H)ππ(riB~(pi,ϵ0,1)/Hi,pi,Hi/Hi)eGH(k,0k,id)ππ(riBϵ0(pi),pi)Fi(k,0k)\displaystyle\begin{CD}(r_{i}\widetilde{B}(p_{i},\epsilon_{0},1),\tilde{p}_{i},H_{i},H_{i}^{\prime})@>{eGH}>{}>(\mathbb{R}^{N},0^{N},H=\mathbb{R}^{N-k},H^{\prime})\\ @V{}V{\pi}V@V{}V{\pi}V\\ (r_{i}\widetilde{B}(p_{i},\epsilon_{0},1)/H_{i},p_{i}^{\prime},H_{i}^{\prime}/H_{i})@>{eGH}>{}>(\mathbb{R}^{k},0^{k},\mathrm{id})\\ @V{}V{\pi}V@V{}V{\pi}V\\ (r_{i}B_{\epsilon_{0}}(p_{i}),p_{i})@>{F_{i}}>{}>(\mathbb{R}^{k},0^{k})\end{CD}

6.2. The existence of local almost product structure

In this subsection, we always consider the rescaled metric ridir_{i}d_{i} on (Xi,pi)(X_{i},p_{i}). Let Brri(pi)B^{r_{i}}_{r}(p_{i}) denote the rr-ball of pip_{i} with respect to the rescaled metric ridir_{i}d_{i}. We will always use the rescaled metric ridir_{i}d_{i} on Brri(pi)B^{r_{i}}_{r}(p_{i}).

The goal of this subsection is to prove the following result.

Lemma 6.3 (Existence of local product structure).

In the context of 6.3, there exists r>0r>0 so that for sufficiently large ii, Brri(pi)B_{r}^{r_{i}}(p_{i}^{\prime}) is, removing some points near the boundary if necessary, biHölder homeomorphic to Br(0k)×𝒩i/H^iB_{r}(0^{k})\times\mathcal{N}_{i}/\hat{H}_{i}, where Br(0k)kB_{r}(0^{k})\subset\mathbb{R}^{k} and 𝒩i\mathcal{N}_{i} is a simply connected nilpotent Lie group with lattice H^i\hat{H}_{i}.

Remark 6.4.

Lemma 6.3 claims the existence of a nil-manifold fibration structure on a neighborhood of pip_{i}^{\prime} (not pip_{i}). The obstruction to constructing an infranil-manifold fibration structure near pip_{i} is that HiH_{i} is not necessarily isomorphic to a lattice. This issue will be addressed in Lemma 6.7.

The proof of Lemma 6.3 follows by applying strategy used in the proof of Theorem A to the collapsing direction. In Theorem A, we first use Theorem 2.26 to find the nilprogression structure of a neighborhood of the identity in the fundamental group; then we apply Theorem 2.8 to show that the fundamental group is isomorphic to a lattice in a nilpotent Lie group. Unfortunately, we cannot directly apply Theorem 2.8 to HiH_{i} and corresponding relative covers, thus we need to consider the groupfication of a neighborhood of the identity in HiH_{i}.

By Theorem 2.26, there exists r>0r>0 such that for all large ii, Hi(p~i,10)H_{i}(\tilde{p}_{i},10) contains a nilprogression PiP_{i} which contains Hi(p~i,4r)H_{i}(\tilde{p}_{i},4r).

Lemma 6.5.

For any fixed 0<ϵ<4r0<\epsilon<4r and sufficiently large ii, the groupfication of Hi(p~i,4r)H_{i}(\tilde{p}_{i},4r) is naturally isomorphic to the groupfication of Hi(p~i,ϵ)H_{i}(\tilde{p}_{i},\epsilon).

Proof.

We have Hi(p~i,ϵ)Hi(p~i,4r)PiH_{i}(\tilde{p}_{i},\epsilon)\subset H_{i}(\tilde{p}_{i},4r)\subset P_{i}. HiH_{i} can be generated by Hi(p~i,ϵi)H_{i}(\tilde{p}_{i},\epsilon_{i}) where ϵi0\epsilon_{i}\to 0. Using the escape norm, the generators of PiP_{i} are ujHi(p~i,ϵi)u_{j}\in H_{i}(\tilde{p}_{i},\epsilon_{i}) and the relations in PiP_{i} are contained in Hi(p~i,ϵ)H_{i}(\tilde{p}_{i},\epsilon). Thus the groupfication of Hi(p~i,4r)H_{i}(\tilde{p}_{i},4r) or Hi(p~i,ϵ)H_{i}(\tilde{p}_{i},\epsilon) is naturally isomorphic to the groupfication of PiP_{i} which is a lattice in a simply connected nilpotent Lie group. ∎

Let H^i\hat{H}_{i} be the groupfication of Hi(p~i,4r)H_{i}(\tilde{p}_{i},4r). For sufficiently large ii, we know that Hi(p~i,r)=Hi\langle H_{i}(\tilde{p}_{i},r)\rangle=H_{i}, thus there is a natural surjective homomorphism si:H^iHis_{i}:\hat{H}_{i}\to H_{i}. We shall use the argument in the proof of Theorem A for H^i\hat{H}_{i}.

We construct a gluing space H^i×Hi(p~i,4r)B¯2rri(p~i)\hat{H}_{i}\times_{H_{i}(\tilde{p}_{i},4r)}\bar{B}^{r_{i}}_{2r}(\tilde{p}_{i}) which is a covering space of B¯2rri(pi)\bar{B}^{r_{i}}_{2r}(p_{i}). Let YiY_{i} be the pre-image of B¯rri(pi)\bar{B}^{r_{i}}_{r}(p_{i}) in H^i×Hi(p~i,4r)B¯2rri(p~i)\hat{H}_{i}\times_{H_{i}(\tilde{p}_{i},4r)}\bar{B}^{r_{i}}_{2r}(\tilde{p}_{i}). Then YiY_{i} is connected since the generators of H^i\hat{H}_{i} is contained in the image of natural injective pseudo-group homomorphism

Hi(p~i,ϵi)H^iH_{i}(\tilde{p}_{i},\epsilon_{i})\hookrightarrow\hat{H}_{i}

and ϵi<<r\epsilon_{i}<<r when ii is large enough.

Thus Hi=H^i/Ker(si)H_{i}=\hat{H}_{i}/\mathrm{Ker}(s_{i}) and define Zi=Yi/Ker(si)Z_{i}=Y_{i}/\mathrm{Ker}(s_{i}). ZiZ_{i} is the union of all HiH_{i}-orbits of B¯rri(pi)\bar{B}^{r_{i}}_{r}(p_{i}^{\prime}), and we have

Zi=π1(B¯rri(pi))riB~(pi,ϵ0,1).Z_{i}=\pi^{-1}(\bar{B}^{r_{i}}_{r}(p_{i}))\subset r_{i}\widetilde{B}(p_{i},\epsilon_{0},1).

Using 6.3 (and forget HiH_{i}^{\prime} for now) and the pre-compactness theorem, we obtain the following:

(6.4) (Yi,p^i,Ker(si),H^i)eGH(Y,p^,S,H^)ππ(Zi,p~i,Hi=H^i/Ker(si))eGH(Y/S=B¯r(0k)×Nk,0N,H=Nk)ππ(B¯rri(pi),pi)Fi(B¯r(0k),0k)\displaystyle\begin{CD}(Y_{i},\hat{p}_{i},\mathrm{Ker}(s_{i}),\hat{H}_{i})@>{eGH}>{}>(Y,\hat{p},S,\hat{H})\\ @V{}V{\pi}V@V{}V{\pi}V\\ (Z_{i},\tilde{p}_{i},H_{i}=\hat{H}_{i}/\mathrm{Ker}(s_{i}))@>{eGH}>{}>(Y/S=\bar{B}_{r}(0^{k})\times\mathbb{R}^{N-k},0^{N},H=\mathbb{R}^{N-k})\\ @V{}V{\pi}V@V{}V{\pi}V\\ (\bar{B}^{r_{i}}_{r}(p_{i}^{\prime}),p_{i}^{\prime})@>{F_{i}}>{}>(\bar{B}_{r}(0^{k}),0^{k})\end{CD}

where SS is the limit group of Ker(si)\mathrm{Ker}(s_{i}), and H=H^/SH=\hat{H}/S.

Lemma 6.6.

SS is a trivial group. Thus Y=B¯rri(0k)×NkY=\bar{B}^{r_{i}}_{r}(0^{k})\times\mathbb{R}^{N-k} and H^=Nk\hat{H}=\mathbb{R}^{N-k} are free translation actions.

Proof.

By [32, 44], for any gKer(si)g\in\mathrm{Ker}(s_{i}), ridi(gp^i,p^i)rr_{i}d_{i}(g\hat{p}_{i},\hat{p}_{i})\geq r. Thus the limit group SS is a discrete group.

We now claim that SS admits no isotropy subgroup. Otherwise assume that idhS\mathrm{id}\neq h\in S and hx=xhx=x for some xYx\in Y. We may assume that xx is not a boundary point, otherwise we take 3r2\frac{3r}{2}-ball instead of rr-ball when we define YiY_{i}, then xx is not on the boundary. Let x¯\bar{x} be the image of xx in Y/SY/S. The tangent cone of x¯\bar{x} is N\mathbb{R}^{N}, thus tangent cone at xx is also N\mathbb{R}^{N}. However, g\langle g\rangle has non-trivial limit group actions on the tangent cone of xx due to volume convergence, thus x¯\bar{x} and xx cannot have the same tangent cone, a contradiction.

Thus SS are free and discrete isometric group actions. Since B¯r(0k)×Nk\bar{B}_{r}(0^{k})\times\mathbb{R}^{N-k} is simply connected, SS is trivial. ∎

Now we have H^=H=Nk\hat{H}=H=\mathbb{R}^{N-k} and the following diagram:

(6.5) (Yi,p^i,Ker(si),H^i)eGH(Y=B¯r(0k)×Nk,0N,id,H^=Nk)ππ(Zi,p~i,Hi=H^i/Ker(si))eGH(B¯r(0k)×Nk,0N,H=Nk)ππ(B¯rri(pi),pi)Fi(B¯r(0k),0k)\displaystyle\begin{CD}(Y_{i},\hat{p}_{i},\mathrm{Ker}(s_{i}),\hat{H}_{i})@>{eGH}>{}>(Y=\bar{B}_{r}(0^{k})\times\mathbb{R}^{N-k},0^{N},\mathrm{id},\hat{H}=\mathbb{R}^{N-k})\\ @V{}V{\pi}V@V{}V{\pi}V\\ (Z_{i},\tilde{p}_{i},H_{i}=\hat{H}_{i}/\mathrm{Ker}(s_{i}))@>{eGH}>{}>(\bar{B}_{r}(0^{k})\times\mathbb{R}^{N-k},0^{N},H=\mathbb{R}^{N-k})\\ @V{}V{\pi}V@V{}V{\pi}V\\ (\bar{B}^{r_{i}}_{r}(p_{i}^{\prime}),p_{i}^{\prime})@>{F_{i}}>{}>(\bar{B}_{r}(0^{k}),0^{k})\end{CD}
Proof of Lemma 6.3.

By Theorem 2.26 and the diagram 6.5, there exists ϵ>0\epsilon>0, for sufficiently large ii, we can find a nilporgression PiH^i(p^i,1)P_{i}^{\prime}\subset\hat{H}_{i}(\hat{p}_{i},1) that contains H^i(p^i,ϵ)={gH^i|ridi(gp^,p^)ϵ}\hat{H}_{i}(\hat{p}_{i},\epsilon)=\{g\in\hat{H}_{i}|r_{i}d_{i}(g\hat{p},\hat{p})\leq\epsilon\}. We may assume ϵ<4r\epsilon<4r, then the map

si:Hi(p~i,ϵ)H^i(p^i,ϵ)s_{i}:H_{i}(\tilde{p}_{i},\epsilon)\to\hat{H}_{i}(\hat{p}_{i},\epsilon)

is injective [32, 44], thus a pseudo-group isomorphism. Since the groupfication of Hi(p~i,ϵ)H_{i}(\tilde{p}_{i},\epsilon) is isomorphic to H^i\hat{H}_{i} by Lemma 6.5, thus H^i\hat{H}_{i} is isomorphic to the groupfication of H^i(p^i,ϵ)\hat{H}_{i}(\hat{p}_{i},\epsilon), which is the groupfication of PiP_{i}^{\prime}.

We can construct a (Nk)(N-k)-dim nilpotent group 𝒩i\mathcal{N}_{i} using the nilprogession PiP_{i}^{\prime}, and endow NiN_{i} a left-invariant metric as in Lemma 4.5. Then H^i\hat{H}_{i} is isomorphic to a lattice in 𝒩i\mathcal{N}_{i}.

Take (0k,e)B¯r(0k)×𝒩i(0^{k},e)\in\bar{B}_{r}(0^{k})\times\mathcal{N}_{i}. By Lemma 4.6, we can construct a map

hi:Brri(p^i)Br(0k,e)h_{i}^{\prime}:B^{r_{i}}_{r}(\hat{p}_{i})\to B_{r}(0^{k},e)

which is an ϵi\epsilon_{i}-GHA and ϵi\epsilon_{i}-almost H^i(p^i,r)\hat{H}_{i}(\hat{p}_{i},r)-equivariant with some ϵi0\epsilon_{i}\to 0.

Then by Lemma 4.7, we can construct a global map, possibly dropping points near the boundary if necessary,

hi:(Yi,p^i)B¯r(0k)×𝒩ih_{i}:(Y_{i},\hat{p}_{i})\to\bar{B}_{r}(0^{k})\times\mathcal{N}_{i}

which is an ϵi\epsilon_{i}-GHA on any rr-ball and hh is ϵi\epsilon_{i}-almost H^i\hat{H}_{i}-equivariant.

Now we can find a normal subgroup H^i\hat{H}_{i}^{\prime} in H^i\hat{H}_{i} of finite index, with H^iBr(e)=\hat{H}_{i}^{\prime}\cap B_{r}(e)=\emptyset. Then by Theorem 3.5, after sufficiently blowing up the metric and dropping points near the boundary if necessary, there exists a H^i/H^i\hat{H}_{i}/\hat{H}_{i}^{\prime}-equivariant ϵi\epsilon_{i}-GHA

fH^i/H^i:Yi/H^iB¯r(0k)×𝒩i/H^i,f_{\hat{H}_{i}/\hat{H}_{i}^{\prime}}:Y_{i}/\hat{H}_{i}^{\prime}\longrightarrow\bar{B}_{r}(0^{k})\times\mathcal{N}_{i}/\hat{H}_{i}^{\prime},

which is locally almost NN-splitting. In particular, fH^i/H^if_{\hat{H}_{i}/\hat{H}_{i}^{\prime}} is biHölder. Thus Yi/Hi=B¯rri(pi)Y_{i}/H_{i}=\bar{B}^{r_{i}}_{r}(p_{i}^{\prime}) is biHölder homeomorphic to B¯r(0k)×𝒩i/H^i\bar{B}_{r}(0^{k})\times\mathcal{N}_{i}/\hat{H}_{i}. ∎

6.3. Local relative fundamental groups

Lemma 6.3 is not a proof of Theorem B since HiH_{i} may not be isomorphic to H^i\hat{H}_{i}; equivalently, Ker(si)\mathrm{Ker}(s_{i}) in 6.5 may not be trivial or the nilpotency rank of HiH_{i} may be strictly less than NkN-k. Then the structure of HiH_{i}^{\prime} is unknown for now. We shall solve this issue by considering another contradiction sequence.

We aim to prove the following in this subsection.

Lemma 6.7.

Take another contradiction sequence if necessary, we may assume that H^i=Hi\hat{H}_{i}=H_{i} and HiH_{i} is a normal subgroup of HiH_{i}^{\prime} with index C(N)\leq C(N).

Now we have the following from 6.5,

(6.6) (Yi,p^i,Ker(si),H^i)eGH(Y=B¯r(0k)×Nk,0N,id,H^=Nk)ππ(Zi,p~i,Hi=H^i/Ker(si))GH(B¯r(0k)×Nk,0N,H=Nk)ππ(B¯rri(pi),pi)eGH(B¯r(0k),0k),\displaystyle\begin{CD}(Y_{i},\hat{p}_{i},\mathrm{Ker}(s_{i}),\hat{H}_{i})@>{eGH}>{}>(Y=\bar{B}_{r}(0^{k})\times\mathbb{R}^{N-k},0^{N},\mathrm{id},\hat{H}=\mathbb{R}^{N-k})\\ @V{}V{\pi}V@V{}V{\pi}V\\ (Z_{i},\tilde{p}_{i},H_{i}=\hat{H}_{i}/\mathrm{Ker}(s_{i}))@>{GH}>{}>(\bar{B}_{r}(0^{k})\times\mathbb{R}^{N-k},0^{N},H=\mathbb{R}^{N-k})\\ @V{}V{\pi}V@V{}V{\pi}V\\ (\bar{B}^{r_{i}}_{r}(p_{i}^{\prime}),p_{i}^{\prime})@>{eGH}>{}>(\bar{B}_{r}(0^{k}),0^{k}),\end{CD}

and Brri(pi)B^{r_{i}}_{r}(p_{i}^{\prime}) is biHölder homeomorphic to Br(0k)×𝒩i/H^iB_{r}(0^{k})\times\mathcal{N}_{i}/\hat{H}_{i}.

Let CC be constant in the generalized Margulis lemma. Consider the quotient space riB~(pi,ϵ0,1)/Gir_{i}\widetilde{B}(p_{i},\epsilon_{0},1)/G_{i} which is a finite cover of riBϵ0(pi)r_{i}B_{\epsilon_{0}}(p_{i}) of index C\leq C. For any gGig\in G_{i} with gHig\notin H_{i}, ridi(gp~i,p~i)1ϵir_{i}d_{i}(g\tilde{p}_{i},\tilde{p}_{i})\geq\frac{1}{\epsilon_{i}}. Thus riB~(pi,ϵ0,1)/Gir_{i}\widetilde{B}(p_{i},\epsilon_{0},1)/G_{i} is isometric to riB~(pi,ϵ0,1)/Hir_{i}\widetilde{B}(p_{i},\epsilon_{0},1)/H_{i} on a large ball. Therefore, for any fixed s>0s>0 and ii large enough, a connected component of the pre-image of Bsri(pi)B_{s}^{r_{i}}(p_{i}) in riB~(pi,ϵ0,1)/Hir_{i}\widetilde{B}(p_{i},\epsilon_{0},1)/H_{i} is a cover of Bsri(pi)B_{s}^{r_{i}}(p_{i}) with index C\leq C; in particular, this cover is contained in a CsCs-ball.

Let H~i\widetilde{H}_{i}^{\prime} be the image of natural map π1(Br20C2ri(pi),pi)π1(Br10Cri(pi),pi)\pi_{1}(B^{r_{i}}_{\frac{r}{20C^{2}}}(p_{i}),p_{i})\to\pi_{1}(B^{r_{i}}_{\frac{r}{10C}}(p_{i}),p_{i}). Let Z~i\widetilde{Z}_{i} be a connected component of the pre-image of Br20C2ri(pi)B^{r_{i}}_{\frac{r}{20C^{2}}}(p_{i}) in riB~(pi,ϵ0,1)/Hir_{i}\widetilde{B}(p_{i},\epsilon_{0},1)/H_{i}. Then Z~i\widetilde{Z}_{i} is a finite cover of Br20C2ri(pi)B^{r_{i}}_{\frac{r}{20C^{2}}}(p_{i}) with index at most CC. In particular, Z~iBr20Cri(pi)Zi\widetilde{Z}_{i}\subset B^{r_{i}}_{\frac{r}{20C}}(p_{i}^{\prime})\subset Z_{i}.

Lemma 6.8.

There exists a natural injective homomorphism i:H^iH~ii:\hat{H}_{i}\to\widetilde{H}_{i}^{\prime}.

Proof.

By Lemma 6.3, Brri(pi)B^{r_{i}}_{r}(p_{i}^{\prime}) is biHölder homeomorphic to Brri(0k)×𝒩i/H^iB^{r_{i}}_{r}(0^{k})\times\mathcal{N}_{i}/\hat{H}_{i} and the diameter of 𝒩i/H^i<ϵi0\mathcal{N}_{i}/\hat{H}_{i}<\epsilon_{i}\to 0. In particular, the image of

π1(B2ϵiri(pi),pi)π1(Br10ri(pi),pi)\pi_{1}(B^{r_{i}}_{2\epsilon_{i}}(p_{i}^{\prime}),p_{i}^{\prime})\to\pi_{1}(B^{r_{i}}_{\frac{r}{10}}(p_{i}^{\prime}),p_{i}^{\prime})

is isomorphic to H^i\hat{H}_{i}.

Now we define the homomorphism i:H^iH~ii:\hat{H}_{i}\to\widetilde{H}_{i}^{\prime} as follows: for any gH^ig\in\hat{H}_{i}, we can take a loop γgB2ϵiri(pi)\gamma_{g}\subset B^{r_{i}}_{2\epsilon_{i}}(p_{i}^{\prime}) at pip_{i}^{\prime} that represents gg in the image of

π1(B2ϵiri(pi),pi)π1(Br10ri(pi),pi).\pi_{1}(B^{r_{i}}_{2\epsilon_{i}}(p_{i}^{\prime}),p_{i}^{\prime})\to\pi_{1}(B^{r_{i}}_{\frac{r}{10}}(p_{i}^{\prime}),p_{i}^{\prime}).

Let π:B2ϵiri(pi)B2ϵiri(pi)\pi:B^{r_{i}}_{2\epsilon_{i}}(p_{i}^{\prime})\to B^{r_{i}}_{2\epsilon_{i}}(p_{i}) be the projection map and define i(g)i(g) to be the element in H~i\widetilde{H}_{i}^{\prime} represented by the loop π(γg)\pi(\gamma_{g}).

The map ii is well-defined because that if two loops γ1,γ2B2ϵiri(pi)\gamma_{1},\gamma_{2}\subset B^{r_{i}}_{2\epsilon_{i}}(p_{i}^{\prime}) are homotpic to each other in Br10ri(pi)B^{r_{i}}_{\frac{r}{10}}(p_{i}^{\prime}), then they must be homotopic to each other in Br10Cri(pi)B^{r_{i}}_{\frac{r}{10C}}(p_{i}^{\prime}) by the local product structure in Lemma 6.3. Thus π(γ1)\pi(\gamma_{1}) is homotopic to π(γ2)\pi(\gamma_{2}) in Br10Cri(pi)B^{r_{i}}_{\frac{r}{10C}}(p_{i}). Then they represent the same element in H~i\widetilde{H}_{i}. Thus ii is well-defined. ii is a homomorphism by the same argument.

Then we prove that ii is injective. Assume that there exists an element gH^ig\in\hat{H}_{i} represented by a loop γgB2ϵi(pi)\gamma_{g}\subset B_{2\epsilon_{i}}(p_{i}^{\prime}) while π(γg)\pi(\gamma_{g}) is contractible in Br10Cri(pi)B^{r_{i}}_{\frac{r}{10C}}(p_{i}). A connected component of the pre-image Br10Cri(pi)B^{r_{i}}_{\frac{r}{10C}}(p_{i}) is contained in Br10ri(pi)B^{r_{i}}_{\frac{r}{10}}(p_{i}^{\prime}). By the homotopy lifting property, γg\gamma_{g} is contractible in Br10ri(pi)B^{r_{i}}_{\frac{r}{10}}(p_{i}^{\prime}), thus gg is the identity element in the image of π1(B2ϵiri(pi),pi)π1(Br10ri(pi),pi)\pi_{1}(B^{r_{i}}_{2\epsilon_{i}}(p_{i}^{\prime}),p_{i}^{\prime})\to\pi_{1}(B^{r_{i}}_{\frac{r}{10}}(p_{i}^{\prime}),p_{i}^{\prime}). Then ii is injective. ∎

Lemma 6.9.

The index |H~i:i(H^i)|C|\widetilde{H}_{i}^{\prime}:i(\hat{H}_{i})|\leq C where CC is the constant in the generalized Margulis lemma.

Proof.

Recall that Z~i\widetilde{Z}_{i} is a finite cover of Br20C2ri(pi)B^{r_{i}}_{\frac{r}{20C^{2}}}(p_{i}) with index at most CC. Assume that |H~i:i(H^i)|>C|\widetilde{H}_{i}^{\prime}:i(\hat{H}_{i})|>C. Then we can find two loops γ1,γ2\gamma_{1},\gamma_{2} in Br20C2ri(pi)B^{r_{i}}_{\frac{r}{20C^{2}}}(p_{i}) at pip_{i}, so that γ1γ21\gamma_{1}\gamma_{2}^{-1} does not represent an element in i(H^i)i(\hat{H}_{i}) while we lift γ1\gamma_{1} and γ2\gamma_{2} in Z~iBr20C(pi)\widetilde{Z}_{i}\subset B_{\frac{r}{20C}}(p_{i}^{\prime}) at pip_{i}^{\prime}, saying the γ1\gamma_{1}^{\prime} and γ2\gamma_{2}^{\prime}, then γ1\gamma_{1}^{\prime} and γ2\gamma_{2}^{\prime} have the same endpoint in Z~i\widetilde{Z}_{i}.

Since Brri(pi)B^{r_{i}}_{r}(p_{i}^{\prime}) is biHölder homeomorphic to Br(0k)×𝒩i/H^iB_{r}(0^{k})\times\mathcal{N}_{i}/\hat{H}_{i}, γ1(γ2)1\gamma_{1}^{\prime}(\gamma_{2}^{\prime})^{-1} is homotopic to a loop γg\gamma_{g} corresponding to some gH^ig\in\hat{H}_{i} in Br(0k)×𝒩i/H^iB_{r}(0^{k})\times\mathcal{N}_{i}/\hat{H}_{i}. Moreover, the homotopy image is contained in Br10Cri(pi)B^{r_{i}}_{\frac{r}{10C}}(p_{i}^{\prime}) and γgB2ϵi(pi)\gamma_{g}\subset B_{2\epsilon_{i}}(p_{i}^{\prime}) as the diameter of 𝒩i/H^i\mathcal{N}_{i}/\hat{H}_{i} converges to 0. We can project this homotopy map to Br10Cri(pi)B^{r_{i}}_{\frac{r}{10C}}(p_{i}), then γ1γ21\gamma_{1}\gamma_{2}^{-1} represents i(g)i(H^i)i(g)\in i(\hat{H}_{i}), a contradiction. ∎

Since H~i\widetilde{H}_{i}^{\prime} contains a nilpotent subgroup i(H^i)i(\hat{H}_{i}) of index C\leq C, then by Lemma 4.8, H~i\widetilde{H}_{i}^{\prime} contains a normal nilpotent subgroup H~i\widetilde{H}_{i} of index C(N)=C!\leq C(N)=C!. Since H^i\hat{H}_{i} is torsion free nilpotent group with rank NkN-k and H~i\widetilde{H}_{i} is a normal subgroup of i(H^i)i(\hat{H}_{i}) with finite index, thus rank(H~i)=Nk(\widetilde{H}_{i})=N-k.

Proof of Lemma 6.7.

Now we return to the setup of 6.1 in the beginning of this section. Recall that H~i\widetilde{H}_{i}^{\prime} is the image of π1(Br20C2ri(pi),pi)π1(Br10Cri(pi),pi)\pi_{1}(B^{r_{i}}_{\frac{r}{20C^{2}}}(p_{i}),p_{i})\to\pi_{1}(B^{r_{i}}_{\frac{r}{10C}}(p_{i}),p_{i}). We may take rir_{i}^{\prime}\to\infty slowly, then we have another contradiction sequence,

(6.7) (ririB~(pi,ϵ0,1),p~i,H~i,H~i)eGH(N,0N,G,G)ππ(ririB~(pi,ϵ0,1)/H~i,pi,H~i/H~i)GH(Y,p,G¯)ππ(ririBϵ0(pi),pi)Fi(k,0k)\displaystyle\begin{CD}(r_{i}^{\prime}r_{i}\widetilde{B}(p_{i},\epsilon_{0},1),\tilde{p}_{i},\widetilde{H}_{i},\widetilde{H}_{i}^{\prime})@>{eGH}>{}>(\mathbb{R}^{N},0^{N},G,G^{\prime})\\ @V{}V{\pi}V@V{}V{\pi}V\\ (r_{i}^{\prime}r_{i}\widetilde{B}(p_{i},\epsilon_{0},1)/\widetilde{H}_{i},p_{i}^{\prime},\widetilde{H}_{i}^{\prime}/\widetilde{H}_{i})@>{GH}>{}>(Y^{\prime},p^{\prime},\bar{G})\\ @V{}V{\pi}V@V{}V{\pi}V\\ (r_{i}^{\prime}r_{i}B_{\epsilon_{0}}(p_{i}),p_{i})@>{F_{i}}>{}>(\mathbb{R}^{k},0^{k})\end{CD}

The reason that we can replace GiG_{i}^{\prime} by H~i\widetilde{H}_{i}^{\prime} is that the element not in H~i\widetilde{H}_{i}^{\prime} will disappear in the limit.

Since the index of H~i\widetilde{H}_{i} in H^i\hat{H}_{i} is C(N)\leq C(N), H~i\widetilde{H}_{i} is generated by short elements

H~i(p^i,ϵi)={gH~i|ridi(gp^i,p^i)ϵi}.\widetilde{H}_{i}(\hat{p}_{i},\epsilon_{i})=\{g\in\widetilde{H}_{i}|r_{i}d_{i}(g\hat{p}_{i},\hat{p}_{i})\leq\epsilon_{i}\}.

Then we can use the same method, for constructing Hi,HiH_{i},H_{i}^{\prime} from the diagram 6.2, to define H¯i\bar{H}_{i} and H¯i\bar{H}_{i}^{\prime} from the diagram 6.7 using the gap lemma. We may assume riϵi0r_{i}^{\prime}\epsilon_{i}\to 0, then H¯i=H~i\bar{H}_{i}=\widetilde{H}_{i}. Since H¯i\bar{H}_{i}^{\prime} is a subgroup of H~i\widetilde{H}_{i}^{\prime}, H¯i\bar{H}_{i} is a normal subgroup of H¯i\bar{H}_{i}^{\prime} of index C(N)\leq C(N).

Then by the same construction of H^i\hat{H}_{i} for HiH_{i}, we can construct the groupfication H¯^i\hat{\bar{H}}_{i} using a pseudo-group in H¯i=H~i\bar{H}_{i}=\widetilde{H}_{i}. There is a natural surjective homomorphism si:H¯^iH~is_{i}:\hat{\bar{H}}_{i}\to\widetilde{H}_{i}. The main improvement is that Ker(si)\mathrm{Ker}(s_{i}) must be trivial, since the nilpotency rank of both H¯^i\hat{\bar{H}}_{i} and H^i\hat{H}_{i} are NkN-k. Therefore H~i\widetilde{H}_{i} is isomorphic to the groupfication H¯^i\hat{\bar{H}}_{i}.

Working on the diagram 6.7 with H~i=H¯i\widetilde{H}_{i}=\bar{H}_{i} and H¯i\bar{H}_{i}^{\prime} if necessary, we may assume that HiH_{i} is normal subgroup of HiH_{i}^{\prime} of index C(N)\leq C(N) and H^i=Hi\hat{H}_{i}=H_{i}. ∎

6.4. Proof of Theorem B

Now we can prove that fi:XiKf_{i}:X_{i}\to K is a fibration map. We summarize the differences in this subsection compared with the proof of Lemma 6.3. The first difference is that, since HiH_{i} is isomorphic to the lattice by Lemma 6.7, HiH_{i}^{\prime} can be identified as a discrete subset of 𝒩iAut(𝒩i)\mathcal{N}_{i}\rtimes\text{Aut}(\mathcal{N}_{i}). Then we can construct an infranil-manifold fiber on XiX_{i}. The second difference is that we apply the gluing argument from Theorem 3.5 carefully so that Fi=fiexpp1F_{i}=f_{i}\circ\mathrm{exp}_{p}^{-1} is the a fibration map near pp, thus fif_{i} is a fibration map.

Proof of Theorem B, the existence of the fibration with an infranil-manifold fiber.

Consider the diagram 6.3

(riB~(pi,ϵ0,1),p~i,Hi,Hi)eGH(N,0N,H=Nk,H)ππ(riB~(pi,ϵ0,1)/Hi,pi,Hi/Hi)GH(k,0k,id)ππ(riBϵ0(pi),pi)Fi(k,0k).\begin{CD}(r_{i}\widetilde{B}(p_{i},\epsilon_{0},1),\tilde{p}_{i},H_{i},H_{i}^{\prime})@>{eGH}>{}>(\mathbb{R}^{N},0^{N},H=\mathbb{R}^{N-k},H^{\prime})\\ @V{}V{\pi}V@V{}V{\pi}V\\ (r_{i}\widetilde{B}(p_{i},\epsilon_{0},1)/H_{i}^{\prime},p_{i}^{\prime},H_{i}/H_{i}^{\prime})@>{GH}>{}>(\mathbb{R}^{k},0^{k},\mathrm{id})\\ @V{}V{\pi}V@V{}V{\pi}V\\ (r_{i}B_{\epsilon_{0}}(p_{i}),p_{i})@>{F_{i}}>{}>(\mathbb{R}^{k},0^{k}).\end{CD}

We assumed that Fi=expp1fiF_{i}=\mathrm{exp}_{p}^{-1}\circ f_{i} is not a fibration around pp. The goal is to show that FiF_{i} is a fibration when ii is large enough, thus a contradiction.

Now we apply Lemma 6.7 to the diagram 6.6,

(6.8) (Zi,p~i,Hi=H^i,Hi)eGH(B¯r(0k)×Nk,0N,H=Nk,H)ππ(B¯rri(pi),pi,Hi/Hi)eGH(B¯r(0k),0k,id)ππ(B¯rri(pi),pi)Fi(B¯r(0k),0k)\displaystyle\begin{CD}(Z_{i},\tilde{p}_{i},H_{i}=\hat{H}_{i},H_{i}^{\prime})@>{eGH}>{}>(\bar{B}_{r}(0^{k})\times\mathbb{R}^{N-k},0^{N},H=\mathbb{R}^{N-k},H^{\prime})\\ @V{}V{\pi}V@V{}V{\pi}V\\ (\bar{B}^{r_{i}}_{r}(p_{i}^{\prime}),p_{i}^{\prime},H_{i}/H_{i}^{\prime})@>{eGH}>{}>(\bar{B}_{r}(0^{k}),0^{k},\mathrm{id})\\ @V{}V{\pi}V@V{}V{\pi}V\\ (\bar{B}^{r_{i}}_{r}(p_{i}),p_{i})@>{F_{i}}>{}>(\bar{B}_{r}(0^{k}),0^{k})\end{CD}

By Lemma 6.7, Hi=H^iH_{i}=\hat{H}_{i} is a lattice in 𝒩i\mathcal{N}_{i} and HiH_{i}^{\prime} can be identified as a discrete subset of 𝒩iAut(𝒩i)\mathcal{N}_{i}\rtimes\text{Aut}(\mathcal{N}_{i}). Use the argument in Lemma 4.6 and 4.7, we can construct a global map

h~i:ZiB¯r(0k)×𝒩i\tilde{h}_{i}:Z_{i}\to\bar{B}_{r}(0^{k})\times\mathcal{N}_{i}

is an ϵi\epsilon_{i}-GHA on any rr-ball and ϵi\epsilon_{i}-almost HiH_{i}^{\prime}-equivariant.

Now we can find a normal subgroup Hi′′H_{i}^{\prime\prime} in HiH_{i}^{\prime} with finite index and Br(e)Hi′′=B_{r}(e)\cap H_{i}^{\prime\prime}=\emptyset where e𝒩ie\in\mathcal{N}_{i}. Then (Zi/Hi′′,Hi/Hi′′)(Z_{i}/H_{i}^{\prime\prime},H_{i}^{\prime}/H_{i}^{\prime\prime}) is eGH close to (B¯r(0k)×𝒩i/Hi′′,Hi/Hi′′)(\bar{B}_{r}(0^{k})\times\mathcal{N}_{i}/H_{i}^{\prime\prime},H_{i}^{\prime}/H_{i}^{\prime\prime}) on any rr-ball. We use Theorem 3.5 to construct a Hi/Hi′′H_{i}^{\prime}/H_{i}^{\prime\prime}-equivariant homeomorphism from Zi/Hi′′Z_{i}/H_{i}^{\prime\prime} to B¯r(0k)×𝒩i/Hi′′\bar{B}_{r}(0^{k})\times\mathcal{N}_{i}/H_{i}^{\prime\prime}, dropping some points near the boundary if necessary. Thus Zi/Hi=B¯rri(pi)Z_{i}/H_{i}^{\prime}=\bar{B}^{r_{i}}_{r}(p_{i}), dropping some points near the boundary if necessary, is biHölder homeomorphic to B¯r(0k)×Ni/Hi\bar{B}_{r}(0^{k})\times N_{i}/H_{i}^{\prime}.

Recall that

F~i:riB~(pi,ϵ0,1)𝜋riBϵi(pi)Fi(k,0k),\tilde{F}_{i}:r_{i}\widetilde{B}(p_{i},\epsilon_{0},1)\overset{\pi}{\longrightarrow}r_{i}B_{\epsilon_{i}}(p_{i})\overset{F_{i}}{\longrightarrow}(\mathbb{R}^{k},0^{k}),

is almost kk-splitting. We need to use Theorem 3.5 more carefully to prove that FiF_{i} is exactly the fibration map. The construction in Theorem 3.5 is to glue local almost NN-splitting maps and the group action orbits. Every time we choose local almost NN-splitting map on Zi/Hi′′Z_{i}/H_{i}^{\prime\prime}, we take first k\mathbb{R}^{k}-component exactly to be F~i\tilde{F}_{i}. Then the gluing by the center of mass keeps the value on the first k\mathbb{R}^{k} component, since different local almost NN-splitting or HiH_{i}^{\prime} orbits have the same F~i\tilde{F}_{i} value at a point. Therefore the gluing only changes the value on the 𝒩i\mathcal{N}_{i} component.

In particular, Fi:Br(pi)kF_{i}:B_{r}(p_{i})\to\mathbb{R}^{k} is the exactly k\mathbb{R}^{k} component of the biHölder homeomorphic from Br(pi)B_{r}(p_{i}) to Br(0k)×𝒩i/HiB_{r}(0^{k})\times\mathcal{N}_{i}/H_{i}^{\prime} constructed above,

Fi:Br(pi)homeoBr(0k)×𝒩i/Hi𝜋Br(0k).F_{i}:B_{r}(p_{i})\overset{\mathrm{homeo}}{\longrightarrow}B_{r}(0^{k})\times\mathcal{N}_{i}/H_{i}^{\prime}\overset{\pi}{\longrightarrow}B_{r}(0^{k}).

Therefore FiF_{i} is a fibration map with the infranil-manifold fiber. ∎

We next show that nilpotent structure of 𝒩i\mathcal{N}_{i}, as described in Theorem B, does not depend on the choice of base point on XiX_{i}, which implies that the structure group in Theorem B can be affine; we refer to the gluing arguments in [9, 35] for further details.

Recall that nilpotent structure of 𝒩i\mathcal{N}_{i} in Theorem B is determined by the escape norm on Hi(p~i,r)H_{i}(\tilde{p}_{i},r). Let A=Hi(p~,r)A=H_{i}(\tilde{p},r), then 𝒩i\mathcal{N}_{i} is constructed from a nilprogression P(u1,,uNk;C1,,CNk)P(u_{1},...,u_{N-k};C_{1},...,C_{N-k}) with the escape norm u1Au2AuNkA||u_{1}||_{A}\leq||u_{2}||_{A}\leq...\leq||u_{N-k}||_{A}. And Hi=H^iH_{i}=\hat{H}_{i} is a lattice of 𝒩i\mathcal{N}_{i}.

Passing to a subsequence if necessary, we may assume that uj+1AujA\frac{||u_{j+1}||_{A}}{||u_{j}||_{A}} converges to a real number or \infty as ii\to\infty. Thus we can find KK and 1=l1<l2<<lKNk1=l_{1}<l_{2}<...<l_{K}\leq N-k such that the following two conditions holds for sufficiently large ii:

ujAulAC,ifforsomes,lsl<j<ls+1,\frac{||u_{j}||_{A}}{||u_{l}||_{A}}\leq C,\ \mathrm{if\ for\ some\ }s,l_{s}\leq l<j<l_{s+1},

and

ujAulA,ifforsomes,l<lsj,\frac{||u_{j}||_{A}}{||u_{l}||_{A}}\to\infty,\ \mathrm{if\ for\ some\ s,}\ l<l_{s}\leq j,

as ii\to\infty.

Fix a large ii, define Gs=u1,,uls1HiG_{s}=\langle u_{1},...,u_{l_{s}-1}\rangle\subset H_{i}, 1sK1\leq s\leq K, then Gs/Gs1G_{s}/G_{s-1} is in the center of Hi/Gs1H_{i}/G_{s-1} due to Theorem 2.19 and the construction of the nilprogression. Let 𝒩i,s\mathcal{N}_{i,s} be the simply connected subgroup of 𝒩i\mathcal{N}_{i} with lattice GsG_{s}. Thus (𝒩i/𝒩i,s1)/(Gs/Gs1)(\mathcal{N}_{i}/\mathcal{N}_{i,s-1})/(G_{s}/G_{s-1}) is a torus bundle over 𝒩i/𝒩i,s\mathcal{N}_{i}/\mathcal{N}_{i,s}.

For sufficiently large ii, we shall show that the constructions of 𝒩i\mathcal{N}_{i} and GsG_{s} above are independent of the choice of base point piXip_{i}\in X_{i}. This will allow us to use a gluing argument in [9, 35] to modify the fibration map and reduce the structure group in Theorem B to be affine.

Consider 6.3, assume x~i,y~iB1ri(p~i)\tilde{x}_{i},\tilde{y}_{i}\in B^{r_{i}}_{1}(\tilde{p}_{i}), for any 0<r<1/1000<r<1/100, define

A1={gHi|ridi(g,gx~i)r},A_{1}=\{g\in H_{i}|r_{i}d_{i}(g,g\tilde{x}_{i})\leq r\},
A2={gHi|ridi(g,gy~i)r}.A_{2}=\{g\in H_{i}|r_{i}d_{i}(g,g\tilde{y}_{i})\leq r\}.

We need to show that for any gA1A2g\in A_{1}\cap A_{2} and sufficiently large ii, 12gA1gA22\frac{1}{2}\leq\frac{||g||_{A_{1}}}{||g||_{A_{2}}}\leq 2, which implies that the escape norms by A1A_{1} and A2A_{2} give the same nilpotent structure of HiH_{i}.

By the definition of escape norm, g1/gA1A1g^{1/||g||_{A_{1}}}\notin A_{1}. Then d(g1/gA1x~i,x~i)rd(g^{1/||g||_{A_{1}}}\tilde{x}_{i},\tilde{x}_{i})\geq r. Since the limit of HiH_{i} is a free translation group Nk\mathbb{R}^{N-k} by Lemma 6.1, d(g1/gA1y~i,y~)2r3d(g^{1/||g||_{A_{1}}}\tilde{y}_{i},\tilde{y})\geq\frac{2r}{3}. Then d(g2/gA1y~i,)rd(g^{2/||g||_{A_{1}}}\tilde{y}_{i},)\geq r. In particular, gA2gA12||g||_{A_{2}}\geq\frac{||g||_{A_{1}}}{2}. Similarly gA1gA22||g||_{A_{1}}\geq\frac{||g||_{A_{2}}}{2}. Thus we have proved that the nilpotent structure of the fiber does not depend on the choice of base point in a small neighborhood of pip_{i}. Then by a connectedness argument, we conclude that the nilpotent structure of the fiber is independent of the choice of base point on XiX_{i} for sufficiently large ii, thus the structure group in Theorem B is affine.

7. Limit of RCD(N1,N)(N-1,N) spaces with bounded covering geometry

In this section we consider a sequence of pointed RCD((N1),N)(-(N-1),N) spaces (Xi,di,N,pi)(X_{i},d_{i},\mathcal{H}^{N},p_{i}) with (1,v)(1,v)-bound covering geometry and assume the following convergence:

(7.1) (Xi,di,N,pi)pmGH(X,d,𝔪,p).(X_{i},d_{i},\mathcal{H}^{N},p_{i})\overset{\mathrm{pmGH}}{\longrightarrow}(X,d,\mathfrak{m},p).

Assume that the rectifiable dimension of XX is kk. We want to show that any kk-regular point in XX is a manifold point.

Theorem 7.1.

In the context of 7.1, assume that xXx\in X is a kk-regular point. Then there exists a neighborhood of xx which is biHölder homeomorphic to an open ball in k\mathbb{R}^{k}.

Let xiXix_{i}\in X_{i} converge to xXx\in X. Let GiG_{i} be the image of natural homomoephism

π1(B12(xi),xi)π1(B1(xi),xi).\pi_{1}(B_{\frac{1}{2}}(x_{i}),x_{i})\to\pi_{1}(B_{1}(x_{i}),x_{i}).

Then let B~(xi,12,1)\widetilde{B}(x_{i},\frac{1}{2},1) be a connected component of the pre-image of B12(xi)B_{\frac{1}{2}}(x_{i}) in the universal cover of B1(xi)B_{1}(x_{i}). Take x~iB~(xi,12,1)\tilde{x}_{i}\in\widetilde{B}(x_{i},\frac{1}{2},1) as a lift of xix_{i}. By the pre-compactness theorem 2.9, passing to a subsequence if necessary,

(7.2) (B~(xi,12,1),p~i,Gi)eGH(Y,x~,G)ππ(B12(xi),xi)GH(B¯12(x),x)\displaystyle\begin{CD}(\widetilde{B}(x_{i},\frac{1}{2},1),\tilde{p}_{i},G_{i})@>{eGH}>{}>(Y,\tilde{x},G)\\ @V{}V{\pi}V@V{}V{\pi}V\\ (B_{\frac{1}{2}}(x_{i}),x_{i})@>{GH}>{}>(\bar{B}_{\frac{1}{2}}(x),x)\end{CD}

Since xx is a kk-regular point, x~\tilde{x} must be a NN-regular point by [23].

We next show that an almost kk-splitting map near xx can be lifted to an almost kk-splitting map near x~\tilde{x}.

Lemma 7.2.

In the context of 7.1, there exists C(N)>0C(N)>0 so that the following holds. Assume that there exist small ϵ,r>0\epsilon,r>0, and

f:Br(x)kf:B_{r}(x)\to\mathbb{R}^{k}

is harmonic and (k,ϵ)(k,\epsilon)-splitting. Define the map

f~:Br2(x~)k,f~=fπ.\tilde{f}:B_{\frac{r}{2}}(\tilde{x})\to\mathbb{R}^{k},\ \tilde{f}=f\circ\pi.

Then f~\tilde{f} is harmonic and (k,Cϵ)(k,C\epsilon)-splitting.

Proof.

By Corollary 4.12 in [1], we can find harmonic maps

fi:B2r3(xi)kf_{i}:B_{\frac{2r}{3}}(x_{i})\to\mathbb{R}^{k}

converging to f|B2r3(xi)f_{|B_{\frac{2r}{3}}(x_{i})} in the H1,2H^{1,2} sense. In particular, fif_{i} is (k,2ϵ)(k,2\epsilon)-splitting for large ii.

Define

f~i:B2r3(x~i)k,f~i=fiπ.\tilde{f}_{i}:B_{\frac{2r}{3}}(\tilde{x}_{i})\to\mathbb{R}^{k},\ \tilde{f}_{i}=f_{i}\circ\pi.

Then f~i\tilde{f}_{i} is a harmonic and (k,Cϵ)(k,C\epsilon)-splitting by the covering lemma 2.12. Passing to a subsequence if necessary, by Theorem 4.4 in [1], f~i\tilde{f}_{i} has a H1,2H^{1,2} limit g~\tilde{g} which is harmonic on Br2(x~)B_{\frac{r}{2}}(\tilde{x}). Then g~\tilde{g} is also (k,Cϵ)(k,C\epsilon)-splitting. By the equivariant convergence, f=g~πf=\tilde{g}\circ\pi on Br2(x)B_{\frac{r}{2}}(x), thus f~=g~\tilde{f}=\tilde{g} is harmonic and (k,Cϵ)(k,C\epsilon)-splitting. ∎

We prove that the geometric transformation theorem holds at regular points in XX.

Lemma 7.3.

In the context of 7.1, for any δ>0\delta>0, there exists ϵ>0\epsilon>0 so that the following holds. Assume that there exists r01r_{0}\leq 1 and a (k,ϵ)(k,\epsilon)-splitting map

f:Br0(x)k.f:B_{r_{0}}(x)\to\mathbb{R}^{k}.

Then for any sr010s\leq\frac{r_{0}}{10}, there exists an k×kk\times k lower triangular matrix TsT_{s} such that

Ts(f):Bs(x)kT_{s}(f):B_{s}(x)\to\mathbb{R}^{k}

is a (k,δ)(k,\delta)-splitting and |Ts|sδ|T_{s}|\leq s^{-\delta}.

Proof.

Take a small ϵ>0\epsilon>0 to be decided later. Since xx is regular, there exists r0>0r_{0}>0 and a (k,ϵ)(k,\epsilon)-splitting map

f:Br0(x)k.f:B_{r_{0}}(x)\to\mathbb{R}^{k}.

Then f~=fπ\tilde{f}=f\circ\pi is (k,Cϵ)(k,C\epsilon)-splitting by Lemma 7.2.

The rectifiable dimension of XX is kk. By [29], any tangent cone of XX can not split an k+1\mathbb{R}^{k+1} factor. In particular, Br0(x)B_{r_{0}}(x) is r0Φ(ϵ|k,N)r_{0}\Phi(\epsilon|k,N)-close to a r0r_{0}-ball in k\mathbb{R}^{k}. By [23], Br0(x~)B_{r_{0}}(\tilde{x}) is r0Φ(ϵ|k,N,v)r_{0}\Phi(\epsilon|k,N,v)-close to r0r_{0}-ball in N\mathbb{R}^{N}.

Now we can find

h~:Br02(x~)Nk\tilde{h}:B_{\frac{r_{0}}{2}}(\tilde{x})\to\mathbb{R}^{N-k}

such that the pair (f~,h~):Br02(x~)N(\tilde{f},\tilde{h}):B_{\frac{r_{0}}{2}}(\tilde{x})\to\mathbb{R}^{N} forms a (N,Φ(ϵ|k,N,v))(N,\Phi(\epsilon|k,N,v))-splitting map. Apply the transformation theorem 2.6 to (f~,h~)(\tilde{f},\tilde{h}). When ϵ\epsilon is small enough, for any sr010s\leq\frac{r_{0}}{10}, we can find a N×NN\times N lower triangular matrix T~s\tilde{T}_{s} such that

T~s(f~,h~):B3s(x~)N\tilde{T}_{s}(\tilde{f},\tilde{h}):B_{3s}(\tilde{x})\to\mathbb{R}^{N}

a (N,δ)(N,\delta)-splitting and |T~s|sδ|\tilde{T}_{s}|\leq s^{-\delta}.

Take TsT_{s} to be the upper left k×kk\times k submatrix of T~s\tilde{T}_{s}. Then TsT_{s} is a k×kk\times k lower triangular matrix and Tsf~T_{s}\tilde{f} is a (k,δ)(k,\delta)-splitting on B3s(x~)B_{3s}(\tilde{x}). By a similar argument in Lemma 7.2, we conclude that TsfT_{s}f is (N,Cδ)(N,C\delta)-splitting on Bs(x)B_{s}(x). Moreover, |Ts||T~s|sδ|T_{s}|\leq|\tilde{T}_{s}|\leq s^{-\delta}. ∎

Proof of Theorem 7.1.

The proof of Theorem 7.1 follows from the proof of canonical Reifenberg theorem in [12] and [22]. Take any small δ>0\delta>0, by Lemma 7.3, there exists r0>0r_{0}>0 and a (k,ϵ)(k,\epsilon)-splitting map

f:B4r0(x)kf:B_{4r_{0}}(x)\to\mathbb{R}^{k}

so that for each yB2r0(x)y\in B_{2r_{0}}(x) and sr0s\leq r_{0}, there exists a k×kk\times k lower triangular matrix Ts,yT_{s,y} so that

Ts,yf:Bs(y)kT_{s,y}f:B_{s}(y)\to\mathbb{R}^{k}

is a (k,δ)(k,\delta)-splitting map with |Ts,y|sδ|T_{s,y}|\leq s^{-\delta}.

We shall show that ff is biHölder from Br0(x)B_{r_{0}}(x) to its image. For any y1,y2Br0(x)y_{1},y_{2}\in B_{r_{0}}(x), take s=d(y1,y2)s=d(y_{1},y_{2}). Since Ts,y1f:Bs(y1)kT_{s,y_{1}}f:B_{s}(y_{1})\to\mathbb{R}^{k} is (k,δ)(k,\delta)-splitting and the dimension of XX is kk, it must be a Φ(δ|k)s\Phi(\delta|k)s-GHA. Thus

d(Ts,y1f(y1),Ts,y1f(y2))(1Φ(δ|k))d(y1,y2).d(T_{s,y_{1}}f(y_{1}),T_{s,y_{1}}f(y_{2}))\geq(1-\Phi(\delta|k))d(y_{1},y_{2}).

Since |Ts,y1|sδ|T_{s,y_{1}}|\leq s^{-\delta} and s=d(y1,y2)s=d(y_{1},y_{2}),

d(f(y1),f(y2))sδd(Ts,y1f(y1),Ts,y1f(y2))(1Φ(δ|k,N))d(y1,y2)1+δ.d(f(y_{1}),f(y_{2}))\geq s^{\delta}d(T_{s,y_{1}}f(y_{1}),T_{s,y_{1}}f(y_{2}))\geq(1-\Phi(\delta|k,N))d(y_{1},y_{2})^{1+\delta}.

On the other hand, since ff is harmonic, |fa|1+Φ(δ|k)|\nabla f^{a}|\leq 1+\Phi(\delta|k), for any a=1,2,ka=1,2,...k, thus

d(f(y1),f(y2))(1+Φ(δ|k))d(y1,y2).d(f(y_{1}),f(y_{2}))\leq(1+\Phi(\delta|k))d(y_{1},y_{2}).

The biHölder estimate holds, completing the proof. ∎

Proof of Theorem 1.9.

The part (a) is exactly Theorem 7.1 and the part (b) follows from the construction in Theorem B. ∎

References

  • [1] L. Ambrosio and S. Honda. Local spectral convergence in RCD*(K,N) spaces. Nonlinear Anal. 177 (2018), 1–23., 2018.
  • [2] Michael T. Anderson. Hausdorff perturbations of Ricci-flat manifolds and the splitting theorem. Duke Mathematical Journal, 68(1):67 – 82, 1992.
  • [3] E Breuillard, B Green, and T Tao. The structure of approximate groups. Publ. Math. Inst. Hautes Etudes Sci., 116:115–221, 2012.
  • [4] E. Brue, A. Naber, and D. Semola. Boundary regularity and stability for spaces with Ricci bounded below. Invent. Math. 228 (2022) no. 2, 777–891, 2022.
  • [5] E. Brue and D. Semola. Constancy of the dimension in codimension one and locality of the unitnormal on RCD(K,N) spaces. Ann. Sc. Norm. Super. Pisa Cl. Sci. (5) 24 (2023), no.3, 1765-1816., 2023.
  • [6] Peter Buser and Hermann Karcher. Gromovs almost flat manifolds. Société mathématique de France, 1981.
  • [7] Jeff Cheeger and Tobias H. Colding. On the structure of spaces with Ricci curvature bounded below. i. J. Differential Geom., 46(3):406–480, 1997.
  • [8] Jeff Cheeger and Tobias H. Colding. On the structure of spaces with Ricci curvature bounded below. ii. J. Differential Geom., 54(1):13–35, 2000.
  • [9] 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.
  • [10] Jeff Cheeger and Mikhael Gromov. Collapsing Riemannian manifolds while keeping their curvature bounded. I. Journal of Differential Geometry, 23(3):309–346, 1986.
  • [11] Jeff Cheeger and Mikhael Gromov. Collapsing Riemannian manifolds while keeping their curvature bounded. II. Journal of Differential Geometry, 32(1):269– 298, 1990.
  • [12] 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.
  • [13] Jeff Cheeger and Aaron Naber. Regularity of Einstein manifolds and the codimension 4 conjecture. Ann. of Math., 182(3):1093–1165, 2015.
  • [14] Q. Deng, J. Santos-Rodríguez, S. Zamora, and X. Zhao. Margulis Lemma on RCD(K,N)(K,N) spaces. arXiv preprint, 2023.
  • [15] Kenji Fukaya. Theory of convergence for riemannian orbifolds. Japan. J. Math. (N.S.), 12(1):121–160, 1986.
  • [16] Kenji Fukaya. Collapsing Riemannian manifolds to ones of lower dimensions. Journal of Differential Geometry, 25(1):139–156, 1987.
  • [17] Kenji Fukaya. A boundary of the set of the Riemannian manifolds with bounded curvatures and diameters. Journal of Differential Geometry, 28(1):1–21, 1988.
  • [18] Kenji Fukaya. Collapsing Riemannian manifolds to ones with lower dimension II. Journal of the Mathematical Society of Japan, 41(2):333–356, 1989.
  • [19] Kenji Fukaya and Takao Yamaguchi. The fundamental groups of almost nonnegatively curved manifolds. Annals of Mathematics, 136(2):253–333, 1992.
  • [20] Mikhael Gromov. Almost flat manifolds. J. Diff. Geom., 13(2):231–241, 1978.
  • [21] Luis Guijarro and Jaime Santos-Rodríguez. On the isometry group of RCD*(K,N)-spaces. manuscripta mathematica, 158:441–461, 2019.
  • [22] S. Honda and Y. Peng. A note on the topological stability theorem from spaces to Riemannian manifolds. Manuscripta Math. 172 (2023), no.3-4, 971-1007., 2023.
  • [23] Hongzhi Huang. Fibrations, and stability for compact group actions on manifolds with local bounded Ricci covering geometry. Front. Math. China, 15(1):69–89, 2020.
  • [24] 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, 2022.
  • [25] Erik Hupp, Aaron Naber, and Kai-Hsiang Wang. Lower Ricci curvature and nonexistence of manifold structure. arXiv:2308.03909, 2023.
  • [26] Vitali Kapovitch. Mixed curvature almost flat manifolds. Geom. Topol. 25 (2021) 2017-2059, 2021.
  • [27] Vitali Kapovitch, Martin Kell, and Christian Ketterer. On the structure of RCD spaces with upper curvature bounds. Mathematische Zeitschrift, 2022, No 4, p. 3469-3502, 2022.
  • [28] Vitali Kapovitch and Burkhard Wilking. Structure of fundamental groups of manifolds with Ricci curvature bounded below. arXiv:1105.5955, 2011.
  • [29] Yu Kitabeppu. A sufficient condition to a regular set of positive measure on RCD spaces. Potential Anal. 51 (2019), no. 2, 179–196., 2019.
  • [30] K.B. Lee and F. Raymond. Rigidity of almost crystallographic groups. Combinatorial methods in topology and algebraic geometry (Rochester, N.Y., 1982), Contemp. Math., vol. 44, Amer. Math. Soc., Providence, RI, 1985, pp. 73-78. ISBN:0-8218-5039-3., 1985.
  • [31] Andrea Mondino and Guofang Wei. On the universal cover and the fundamental group of an RCD(K,N)*(K,N)-space. J. Reine Angew. Math., 2019(753):211–237, 2019.
  • [32] Jiayin Pan and Jikang Wang. Some topological results of Ricci limit spaces. Transactions of the American Mathematical Society, 375(12):8445–8464, 2022.
  • [33] G. De Philippis and N. Gigli. Non-collapsed spaces with Ricci curvature bounded from below. J.Ec. polytech. Math. 5 (2018), 613-650., 2018.
  • [34] Xiaochun Rong. A new proof of the Gromov’s theorem on almost flat manifolds. arXiv:1906.03377v2, 2019.
  • [35] Xiaochun Rong. Collapsed manifolds with local ricci bounded covering geometry. arXiv:2211.09998, 2022.
  • [36] Ernst A. Ruh. Almost flat manifolds. J. Differential Geom., 17(1):1–14, 1982.
  • [37] Jaime Santos-Rodriguez and Sergio Zamora. On fundamental groups of RCD spaces. Journal für die reine und angewandte Mathematik (Crelles Journal), vol. 2023, no. 799, 2023, pp. 249-286., 2023.
  • [38] Christina Sormani and Guofang Wei. Universal covers for Hausdorff limits of noncompact spaces. Transactions of the American Mathematical Society, 356(3):1233–1270, 2004.
  • [39] Gerardo Sosa. The isometry group of an RCD* space is Lie. Potential Analysis, 49:267–286, 2018.
  • [40] Jikang Wang. On the limit of simply connected manifolds with discrete isometric cocompact group actions. arXiv:2307.07658, 2023.
  • [41] Jikang Wang. RCD*(K,N) spaces are semi-locally simply connected. Journal für die reine und angewandte Mathematik (Crelles Journal), vol. 2024, no. 806, 2024, pp. 1-7., 2024.
  • [42] Jikang Wang. Ricci limit spaces are semi-locally simply connected. Journal of Differential Geometry, J. Differential Geom. 128(3), 1301-1314, (November 2024), 2024.
  • [43] Shicheng Xu. Precompactness of domains with lower Ricci curvature bound under Gromov-Hausdorff topology. arXiv:2311.05140, 2023.
  • [44] Sergio Zamora. Limits of almost homogeneous spaces and their fundamental groups. Groups Geom. Dyn. 18 (2024), no. 3, pp. 761–798, 2024.
  • [45] Sergio Zamora and Xingyu Zhu. Topological rigidity of small RCD(K,N) spaces with maximal rank. arXiv:2406.10189, 2024.
  • [46] Shengxuan Zhou. Examples of Ricci limit spaces with infinite holes. arXiv:2404.00619, 2024.