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

Convergence of Ricci-limit spaces under bounded Ricci curvature and local covering geometry I

Zuohai Jiang Beijing international center for mathematical research, peking university, Beijing, China [email protected] Lingling Kong School of Mathematical and statistical, Northeast Normal Universiy, Changchun China [email protected]  and  Shicheng Xu School of Mathematical Sciences, Capital Normal Universiy, Beijing China Academy for Multidisciplinary Studies, Capital Normal University, Beijing China [email protected]
Abstract.

We extend Cheeger-Gromov’s and Anderson’s convergence theorems to regular limit spaces of manifolds with bounded Ricci curvature and local covering geometry, by establishing the C1,αC^{1,\alpha}-regularities that are the best one may expect on those Ricci-limit spaces. As an application we prove an optimal generalization of Fukaya’s fibration theorem on collapsed manifolds with bounded Ricci curvature, which also improves the original version to C1,αC^{1,\alpha} limit spaces.

2010 Mathematics Subject Classification:
53C23, 53C21, 53C20, 5324
The second author would like to thank Capital Normal University for a warm hospitality during his visit

0. Introduction

One of the fundamental tools in the study of geometry and topology of manifolds with curvature bound is the Cheeger-Gromov’s convergence theorem [8, 34], which implies that the space n(v,D)\mathcal{M}^{n}(v,D) of all closed Riemannian nn-manifolds of sectional curvature |sec|1|\sec|\leq 1, volume v>0\geq v>0 and diameter D<+\leq D<+\infty is precompact in the C1,αC^{1,\alpha}-topology (cf. also [53, 32, 47]). This C1,αC^{1,\alpha}-convergence theorem had been generalized by Anderson [4] (cf. [31]) and Anderson-Cheeger [5] to Riemannian manifolds with (lower) bounded Ricci curvature. That is, the space |Ric|n(i0,D)\mathcal{M}_{|Ric|}^{n}(i_{0},D) (resp. Ricn(i0,D)\mathcal{M}_{Ric}^{n}(i_{0},D)) consisting of all closed Riemannian nn-manifolds of Ricci curvature |Ric|n1|\operatorname{Ric}|\leq n-1 (resp. Ric(n1)\operatorname{Ric}\geq-(n-1)), injectivity radius i0>0\geq i_{0}>0 and diameter D<+\leq D<+\infty is precompact in the C1,αC^{1,\alpha}-topology (resp. C0,αC^{0,\alpha}-topology). A weaker replacement of the injectivity radius is (δ,ρ)(\delta,\rho)-Reifenberg condition [11] (cf. [40]) for 0δδ(n)0\leq\delta\leq\delta(n), i.e., around each point xx in such a manifold the Gromov-Hausdorff distance

dGH(Br(x),Brn(0))δr,for all 0<rρ,d_{GH}\left(B_{r}(x),B_{r}^{n}(0)\right)\leq\delta\cdot r,\qquad\text{for all $0<r\leq\rho$}, (0.1)

where Brn(0)B_{r}^{n}(0) is the rr-ball at the origin in n\mathbb{R}^{n}, δ(n)\delta(n) is a constant depending only on nn, and ρ\rho is a small positive constant. A point satisfying (0.1) is called a (δ,ρ)(\delta,\rho)-Reifenberg point. By Anderson [4] and Colding [21], it was well known that Anderson’s convergence theorem still holds after replacing the lower bound of injectivity radius by the (δ,ρ)(\delta,\rho)-Reifenberg condition.

By the convergence theorems above, a limit space of manifolds in Ricn(i0,D)\mathcal{M}^{n}_{Ric}(i_{0},D) (resp. |Ric|n(i0,D)\mathcal{M}^{n}_{|Ric|}(i_{0},D)) in the Gromov-Hausdorff topology is always a smooth manifold with a CαC^{\alpha}-Riemannian metric (resp. C1,αC^{1,\alpha}-Riemannian metric). However, without a positive injectivity radius bound a Ricci limit space may contain non-Euclidean points. A collapsed Ricci limit space may even have non integer Hausdorff dimension, see [51]. The geometric properties of Ricci limit spaces had played a fundamental role in solving some open problems and conjectures on manifolds with (lower) Ricci curvature bound; see for example [10], [11], [46].

A Riemannian manifold (M,g)(M,g) with normalized curvature bound is called ϵ\epsilon-collapsed, if the volume of any unit ball B1(x)B_{1}(x) at xMx\in M is less than ϵ\epsilon. As the counterpart of Cheeger-Gromov’s convergence theorem, ϵ\epsilon-collapsed Riemannian manifolds with bounded sectional curvature had been extensively studied by Cheeger-Fukaya-Gromov [12] (cf. Cheeger-Gromov [14, 15], and Fukaya [27, 28, 29]). In contrast, the geometry and topology of collapsed manifolds with bounded Ricci curvature are much more complicated and rarely known at present. Many progresses on the collapsed manifolds with lower bounded Ricci curvature are achieved recently under some additional geometric assumptions, such as bounded local covering geometry, see [57], [18, 17], [50], [40], [39], [42, 43], etc. and the survey paper [41].

According to [40] (cf. [16, 13]), a complete Riemannian nn-manifold (M,g)(M,g) with |RicM|n1|\operatorname{Ric}_{M}|\leq n-1 is said to have (r,v)(r,v)-local covering geometry, if for any xMx\in M, the local rewinding volume of Br(x)B_{r}(x), Vol(Br(x~))vrn>0\operatorname{Vol}(B_{r}(\tilde{x}))\geq vr^{n}>0, where x~\tilde{x} is a preimage point of xx in the (incomplete) Riemannian universal cover π:(Br(x)~,x~)(Br(x),x)\pi\mathrel{\mathop{\mathchar 58\relax}}(\widetilde{B_{r}(x)},\tilde{x})\to(B_{r}(x),x). Moreover, (M,g)(M,g) is said to have (δ,r)(\delta,r)-Reifenberg local covering geometry, if for any xMx\in M, x~\tilde{x} is a (δ,r)(\delta,r)-Reifenberg point. By Cheeger-Fukaya-Gromov [12, Theorem 1.3], for any δ>0\delta>0, there is r(n,δ)>0r(n,\delta)>0 such that any collapsed complete manifold with sectional curvature bound |sec|1|\sec|\leq 1 admits (δ,r(n,δ))(\delta,r(n,\delta))-Reifenberg local covering geometry.

In this paper, we prove the C1,αC^{1,\alpha}-regularity of those regular limit spaces under bounded Ricci curvature and local covering geometry, which naturally extends the above Cheeger-Gromov’s and Anderson’s convergence theorems.

Let 𝒳n,r,vm(δ,ρ)\mathcal{X}_{n,r,v}^{m}(\delta,\rho) be the set consisting of all compact Ricci-limit spaces of Riemannian nn-manifolds with |RicM|n1|\operatorname{Ric}_{M}|\leq n-1 and (r,v)(r,v)-local covering geometry, such that the dimension of XX is mm in the sense of Colding-Naber [22], and any point xx in X𝒳n,r,vm(δ,ρ)X\in\mathcal{X}_{n,r,v}^{m}(\delta,\rho) is (δ,ρ)(\delta,\rho)-Reifenberg.

Theorem 0.1 (bounded local covering geometry).

Given r,v,ρ>0r,v,\rho>0 and positive integers mnm\leq n, there are constants δ=δ(n),T(n,r,v,ρ)>0\delta=\delta(n),T(n,r,v,\rho)>0 such that the followings hold.

  1. (0.1.1)

    Any X𝒳n,r,vm(δ,ρ)X\in\mathcal{X}_{n,r,v}^{m}(\delta,\rho) is a C1,αC^{1,\alpha}-Riemannian manifold (X,h)(X,h) with a positive C1,αC^{1,\alpha}-harmonic radius r0(n,r,v,ρ,α)>0\geq r_{0}(n,r,v,\rho,\alpha)>0 for any 0<α<10<\alpha<1. And for each 0<ϵT(n,r,v,ρ)0<\epsilon\leq T(n,r,v,\rho), there is a nearby Riemannian metric h(ϵ)h(\epsilon) on (X,h)(X,h) such that

    1. (a)

      h(ϵ)hC1,αΨ(ϵ|n,r,v,ρ,α)\|h(\epsilon)-h\|_{C^{1,\alpha}}\leq\Psi(\epsilon\,|\,n,r,v,\rho,\alpha), where Ψ(ϵ|n,r,v,ρ,α)0\Psi(\epsilon\,|\,n,r,v,\rho,\alpha)\to 0 as ϵ0\epsilon\to 0 with other parameters n,,αn,\dots,\alpha fixed,

    2. (b)

      the sectional curvature of h(ϵ)h(\epsilon) satisfies |sech(ϵ)|C(n,r,v,ρ)ϵ12|\sec_{h(\epsilon)}|\leq C(n,r,v,\rho)\epsilon^{-\frac{1}{2}},

    3. (c)

      the kthk^{\text{th}}-derivative of curvature tensor |kRm(h(ϵ))|h(ϵ)C(n,r,v,ρ,k,ϵ),\left|\nabla^{k}\operatorname{Rm}(h(\epsilon))\right|_{h(\epsilon)}\leq C(n,r,v,\rho,k,\epsilon),

    where the constants r0(n,r,)r_{0}(n,r,\dots) and C(n,r,)C(n,r,\dots) depend only on the given parameters.

  2. (0.1.2)

    The subset 𝒳n,r,vm(δ,ρ,D)={X𝒳n,r,vm(δ,ρ):diam(X)D}\mathcal{X}_{n,r,v}^{m}(\delta,\rho,D)=\{X\in\mathcal{X}_{n,r,v}^{m}(\delta,\rho)\mathrel{\mathop{\mathchar 58\relax}}\operatorname{diam}(X)\leq D\} is compact in the C1,αC^{1,\alpha}-topology. In particular, 𝒳n,r,vm(δ,ρ,D)\mathcal{X}_{n,r,v}^{m}(\delta,\rho,D) contains only finitely many diffeomorphism types.

Recently Naber and Zhang [50] proved the ϵ\epsilon-regularity theorem for locally full-rank collapsed Riemannian manifolds with (lower) bounded Ricci curvature, which says that there are uniform constants ϵ=ϵ(n),r(n,α)>0\epsilon=\epsilon(n),r(n,\alpha)>0 such that for any complete Riemannan nn-manifold MM with |RicM|n1|\operatorname{Ric}_{M}|\leq n-1 and 0<ρ10<\rho\leq 1, if an open ball Bρ(x)B_{\rho}(x) is ϵρ\epsilon\rho-Gromov-Hausdorff close to a ρ\rho-ball in a lower dimensional Euclidean space m\mathbb{R}^{m}, then the subgroup Γϵ,ρ\Gamma_{\epsilon,\rho} generated by loops at xx of length <2ϵρ<2\epsilon\rho in the fundamental group π1(Bρ(x),x)\pi_{1}(B_{\rho}(x),x) has the full rank nmn-m if and only if the preimages of xx in the universal cover of Bρ/2(x)B_{\rho/2}(x) admit a uniform C1,αC^{1,\alpha}-harmonic radius r(n,α)ρ>0\geq r(n,\alpha)\rho>0. Recall that by the generalized Margulis lemma [46], the local fundamental group Γϵ,ρ\Gamma_{\epsilon,\rho} contains a nilpotent subgroup NN of finite index ω(n)\leq\omega(n), where the rank of Γϵ,ρ\Gamma_{\epsilon,\rho} is defined to be that of NN, which is no more than nmn-m (see e.g., [50, Theorem 2.27]).

Let 𝒴nm(δ,ρ)\mathcal{Y}_{n}^{m}(\delta,\rho) be the set consisting of all compact Ricci-limit spaces of locally full-rank collapsed manifolds with |Ric|n1|\operatorname{Ric}|\leq n-1, such that the dimension of X𝒴nm(δ,ρ)X\in\mathcal{Y}^{m}_{n}(\delta,\rho) is mm, and any point in XX is (δ,ρ)(\delta,\rho)-Reifenberg. By Naber-Zhang’s ϵ\epsilon-regularity, 𝒴nm(δ,ρ)𝒳n,r,vm(δ,ρ)\mathcal{Y}_{n}^{m}(\delta,\rho)\subset\mathcal{X}_{n,r,v}^{m}(\delta,\rho) for some r=r(n,ρ)r=r(n,\rho) and v>0v>0.

Corollary 0.2 (locally full-rank collapsed limit spaces).

Let δ=δ(n)>0\delta=\delta(n)>0 be the constant in Theorem 0.1. The conclusions in Theorem 0.1 hold for 𝒴nm(δ,ρ)\mathcal{Y}_{n}^{m}(\delta,\rho), where the dependence of constants T,CT,C and r0r_{0} on n,r,v,ρn,r,v,\rho can be simplified to nn and ρ\rho only.

Since all closed manifolds in |Ric|m(i0,D)\mathcal{M}_{|Ric|}^{m}(i_{0},D) are contained in 𝒴nm(δ(n),ρ,D)\mathcal{Y}^{m}_{n}(\delta(n),\rho,D) for some ρ=ρ(n,i0)\rho=\rho(n,i_{0}), Theorem 0.1 generalizes Anderson’s C1,αC^{1,\alpha}-convergence theorem.

Let us recall that in general, a positive lower volume bound for manifolds with bounded Ricci curvature and diameter is weaker than a positive injectivity radius bound, though they are equivalent under sectional curvature bound. Similar to the Anderson’s [4] and Anderson-Cheeger’s [5] convergence theorems, Theorem 0.1 fails after loosing the (δ,ρ)(\delta,\rho)-Reifenberg condition to a positive volume lower bound on limit spaces; it shares the same counterexamples as those for Anderson’s convergence (see [4]).

Next, we will generalize the Cheeger-Gromov’s convergence to limit spaces of manifolds with bounded Ricci curvature and Reifenberg-bounded local covering geometry, where the Reifenberg condition (resp. positive injectivity radius) on the limit spaces is replaced by a positive lower bound of the mm-Hausdorff measure (resp. the volume).

Let 𝒵n,δ,rm(τ)\mathcal{Z}_{n,\delta,r}^{m}(\tau) be the set consisting of all compact Ricci-limit spaces of Riemannian nn-manifolds with |RicM|n1|\operatorname{Ric}_{M}|\leq n-1 and (δ,r)(\delta,r)-Reifenberg local covering geometry, such that each element X𝒵n,δ,rm(τ)X\in\mathcal{Z}_{n,\delta,r}^{m}(\tau) is τ\tau-almost regular in the sense that for any xXx\in X, any tangent cone (TxX,x)(T_{x}X,x) at xx is τ\tau-close to (m,0)(\mathbb{R}^{m},0) in the pointed Gromov-Hausdorff topolgy.

Theorem 0.3 (Reifenberg-bounded local covering geometry).

Let δ=δ(n)>0\delta=\delta(n)>0 be the constant in Theorem 0.1.

  1. (0.3.1)

    Any X𝒵n,δ,rm(δ)X\in\mathcal{Z}_{n,\delta,r}^{m}(\delta) is a C1,αC^{1,\alpha}-Riemannian manifold (X,h)(X,h) such that for any point xXx\in X, Vol(B1(x))w>0\operatorname{Vol}(B_{1}(x))\geq w>0 implies that the C1,αC^{1,\alpha}-harmonic radius at xx is no less than rh(n,r,w,α)>0r_{h}(n,r,w,\alpha)>0 for any 0<α<10<\alpha<1.

  2. (0.3.2)

    Let 𝒵n,δ,rm(δ,w,D)={X𝒵n,δ,rm(δ):diam(X)D,Vol(X)w>0}\mathcal{Z}_{n,\delta,r}^{m}(\delta,w,D)=\{X\in\mathcal{Z}_{n,\delta,r}^{m}(\delta)\mathrel{\mathop{\mathchar 58\relax}}\operatorname{diam}(X)\leq D,\operatorname{Vol}(X)\geq w>0\}. There is ρ=ρ(n,r,w,D)>0\rho=\rho(n,r,w,D)>0 such that 𝒵n,δ,rm(δ,w,D)𝒴nm(δ,ρ,D)\mathcal{Z}_{n,\delta,r}^{m}(\delta,w,D)\subset\mathcal{Y}_{n}^{m}(\delta,\rho,D).

    In particular, the conclusions in Theorem 0.1 hold for 𝒵n,δ,rm(δ,w,D)\mathcal{Z}_{n,\delta,r}^{m}(\delta,w,D).

Theorem 0.3 generalizes the local estimate [16], [19] on the injectivity radius of manifolds with bounded sectional curvature to harmonic radius of limit spaces under bounded Ricci curvature and Reifenberg local covering geometry.

We point out that the Theorem 0.3 fails for regular Ricci-limit spaces under bounded Ricci curvature and (r,v)(r,v)-local covering geometry in the sense of rewinding volume, since it would contains all non-collapsed Ricci-flat manifolds.

Remark 0.4.

For those regular limit spaces of collapsed manifolds with two-sided bounded sectional curvature, the C1,αC^{1,\alpha}-regularity and compactness in Theorems 0.1-0.3 are well-known to experts and can be easily derived by [28]. In fact, they are direct corollaries of Cheeger-Gromov’s convergence theorem, because those limit spaces of Hausdorff dimension mm can be smoothed to metrics whose sectional curvature is bounded two-sided uniformly by C(n,w,D)C(n,w,D), where diam(X)D\operatorname{diam}(X)\leq D, and mm-Hausdorff measure Hm(X)w>0H^{m}(X)\geq w>0; for details see Section 2, and also compare [28, Theorem 0.9] and [28, Corollary 0.11], which states XX is a smooth manifold with a continuous metric tensor hh inducing a C1,αC^{1,\alpha} distance function.

However, the smoothed metrics on limit spaces of manifolds with bounded Ricci curvature generally admit no uniformly bounded sectional curvature. In order to derive the C1,αC^{1,\alpha}-convergence, one has to construct C1,αC^{1,\alpha}-harmonic coordinates directly on a limit space. This is the new ingredient in Theorem 0.1.

For those non-collapsed nn-manifolds with |RicM|n1|\operatorname{Ric}_{M}|\leq n-1 and (δ,r)(\delta,r)-Reifenberg local covering geometry, the C1,αC^{1,\alpha}-precompactness of non-collapsed has been proved in [17, Theorem E].

It should be pointed out that, though a limit space XX in Theorems 0.1-0.3 admits the sythetic CD((n1),n)CD(-(n-1),n) curvature condition (cf. [30]) or Bakry-Émery-Ricci curvature lower bound in a generalized sense by [48], their weighted measures cannot be used to detect how much XX is collapsed as in Theorem 0.3. On the other hand, for the original mm-Hausdorff measure on XX, we do not know whether the volume comparison is satisfied; compare [55].

Remark 0.5.

It is well known by [54] that once XX admits a positive C1,αC^{1,\alpha}-harmonic radius r>0r>0, it can be smoothed to a new metric hϵh_{\epsilon}, whose sectional curvature |sechϵ|C(r,m,ϵ)|\operatorname{sec}_{h_{\epsilon}}|\leq C(r,m,\epsilon). The smoothed metric h(ϵ)h(\epsilon) in Theorem 0.1 has a better order (0.1.1.b-c), which arises from the Ricci flow solutions g(ϵ)g(\epsilon) on such manifolds, where h(ϵ)h(\epsilon) is their limit metric under Gromov-Hausdorff topology.

Indeed, by [40] (see also Lemma 1.12 below), for any X𝒳n,r,vm(δ,ρ),X\in\mathcal{X}_{n,r,v}^{m}(\delta,\rho), the manifolds in its converging sequence with bounded Ricci curvature and (r,v)(r,v)-local covering geometry satisfy (δ,r)(\delta,r^{\prime})-Reifenberg local covering geometry for some r=r(n,r,v,ρ)>0r^{\prime}=r^{\prime}(n,r,v,\rho)>0. Then by Dai-Wei-Ye [23] (see also Theorem 1.8 below), the solution g(t)g(t) of Ricci flow equation with initial value gg exists in (0,T(n,r)](0,T(n,r^{\prime})] for some constant T(n,r)>0T(n,r^{\prime})>0, and satisfies (0.1.1.a-c) on the local universal cover for 0<ϵ=tT(n,r)0<\epsilon=t\leq T(n,r^{\prime}). Hence any limit space in 𝒳n,r,vm(δ,r)\mathcal{X}_{n,r,v}^{m}(\delta,r^{\prime}) admits a nearby metric h(t)h(t) that locally is a quotient orbit space of manifolds with bounded sectional curvature C(n,r)t1/2C(n,r^{\prime})t^{-1/2}.

By elementary facts on Riemannian submersions (e.g., see the proof of Lemma 3.1 below), the curvature condition on g(t)g(t) naturally passes to the quotient metric h(t)h(t) in a harmonic coordinate chart (see the diagram (0.2) below). Note that, though the lower curvature bound can be always passed to h(t)h(t) by the O’Neill’s formula, the radius of harmonic coordinates on the quotient is crucial for the upper curvature bound. There are limit spaces, e.g. [26, §1-e, Example 1.13], whose sectional curvature blows up as the volume goes to zero.

We now give an application of Theorem 0.1. As a parametrized version of Gromov’s almost flat manifold theorem ([33], [59]), Fukaya [27] constructed a bundle structure whose fibers absorb all collapsing directions on a manifold MM that is Gromov-Haussdorff close to a lower dimensional manifold under bounded sectional curvature.

Theorem 0.6 (Fukaya’s fibration theorem [27, 12]).

Given constants n2n\geq 2, 1i0>01\geq i_{0}>0, there are constants ϵ(n)>0\epsilon(n)>0 and C(n)>0C(n)>0 such that the following holds.

Let (M,g)(M,g) and (N,h)(N,h) be a closed Riemannian nn-manifold and m(n)m(\leq n)-manifold respectively, whose sectional curvature and injectivity radius satisfy

|sec(M,g)|1,|sec(N,h)|1,inj.rad(N,h)i0.|\sec_{(M,g)}|\leq 1,\quad|\sec_{(N,h)}|\leq 1,\quad\operatorname{inj.rad}(N,h)\geq i_{0}.

If dGH(M,N)ϵi0d_{GH}(M,N)\leq\epsilon\cdot i_{0} with ϵ<ϵ(n)\epsilon<\epsilon(n), then there is a CC^{\infty}-smooth fibration f:MNf\mathrel{\mathop{\mathchar 58\relax}}M\to N such that

  1. (0.6.1)

    ff is a ϰ(ϵ|n)\varkappa(\epsilon\,|\,n)-almost Riemannian submersion, i.e., for any vector ξ\xi perpendicular to an ff-fiber, eϰ(ϵ|n)|ξ|g|df(ξ)|heϰ(ϵ|n)|ξ|g,e^{-\varkappa(\epsilon\,|\,n)}|\xi|_{g}\leq|df(\xi)|_{h}\leq e^{\varkappa(\epsilon\,|\,n)}|\xi|_{g}, where after fixing nn, ϰ(ϵ|n)0\varkappa(\epsilon\,|\,n)\to 0 as ϵ0\epsilon\to 0.

  2. (0.6.2)

    The intrinsic diameter of any ff-fiber Fx=f1(x)F_{x}=f^{-1}(x) over xNx\in N satisfies diamg(Fx)C(n)dGH(M,N).\operatorname{diam}_{g}(F_{x})\leq C(n)\cdot d_{GH}(M,N).

  3. (0.6.3)

    The second fundamental form is bounded by |2f|C(n)i01.\left|\nabla^{2}f\right|\leq C(n)i_{0}^{-1}.

  4. (0.6.4)

    FxF_{x} is diffeomorphic to an infra-nilmanifold.

Remark 0.7.

The formulation of Theorem 0.6 is similar to [12, Theorem 2.6], where the estimates are better than its original versions [27, 29], but depend on the higher regularities of gg and hh, called AA-regular in [12], i.e., the curvature tensor satisfies |iRm|Ai|\nabla^{i}\operatorname{Rm}|\leq A_{i} for all integer i0i\geq 0. It is well-known that the dependence of {Ai}i1\{A_{i}\}_{i\geq 1} in (0.6.1-3) can be removed in several ways, e.g. see [58] for a simple proof, and also Theorem 0.8 below.

The last main result in this paper is an optimal generalization of Fukaya’s fibration theorem on collapsed manifolds under bounded Ricci curvature. Let δ(n)>0\delta(n)>0 be the constant in Theorem 0.1.

Theorem 0.8.

Given ρ>0\rho>0 and positive integers n,m(n)n,m(\leq n), there exist constants ϵ(n),C(n)\epsilon(n),C(n) such that the following holds.

Let (M,g)(M,g) be a closed Riemannian nn-manifold with |RicM|n1|\operatorname{Ric}_{M}|\leq n-1 and (δ(n),ρ)(\delta(n),\rho)-Reifenberg local covering geometry, and (X,h)𝒴nm(δ(n),ρ)(X,h)\in\mathcal{Y}_{n}^{m}(\delta(n),\rho) for 0<ρ10<\rho\leq 1. If dGH(M,X)ϵρd_{GH}(M,X)\leq\epsilon\cdot\rho with ϵ<ϵ(n)\epsilon<\epsilon(n), then there is a CC^{\infty}-smooth fibration f:MXf\mathrel{\mathop{\mathchar 58\relax}}M\to X that satisfies (0.6.1)-(0.6.4) after replacing i0i_{0} with ρ\rho.

Theorem 0.8 also holds for closed Riemannian nn-manifolds with |RicM|n1|\operatorname{Ric}_{M}|\leq n-1 and (r,v)(r,v)-local covering geometry, because by Lemma 1.12 it admits a uniform Reifenberg-bounded local covering geometry.

The existence of a fibration that is a ϰ(ϵ|n)\varkappa(\epsilon\,|\,n)-Gromov-Hausdorff approximation satisfying (0.6.1) and (0.6.4) was already well known (cf. [50, Proposition 6.6]). In fact, after removing the upper Ricci curvature bound in Theorem 0.8, a smooth fibration that is ϰ(ϵ|n)\varkappa(\epsilon\,|\,n)-Gromov-Hausdorff approximation was constructed in both [40] (by smoothing methods based on Perelman’s pseudo-locality [52] for the Ricci flow) and [39] (by gluing δ\delta-splitting maps together via center of mass), where the uniform regularity is Hölder. Such fibrations are also constructed between Alexandrov spaces [25] recently.

What is new for the fibration in Theorem 0.8 is that, it provides the best possible regularity (0.6.3), which is first known in the literature even for the case that (X,h)(X,h) is an Euclidean 11-ball (cf. [50, Proposition 6.6]).

Remark 0.9.

The fibration in Theorem 0.8 is constructed via gluing the locally defined Cheeger-Colding’s δ\delta-splitting maps together. However, the optimal regularities are not direct consequences of neither those smoothing methods (e.g., [23] [54]), nor the Cheeger-Colding’s L2L^{2}-estimates [10, 11] on the δ\delta-splitting map. The subtle point is the balance between (0.6.1) and (0.6.3). For example, if a fiber bundle ftf_{t} is constructed with respect to a smoothed metric g(t)g(t) by the earlier known methods, then ft:(M,g)(X,h)f_{t}\mathrel{\mathop{\mathchar 58\relax}}(M,g)\to(X,h) is a ϰ(ϵ,t|n)\varkappa(\epsilon,t\,|\,n)-almost Riemannian submersion depending also on tt. In order to derive (0.6.1), tt has to approach 0, while secg(t)\operatorname{sec}_{g(t)} and hence 2ft\nabla^{2}f_{t} generally blows up as t0t\to 0. Similar issue also happens in applying Cheeger-Colding’s L2L^{2}-estimates. Instead, we apply the C2,αC^{2,\alpha}-compactness of harmonic coordinate charts on the local covers, which is crucial in deriving the optimal regularities for Theorem 0.8. For details, see Remarks 7.2 and 7.3 below.

Remark 0.10.

Compared with those earlier versions of the fibration Theorem in [29], [12], [50], [40], [39], etc., another improvement here is that, (X,h)(X,h) has only C1,αC^{1,\alpha}-regularity that may even admit no standard exponential map (see [37], cf. [7]). We will apply the center of mass technique with respect to a smoothed nearby metric h(t0)h(t_{0}) offered by Theorem 0.1, which admits a convex radius depends on t0t_{0}, such that (0.6.1)-(0.6.4) are proved for the original metric hh with fixed t0t_{0}.

At the core of Theorem 0.1 is the proof of the existence of harmonic coordinates, i.e., the charts for which the coordinate functions are harmonic functions on balls of a uniform size (depending only on the constants given), and uniform C1,αC^{1,\alpha}-norm estimates of the metric tensor in these coordinates [8, 45]. The main ingredients are as follows.

Let (Mi,gi)(M_{i},g_{i}) be a sequence of Riemannian nn-manifolds with |RicMi|n1|\operatorname{Ric}_{M_{i}}|\leq n-1 and (r,v)(r,v)-local covering geometry that converges to X𝒳n,r,vm(δ,ρ)X\in\mathcal{X}_{n,r,v}^{m}(\delta,\rho) in the Gromov-Hausdorff topology. By Lemma 1.12 and [4], for δ=δ(n)>0\delta=\delta(n)>0, the C1,αC^{1,\alpha}-harmonic radius of x~i\tilde{x}_{i} in the universal cover of Br(xi)B_{r}(x_{i}) admits a uniform lower bound r0(n,r,v,ρ)>0r_{0}(n,r,v,\rho)>0. For simplicity we assume r=ρr=\rho.

According to the precompactness principle for domains with boundary [63] (see Theorem 1.9 below), there is a normal cover U^i\widehat{U}_{i} of Bρ2(xi,gi)B_{\frac{\rho}{2}}(x_{i},g_{i}) equipped with its length metric such that by passing to a subsequence, it converges equivariantly in the pointed Gromov-Hausdorff topology:

(U^i,x^i,Γi)iGH(Y,x^,G)πiπ(Bρ2(xi,gi),xi)iGHY/G,\begin{CD}(\widehat{U}_{i},\hat{x}_{i},\Gamma_{i})@>{GH}>{i\to\infty}>(Y,\hat{x}_{\infty},G)\\ @V{\pi_{i}}V{}V@V{\pi_{\infty}}V{}V\\ (B_{\frac{\rho}{2}}(x_{i},g_{i}),x_{i})@>{GH}>{i\to\infty}>Y/G,\end{CD} (0.2)

where Γi\Gamma_{i} is the deck-transformation of Γ12,ρ(xi)\Gamma_{\frac{1}{2},\rho}(x_{i}), GG is the limit group of Γi\Gamma_{i}, and the quotient Y/GY/G is locally isometric to the limit ball Bρ2(x)B_{\frac{\rho}{2}}(x_{\infty}) on XX. Moreover, by the definition of U^i\widehat{U}_{i} (see Remark 1.11 below), it still admits a uniform C1,αC^{1,\alpha}-harmonic radius lower bound. Hence YY is a C1,αC^{1,\alpha}-Riemannian manifold.

In order to present the idea shortly, we first assume that YY is a smooth Riemannian manifold. Since the tangent cone of Y/GY/G is the quotient space of n\mathbb{R}^{n}, which is either isometric to m\mathbb{R}^{m} or definitely away from m\mathbb{R}^{m}, the (δ,ρ)(\delta,\rho)-Reifenberg condition implies that Y/GY/G is regular. By the standard theory of isometric actions on Riemannian manifolds (e.g. [35, §1]), Y/GY/G is also a Riemannian manifold. Then we construct a harmonic coordinate on XX in the following two steps.

Step 1. Following Cheeger-Colding [10, 11], we construct a harmonic δ\delta-splitting map φi:Bδ1/4ρ(xi,δ1gi)m\varphi_{i}\mathrel{\mathop{\mathchar 58\relax}}B_{\delta^{-1/4}\rho}(x_{i},\delta^{-1}g_{i})\to\mathbb{R}^{m}. We lift φi\varphi_{i} to a harmonic δ\delta-splitting map φ^i=φiπi\hat{\varphi}_{i}=\varphi_{i}\circ\pi_{i} on U^i\hat{U}_{i}, and then by appending other harmonic functions, we complete it to a harmonic coordinate chart (φ^i,ψ^i):Bδ1/4ρ(x^i,δ1g^i,U^i)n(\hat{\varphi}_{i},\hat{\psi}_{i})\mathrel{\mathop{\mathchar 58\relax}}B_{\delta^{-1/4}\rho}(\hat{x}_{i},\delta^{-1}\hat{g}_{i},\widehat{U}_{i})\to\mathbb{R}^{n}.

Step 2. By taking limit of (φ^i,ψ^i)(\hat{\varphi}_{i},\hat{\psi}_{i}), we get a harmonic coordinate chart (φ^,ψ^)(\hat{\varphi}_{\infty},\hat{\psi}_{\infty}) on YY such that φ^\hat{\varphi}_{\infty} takes the same value on each GG-orbit. Hence φ^\hat{\varphi}_{\infty} descends to a smooth map φ\varphi_{\infty} on XX. For simplicity such harmonic coordinate chart is called to be adapted for a submersion π\pi, i.e., each yjy^{j} (j=1,,m)(j=1,\dots,m) takes the same value along every π\pi-fiber.

By the C1,αC^{1,\alpha}-precompactness on U^i\widehat{U}_{i} via harmonic coordinates, (φ^,ψ^)(\hat{\varphi}_{\infty},\hat{\psi}_{\infty}) admits a small Hessian up to a definite rescaling on the metric. By the technical result below, φ\varphi_{\infty} gives rise to a harmonic coordinate chart of definite size on XX.

Theorem 0.11.

Given any r>0,0<α<1,0<Q102r>0,0<\alpha<1,0<Q\leq 10^{-2} and integers n,m(n)n,m(\leq n), there is a constant τ(n,r,α,Q)>0\tau(n,r,\alpha,Q)>0 such that the following holds.

Let π:(Y,g)(X,h)\pi\mathrel{\mathop{\mathchar 58\relax}}(Y,g)\to(X,h) be a Riemannian submersion from a Riemannian nn-manifold (may not complete) to a Riemannian mm-manifold. Suppose that there is an adapted C1,αC^{1,\alpha}-harmonic coordinate chart (y1,,yn):Br(p)n(y^{1},\dots,y^{n})\mathrel{\mathop{\mathchar 58\relax}}B_{r}(p)\to\mathbb{R}^{n} at pYp\in Y with (α,Q)(\alpha,Q)-C1,αC^{1,\alpha}-control (see (1.1.1)-(1.1.2) below). If the Hessian of each adapted coordinate function satisfies

HessyjC0,α(Br(p))τ(n,r,α,Q),j=1,,m,\|\operatorname{Hess}y^{j}\|_{C^{0,\alpha}(B_{r}(p))}\leq\tau(n,r,\alpha,Q),\ \ j=1,\dots,m, (0.3)

where the C0,αC^{0,\alpha}-norm is taken in the coordinates (y1,,yn)(y^{1},\dots,y^{n}), then there is a C1,αC^{1,\alpha}-harmonic coordinate chart (x1,,xm):Br/2(p¯)m(x^{1},\dots,x^{m})\mathrel{\mathop{\mathchar 58\relax}}B_{r/2}(\bar{p})\to\mathbb{R}^{m} at p¯=π(p)\bar{p}=\pi(p) with (α,2Q)(\alpha,2Q)-C1,αC^{1,\alpha}-control.

Note that the limit space YY of the normal covers of balls in (0.2) is only a C1,αC^{1,\alpha}-Riemannian manifold, and in general, a C1,αC^{1,\alpha}-Riemannian manifold may even admit no standard exponential map (see [37], cf. [7]). In order to guarantee the arguments above, we will show that π\pi_{\infty} is still a smooth submersion between C1,αC^{1,\alpha}-Riemannian manifolds. This can be seen by applying the Ricci flow on (Mi,gi)(M_{i},g_{i}) to obtain smooth limits YtY_{t} and XtX_{t}, which share the same limit group action in (0.2); see Proposition 4.1 below. Thus a harmonic coordinate chart on XX can be constructed as above.

The organization of this paper is as following. In section 1, we will supply some notations and preliminary facts that will be used later. In section 2 we give a simple proof of the C1,αC^{1,\alpha}-compactness for limit spaces under bounded sectional curvature. Section 3 is devoted to the proof of Theorem 0.11. In section 4, we shall prove that each element X𝒳n,r,vm(δ,ρ)X\in\mathcal{X}_{n,r,v}^{m}(\delta,\rho) is a smooth manifold with a C1,αC^{1,\alpha}-Riemannian metric. In section 5, we will construct C1,αC^{1,\alpha}-harmonic coordinate charts on XX, and complete the proof of Theorem 0.1. Theorem 0.3 and Theorem 0.8 will be proved in section 6 and 7 respectively.

Acknowledgement. The authors are grateful to Professor Xiaochun Rong for his interest and very helpful discussion on the results in this paper. Z. J. is supported by China Postdoctoral Science Foundation Grant No. 8206300494. S. X. is supported in part by Beijing Natural Science Foundation Grant No. Z19003 and National Natural Science Foundation of China Grant No. 11871349.

1. Preliminaries

In this section, we will supply some notations and basic results that will be used through the rest of the paper.

1.1. C1,αC^{1,\alpha}-Convergence

In this subsection, we will introduce the concepts such as the harmonic radius of a Riemannian manifold, C1,αC^{1,\alpha}-convergence of a sequence of Riemannian manifolds. After that, we will give the well-known Cheeger-Gromov’s and Anderson’s C1,αC^{1,\alpha}-convergence theorems [8], [34], [4], [5] (cf. also [53], [32], [47], [31], [38]).

Definition 1.1.

Given α(0,1)\alpha\in(0,1) and Q>0Q>0. Let (M,g)(M,g) be a smooth nn-manifold with a C1,αC^{1,\alpha}-Riemannian metric gg. For any qMq\in M, we define the C1,αC^{1,\alpha}-harmonic radius at qq as the largest number rh=rh(α,Q)(q,g)r_{h}=r_{h}(\alpha,Q)(q,g) such that on the geodesic ball Brh(q,g)B_{r_{h}}(q,g) of radius rhr_{h} centered at qq, there is a harmonic coordinate chart φ=(x1,,xn):Brh(q,g)Ωn\varphi=(x^{1},\dots,x^{n})\mathrel{\mathop{\mathchar 58\relax}}B_{r_{h}}(q,g)\to\Omega\subset\mathbb{R}^{n} such that the metric tensor admits the following (α,Q)(\alpha,Q)-C1,αC^{1,\alpha}-control:

  1. (1.1.1)

    eQδijgijeQδije^{-Q}\delta_{ij}\leq g_{ij}\leq e^{Q}\delta_{ij} as bilinear forms, where gij=g(xi,xj)g_{ij}=g\left(\frac{\partial}{\partial x^{i}},\frac{\partial}{\partial x^{j}}\right), for i,j=1,,ni,j=1,\dots,n, and δij\delta_{ij} are the Kronecker symbols,

  2. (1.1.2)

    rh1+αkgijC0,α(Ω)eQr_{h}^{1+\alpha}\left\|\partial_{k}g_{ij}\right\|_{C^{0,\alpha}(\Omega)}\leq e^{Q}, which means

    k=1nrhsupxΩ|kgij(x)|+k=1nrh1+αsupy,zΩ,yz|kgij(y)kgij(z)|dg(y,z)αeQ,\sum_{k=1}^{n}r_{h}\sup_{x\in\Omega}\left|\partial_{k}g_{ij}(x)\right|+\sum_{k=1}^{n}r_{h}^{1+\alpha}\sup_{y,z\in\Omega,y\neq z}\frac{\left|\partial_{k}g_{ij}(y)-\partial_{k}g_{ij}(z)\right|}{d_{g}(y,z)^{\alpha}}\leq e^{Q},

    holds for k=xk\partial_{k}=\frac{\partial}{\partial x^{k}} and the distance dgd_{g} associated with gg.

The harmonic radius rh(α,Q)(M,g)r_{h}(\alpha,Q)(M,g) of (M,g)(M,g) is now defined by rh(α,Q)(M,g)=infqMrh(q,g)r_{h}(\alpha,Q)(M,g)=\inf_{q\in M}r_{h}(q,g). For simplicity we will omit α\alpha and QQ when there is no confusion.

In general, the C1,αC^{1,\alpha}-norms of the components, gijg_{ij}, of metric gg in the coordinates {xi}i=1n\{x_{i}\}_{i=1}^{n} are defined on the Euclidean domain Ω\Omega. For convenience, we also denote the C1,αC^{1,\alpha}-norm of gijg_{ij} on Ω=φ(Brh(q,g))\Omega=\varphi(B_{r_{h}}(q,g)) by gijC1,α(Brh(q,g))\left\|g_{ij}\right\|_{C^{1,\alpha}(B_{r_{h}}(q,g))}.

Note that, g(xi,xj)g\left(\frac{\partial}{\partial x^{i}},\frac{\partial}{\partial x^{j}}\right) and kgij\partial_{k}g_{ij} in (1.1.1) and (1.1.2) can be replaced equivalently with g(xi,xj)g(\nabla x^{i},\nabla x^{j}) and xkg~ij\nabla x^{k}\tilde{g}_{ij} respectively, where xk\nabla x^{k} is the gradient of the coordinate function xkx^{k} with respect to metric gg.

Remark 1.2.

By definition, rh(q,gδ)=δ1rh(q,g)r_{h}(q,g_{\delta})=\delta^{-1}r_{h}(q,g) for gδ=δ2gg_{\delta}=\delta^{-2}g.

Now we give the concept of C1,αC^{1,\alpha}-convergence of a sequence of Riemannian manifolds.

Definition 1.3.

Let MM be a closed smooth nn-manifold. Let gig_{i} and gg be complete C1,αC^{1,\alpha}-smooth Riemannian metrics on MM. We say that gig_{i} converges to gg in the sense of C1,αC^{1,\alpha}-norm if for any pMp\in M, there exists a coordinate chart around pp, (x1,,xn):UΩn\left(x^{1},\dots,x^{n}\right)\mathrel{\mathop{\mathchar 58\relax}}U\to\Omega\subset\mathbb{R}^{n}, such that gi,st=gi(xs,xt)g_{i,st}=g_{i}\left(\frac{\partial}{\partial x^{s}},\frac{\partial}{\partial x^{t}}\right) C1,αC^{1,\alpha}-converges to gst=g(xs,xt)g_{st}=g\left(\frac{\partial}{\partial x^{s}},\frac{\partial}{\partial x^{t}}\right) as ii\to\infty. i.e.,

gi,stgstC1,α(Ω)0,asi,\left\|g_{i,st}-g_{st}\right\|_{C^{1,\alpha}(\Omega)}\to 0,\ \text{as}\ i\to\infty, (1.1)

where the C1,αC^{1,\alpha}-norm fC1,α(Ω)\left\|f\right\|_{C^{1,\alpha}(\Omega)} of a smooth function ff is defined by

fC1,α(Ω)=supxΩ|f(x)|+k=1nsupxΩ|kf(x)|+k=1nsupyzΩ|kf(y)kf(z)|dg(y,z)α\left\|f\right\|_{C^{1,\alpha}(\Omega)}=\sup_{x\in\Omega}\left|f(x)\right|+\sum_{k=1}^{n}\sup_{x\in\Omega}\left|\partial_{k}f(x)\right|+\sum_{k=1}^{n}\sup_{y\neq z\in\Omega}\frac{|\partial_{k}f(y)-\partial_{k}f(z)|}{d_{g}(y,z)^{\alpha}}

and k=xk\partial_{k}=\frac{\partial}{\partial x^{k}}.

In practice, an open cover of coordinate charts (xj1,,xjn):UjΩjn\left(x_{j}^{1},\dots,x_{j}^{n}\right)\mathrel{\mathop{\mathchar 58\relax}}U_{j}\to\Omega_{j}\subset\mathbb{R}^{n} are usually fixed as the background coordinate charts for the C1,αC^{1,\alpha}-convergence on MM.

Definition 1.4.

Let (Mj,gj)(M_{j},g_{j}) and (M,g)(M,g) be closed smooth nn-manifolds with C1,αC^{1,\alpha}-smooth Riemannian metrics. We say that (Mj,gj)(M_{j},g_{j}) converges to (M,g)(M,g) in the C1,αC^{1,\alpha}-topology if there exists an integer j0>0j_{0}>0 such that the following holds: for each jj0j\geq j_{0} there exists C2,αC^{2,\alpha}-diffeomorphism Φj:MMj\Phi_{j}\mathrel{\mathop{\mathchar 58\relax}}M\to M_{j} such that the pullback metric Φjgj\Phi_{j}^{*}g_{j} converges to gg in the sense of C1,αC^{1,\alpha}-norm.

Note that, the pullback of gjg_{j} by a diffeomorphism is crucial in Definition 1.4: even if (M,gj)(M,g_{j}) converges to (M,g)(M,g) on the same manifold MM in the sense of C1,αC^{1,\alpha}-topology, it does not mean that gjg_{j} converges to gg in the C1,αC^{1,\alpha}-norm. A counterexample can be found in [38, Remark 3 below Main theorem].

We say that a sequence of pointed complete Riemannian manifold (Mi,gi,pi)(M_{i},g_{i},p_{i}) converges in the C1,αC^{1,\alpha}-topology to a limit (M,g,p)(M,g,p) if

  1. (1.4.1)

    there exists an exhaustion of MM by open subsets {Ui}i=1\{U_{i}\}_{i=1}^{\infty} such that UiUi+1U_{i}\subseteq U_{i+1} and M=UiM=\bigcup U_{i};

  2. (1.4.2)

    there exists a sequence of C2,αC^{2,\alpha}-embeddings ϕi:UiMi\phi_{i}\mathrel{\mathop{\mathchar 58\relax}}U_{i}\to M_{i} such that

    ϕi(p)=pi,andϕigiC1,αg\phi_{i}(p)=p_{i},\ \text{and}\ \phi_{i}^{*}g_{i}\overset{C^{1,\alpha}}{\longrightarrow}g

    uniformly on any compact subset of MM.

Let us view (Ω,gi)(\Omega,g_{i}) as the domain in Definition 1.1 with the pullback metric by φi1\varphi_{i}^{-1}, where φi\varphi_{i} is a C1,αC^{1,\alpha}-harmonic coordinate chart φi:Brh(qi,gi)Ωn\varphi_{i}\mathrel{\mathop{\mathchar 58\relax}}B_{r_{h}}(q_{i},g_{i})\to\Omega\subset\mathbb{R}^{n}. Then the Cartesian coordinates on n\mathbb{R}^{n} (x1,,xn):(Ω,gi)n(x^{1},\dots,x^{n})\mathrel{\mathop{\mathchar 58\relax}}(\Omega,g_{i})\to\mathbb{R}^{n} is harmonic with respect to gig_{i}, which satisfies the (α,Q)(\alpha,Q)-C1,αC^{1,\alpha}-control (1.1.1)-(1.1.2). In the harmonic coordinates for a metric tensor gg, the Ricci curvature satisfies the following equation:

gij2grsxixj+B(gklxm,gkl)=2(Ricg)rs,g^{ij}\frac{\partial^{2}g_{rs}}{\partial x^{i}\partial x^{j}}+B(\frac{\partial g_{kl}}{\partial x^{m}},g_{kl})=-2(\operatorname{Ric}_{g})_{rs},

where BB is a quadratic term in gklxm\frac{\partial g_{kl}}{\partial x^{m}} for m=1,,nm=1,\dots,n (cf. [24]). By the standard LpL^{p}-estimate for elliptic PDEs, the L2,pL^{2,p}-norm of gklg_{kl} admits a uniform bound that depends on L1,pL^{1,p}-norm of gijg^{ij}, LpL^{p}-bound on the term BB and the LpL^{p}-bound on (Ricg)rs(\operatorname{Ric}_{g})_{rs} for any 1<p<+1<p<+\infty. Hence a subsequence of metric tensors gig_{i} converges to a limit C1,αC^{1,\alpha}-metric gg in the C1,αC^{1,\alpha}-norm if |Ricgi|n1,|\operatorname{Ric}_{g_{i}}|\leq n-1, where α=1np\alpha=1-\frac{n}{p} for any p>np>n. More generally, if gig_{i} converges to gg in the C1,αC^{1,\alpha}-norm with respect to another fixed coordinates on Ω\Omega, then similarly by the elliptic LpL^{p} regularity, the harmonic coordinates (xi1,,xin)(x^{1}_{i},\cdots,x^{n}_{i}) of gig_{i} admit a uniform L3,pL^{3,p}-bound, which implies that they converge to the harmonic coordinates of gg in the C2,αC^{2,\alpha}-norm.

Conversely, given a harmonic coordinate chart (x1,,xn):(Ω,g)n(x^{1},\dots,x^{n})\mathrel{\mathop{\mathchar 58\relax}}(\Omega,g)\to\mathbb{R}^{n} at pp with (α,Q)(\alpha,Q)-C1,αC^{1,\alpha}-control for gg, the Dirichlet problem associated with C1,αC^{1,\alpha}-nearby metric gig_{i} can be solved on (Ω,gi)(\Omega,g_{i}), with boundary value xik=xkx_{i}^{k}=x^{k} for each k=1,,nk=1,\dots,n. And the Schauder estimates give almost the same C1,αC^{1,\alpha}-control in the interior of Ω\Omega.

Therefore, the C1,αC^{1,\alpha}-harmonic radius under bounded Ricci curvature is continuous in the sense of C1,αC^{1,\alpha}-topology, i.e., the following proposition.

Proposition 1.5 ([4],[5]).

Let (Mi,gi)(M_{i},g_{i}) be a sequence of Riemannian manifolds with |Ric(Mi,gi)|n1|\operatorname{Ric}_{(M_{i},g_{i})}|\leq n-1, which C1,αC^{1,\alpha}-converges to a C1,αC^{1,\alpha}-Riemannian manifold (M,g)(M,g). Then

rh(M,g)=limirh(Mi,gi),r_{h}(M,g)=\lim_{i\to\infty}r_{h}(M_{i},g_{i}),

The same holds for rh(zi,gi)r_{h}(z_{i},g_{i}) and rh(z,g)r_{h}(z,g) as zi(Mi,gi)z_{i}\in(M_{i},g_{i}) converges to z(M,g)z\in(M,g).

Cheeger-Gromov-Anderson’s C1,αC^{1,\alpha}-convergence theorem says that the diffeomorphism types of the whole manifolds are also stable under the C1,αC^{1,\alpha}-topology, which provides a fundamental tool in this paper.

Theorem 1.6 ([4]).

The space |Ric|n(i0,D)\mathcal{M}_{|Ric|}^{n}(i_{0},D) of all closed Riemannian nn-manifolds (M,g)(M,g) such that

|Ric(M,g)|n1,inj.rad(M,g)i0>0,diam(M,g)D\left|\operatorname{Ric}_{(M,g)}\right|\leq n-1,\quad\operatorname{inj.rad}(M,g)\geq i_{0}>0,\quad\operatorname{diam}(M,g)\leq D (1.2)

is precompact in the C1,αC^{1,\alpha}-topology for any 0<α<10<\alpha<1. More precisely, any sequence of nn-manifolds {(Mi,gi)}|Ric|n(i0,D)\left\{(M_{i},g_{i})\right\}\subseteq\mathcal{M}_{|Ric|}^{n}(i_{0},D) admits a subsequence (Mi1,gi1)(M_{i_{1}},g_{i_{1}}) that converges to a closed smooth manifold (M,g)(M,g) with a C1,αC^{1,\alpha}-Riemannian metric gg via CC^{\infty}-smooth diffeomorphisms fi1:MMi1f_{i_{1}}\mathrel{\mathop{\mathchar 58\relax}}M\to M_{i_{1}} in the C1,αC^{1,\alpha}-topology.

In particular, there are only finitely many diffeomorphism types of nn-manifolds satisfying (1.2).

Remark 1.7.

Theorem 1.6 also holds for bounded domains in Riemannian manifolds, and for pointed complete but noncompact manifolds, after restricting to those compact subsets definitely away from the incomplete boundary (see [4, Main Lemma 2.2]).

1.2. Ricci Flows

Let (M,g)(M,g) be a closed Riemannian manifold. The Ricci flow was introduced by Hamilton [36] as the solution of the following degenerate parabolic PDE,

tg(t)=2Ricg(t),g(0)=g.\frac{\partial}{\partial t}g(t)=-2\operatorname{Ric}_{g(t)},\qquad g(0)=g. (1.3)

The solution always exists for a short time t>0t>0, and if it admits a finite maximal flow time Tmax<+T_{max}<+\infty, then the curvature tensor blows up as tTmaxt\to T_{max}, i.e., max|Rm(g(t))|g(t)+\max\left|\operatorname{Rm}(g(t))\right|_{g(t)}\to+\infty.

A basic property of Ricci flow is that it improves the regularity of the initial metric ([60, 61]), which depends on the flow time. The existence of a uniform definite flow time is important in practice.

Dai-Wei-Ye [23] proved that a uniform flow time T(n,r0)T(n,r_{0}) exists for a closed nn-manifold (M,g)(M,g) satisfying |Ric(M,g)|n1\left|\operatorname{Ric}_{(M,g)}\right|\leq n-1 and the conjugate radius conj.rad(M,g)r0>0\operatorname{conj.rad}(M,g)\geq r_{0}>0. As already pointed out by [17], the conjugate radius condition in their proof is only used to derive a uniform L2,pL^{2,p}-harmonic coordinates for the lifted metric on Br0(0)TxMB_{r_{0}}(0)\subseteq T_{x}M for all p1p\geq 1 (see [23, Remark 1]) and xMx\in M, which is required to apply the weak maximum principle [23, Theorem 2.1]. Since the same holds at a preimage point x~\tilde{x} on the universal covering space of a ρ\rho-ball Bρ(x)B_{\rho}(x) on (M,g)(M,g) when (M,g)(M,g) has (δ,ρ)(\delta,\rho)-local covering geometry, [23, Theorem 1.1] can be reformulated into the following form.

Theorem 1.8 ([23], cf. [17, Theorem 1.5]).

Given n,ρ>0n,\rho>0, there exist constants δ(n),T(n,ρ)>0\delta(n),T(n,\rho)>0 and C(n,ρ)>0C(n,\rho)>0 such that for any 0<δδ(n)0<\delta\leq\delta(n), if (M,g)(M,g) is a closed nn-manifold with |RicM|n1|\operatorname{Ric}_{M}|\leq n-1 and (δ,ρ)(\delta,\rho)-Reifenberg local covering geometry, then the Ricci flow equation (1.3) has a unique smooth solution g(t)g(t) for 0<tT(n,ρ)0<t\leq T(n,\rho) satisfying

{|g(t)g|g4t;|Rm(g(t))|g(t)C(n,ρ)t12;|kRm(g(t))|g(t)C(n,ρ,k,t);|Ric(M,g(t))|g(t)2(n1),\left\{\begin{array}[]{llll}&\left|g(t)-g\right|_{g}\leq 4t;\\ &\left|\operatorname{Rm}(g(t))\right|_{g(t)}\leq C(n,\rho)t^{-\frac{1}{2}};\\ &\left|\nabla^{k}\operatorname{Rm}(g(t))\right|_{g(t)}\leq C(n,\rho,k,t);\\ &\left|\operatorname{Ric}(M,g(t))\right|_{g(t)}\leq 2(n-1),\end{array}\right. (1.4)

where Rm(g(t))\operatorname{Rm}(g(t)) denotes the curvature tensor of g(t)g(t), kRm(g(t))\nabla^{k}\operatorname{Rm}(g(t)) the kthk^{th}-convariant derivative of Rm(g(t))\operatorname{Rm}(g(t)), whose norm is measured in g(t)g(t).

1.3. Gromov-Hausdorff precompactness for the covering spaces of open balls

Let (Mi,gi)(M_{i},g_{i}) be a sequence of complete Riemannian nn-manifolds with Ric(Mi,gi)(n1)\operatorname{Ric}_{(M_{i},g_{i})}\geq-(n-1). Let Br(xi,gi)B_{r}(x_{i},g_{i}) be an open rr-ball in (Mi,gi)(M_{i},g_{i}), and B~(xi,r)\widetilde{B}(x_{i},r) the Riemannian universal cover of Br(xi,gi)B_{r}(x_{i},g_{i}). Then B~(xi,r)\widetilde{B}(x_{i},r) may not admit a convergence subsequence in the pointed Gromov-Hausdorff topology (see [62, Example 3.2]).

Recently the third author [63] proved a precompactness principle for open domains in complete Riemannian manifold with Ric(n1)\operatorname{Ric}\geq-(n-1), which particularly is suitable for the covering spaces of open balls.

Let B^(x,r,R)\widehat{B}(x,r,R) be a component of the preimage π1(Br(x,g))\pi^{-1}(B_{r}(x,g)) in the Riemannian universal cover π:B~(x,R)BR(x,g)\pi\mathrel{\mathop{\mathchar 58\relax}}\widetilde{B}(x,R)\to B_{R}(x,g). Then B^(x,r,R)\widehat{B}(x,r,R) is a normal GG-cover of Br(x,g)B_{r}(x,g), where

G=Γr/R,R(x)=Image[π1(Br(x,g),x)π1(BR(x,g),x)],and\displaystyle G=\Gamma_{r/R,R}(x)=\operatorname{Image}[\pi_{1}(B_{r}(x,g),x)\to\pi_{1}(B_{R}(x,g),x)],\quad\text{and}
π1(B^(x,r,R))=Kernel[π1(Br(x,g),x)π1(BR(x,g),x)].\displaystyle\pi_{1}(\widehat{B}(x,r,R))=\operatorname{Kernel}[\pi_{1}(B_{r}(x,g),x)\to\pi_{1}(B_{R}(x,g),x)].

We endow B^(x,r,R)\widehat{B}(x,r,R) with a base point x^\hat{x} in the preimage of xx and its length metric induced from B~(x,R)\widetilde{B}(x,R).

Theorem 1.9 ([63]).

For any R>r>0R>r>0, let ^(r,R)\widehat{\mathcal{B}}(r,R) be the set consisting of the Riemannian normal covers B^(x,r,R)\widehat{B}(x,r,R) of all open balls in complete Riemannian nn-manifolds with Ric(n1)\operatorname{Ric}\geq-(n-1). Then ^(r,R)\widehat{\mathcal{B}}(r,R) is precompact in the pointed Gromov-Hausdorff topology.

More generally, let W(r)=jBr(pj,g)W(r)=\cap_{j}B_{r}(p_{j},g) be a non-empty intersection of open rr-balls Br(pj,g)B_{r}(p_{j},g) in (M,g)(M,g), and W~(r)\widetilde{W}(r) be the Riemannian universal cover of W(r)W(r). For R>r>0R>r>0, we define W^(r,R)\widehat{W}(r,R) to be a component of the preimage of W(r)W(r) in the Riemannian universal cover π:W~(R)W(R)=jBR(pj,g)\pi\mathrel{\mathop{\mathchar 58\relax}}\widetilde{W}(R)\to W(R)=\cap_{j}B_{R}(p_{j},g).

Theorem 1.10 ([63]).

For any R>r>0R>r>0, let 𝒲^(r,R)\widehat{\mathcal{W}}(r,R) be the set consisting of the normal covers W^(r,R)\widehat{W}(r,R) endowed with length metric of W(r)W(r) in complete Riemannian nn-manifolds with Ric(n1)\operatorname{Ric}\geq-(n-1). Then 𝒲^(r,R)\widehat{\mathcal{W}}(r,R) is precompact in the pointed Gromov-Hausdorff topology.

Note that for each jj, a component of πj1(W(R))\pi_{j}^{-1}(W(R)) in the universal cover πj:B~(pj,R)BR(pj,g)\pi_{j}\mathrel{\mathop{\mathchar 58\relax}}\widetilde{B}(p_{j},R)\to B_{R}(p_{j},g) is a normal cover of W(R)W(R), such that W~(R)\widetilde{W}(R) covers πj1(W(R))\pi_{j}^{-1}(W(R)). By the definition of W^(r,R)\widehat{W}(r,R), we derive that

W^(r,R)\widehat{W}(r,R) is a normal cover of πj1(W(r))B^(pj,r,R)\pi_{j}^{-1}(W(r))\subset\widehat{B}(p_{j},r,R). (1.5)

This fact will be applied in Section 4.

Remark 1.11.

Since B^(x,r/2,r)\widehat{B}(x,r/2,r) is a component of the preimage π1(Br/2(x,g))\pi^{-1}(B_{r/2}(x,g)) in the Riemannian universal cover π:B~(x,r)Br(x,g)\pi\mathrel{\mathop{\mathchar 58\relax}}\widetilde{B}(x,r)\to B_{r}(x,g), the (δ,r)(\delta,r)-Reifenberg local covering geometry condition is naturally passed to B^(x,r/2,r)\widehat{B}(x,r/2,r). That is, if for any xMx\in M, x~B~(x,r)\tilde{x}\in\widetilde{B}(x,r) is a (δ,r)(\delta,r)-Reifenberg point, then x^B^(x,r/2,r)\hat{x}\in\widehat{B}(x,r/2,r) is (δ,r/4)(\delta,r/4)-Reifenberg.

Finally we recall that, by [40] the local covering geometry via rewinding volume is equivalent to that by the Reifenberg condition for a manifold that is locally close to be Euclidean.

Lemma 1.12 ([40, Lemma 2.1]).

Given positive integer nn and real numbers δ,v>0\delta,v>0, there are constants ϵ(n),ρ(n,v,δ)>0\epsilon(n),\rho(n,v,\delta)>0 such that the following holds.

Let (M,g)(M,g) be a Riemannian nn-manifold with RicM(n1)\operatorname{Ric}_{M}\geq-(n-1) and

Vol(B1(x~))v>0,\operatorname{Vol}(B_{1}(\tilde{x}))\geq v>0,

where x~\tilde{x} is a preimage point of xx in the universal cover of B1(x)B_{1}(x). If

dGH(B1(x),B1m(0))ϵ(n),d_{GH}(B_{1}(x),B_{1}^{m}(0))\leq\epsilon(n),

then x~\tilde{x} is a (δ,ρ(n,v,δ))(\delta,\rho(n,v,\delta))-Reifenberg point, i.e.,

dGH(Br(x~),Brn(0))δr, 0<rρ(n,v,δ).d_{GH}(B_{r}(\tilde{x}),B_{r}^{n}(0))\leq\delta r,\qquad\forall\;0<r\leq\rho(n,v,\delta).
Proof.

Lemma 1.12 essentially is a restatement of [40, Lemma 2.1]. For the reader’s convenience we give a simple proof from a different viewpoint.

Let us argue by contradiction. Assume that there is a sequence (Mi,gi)(M_{i},g_{i}) with RicMi(n1)\operatorname{Ric}_{M_{i}}\geq-(n-1), Vol(B1(x~i))v>0\operatorname{Vol}(B_{1}(\tilde{x}_{i}))\geq v>0, dGH(B1(x),B1m(0))ϵi0d_{GH}(B_{1}(x),B_{1}^{m}(0))\leq\epsilon_{i}\to 0, but x~i\tilde{x}_{i} is not a (δ,ρ)(\delta,\rho)-Reifenberg point for any fixed δ,ρ>0\delta,\rho>0 for all sufficient large ii.

By passing to a subseqence, let us consider the equivariant pointed Gromov-Hausdorff convergence as in (0.2)

(B^(xi,1/2,1),x^i,Γi)GH(Y,x^,G)πiπ(B1/2(xi,gi),xi)GH(B1/2m(0),0)m.\begin{CD}(\widehat{B}(x_{i},1/2,1),\hat{x}_{i},\Gamma_{i})@>{GH}>{}>(Y,\hat{x},G)\\ @V{\pi_{i}}V{}V@V{}V{\pi}V\\ (B_{1/2}(x_{i},g_{i}),x_{i})@>{GH}>{}>(B_{1/2}^{m}(0),0)\subset\mathbb{R}^{m}.\end{CD}

Since Y/G=B1/2m(0)Y/G=B_{1/2}^{m}(0), a standard blowing-up argument implies that the quotient of any tangent cone Tx^T_{\hat{x}} modular the infinitesimal actions dGdG induced by GG is m\mathbb{R}^{m}, where the lines on m\mathbb{R}^{m} can be lifted onto Tx^T_{\hat{x}}. At the same time, by the fact that YY is a Ricci-limit space of a non-collapsing sequence, Tx^T_{\hat{x}} is a metric cone. Hence Tx^T_{\hat{x}} splits to m×C(Σ)\mathbb{R}^{m}\times C(\Sigma), and dGdG acts transitively on C(Σ)C(\Sigma). Hence Tx^T_{\hat{x}} is n\mathbb{R}^{n}.

In particular, there is ρ=ρ(δ,x^)>0\rho=\rho(\delta,\hat{x})>0 such that x^\hat{x} is a (δ/10,ρ)(\delta/10,\rho)-Reifenberg point. Then by Colding’s volume convergence ([21]), x~i\tilde{x}_{i} is (δ,ρ)(\delta,\rho)-Reifenberg point for ii large, a contradiction. ∎

2. C1,αC^{1,\alpha}-regularity of limit spaces under bounded sectional curvature

As the starting point of this paper, we give a simple proof of Theorem 0.3 for the limit spaces of closed manifolds with bounded sectional curvature.

In fact, it is essentially a corollary of [28, Theorem 0.9, Corollary 0.11], which now can be improved as follows.

Theorem 2.1.

Let (Mi,gi)(M_{i},g_{i}) be a sequence of closed Riemannian nn-manifolds with |secgi|1|\operatorname{sec}_{g_{i}}|\leq 1 such that (Mi,gi)GHX(M_{i},g_{i})\overset{GH}{\longrightarrow}X. Then XX admits a stratification X=S0(X)S1(X)Sk(X)X=S_{0}(X)\supset S_{1}(X)\supset\cdots\supset S_{k}(X) for 0kn0\leq k\leq n (some of them may be the same) such that

  1. (2.1.1)

    the Hausdorff dimension of XX is equal to kk.

  2. (2.1.2)

    Si(X)Si+1(X)=(Rki(X),hki)S_{i}(X)\setminus S_{i+1}(X)=(R^{k-i}(X),h_{k-i}) is a smooth (ki)(k-i)-manifold with a C1,αC^{1,\alpha}-Riemannian metric.

    Moreover, if the kk-Hausdorff measure Hk(X)v>0H^{k}(X)\geq v>0 and the diameter of XX D\leq D, then for any point pSi(X)Si+1(X)p\in S_{i}(X)\setminus S_{i+1}(X) that is ϵ\epsilon-away from Si+1(X)S_{i+1}(X), the C1,αC^{1,\alpha}-harmonic radius at pp is no less than rh(ϵ|n,r,α,Q,v,D)>0r_{h}(\epsilon\,|\,n,r,\alpha,Q,v,D)>0, where Q>0Q>0, 0<α<10<\alpha<1.

Note that the original proof of [28, Theorem 0.9] depends on orthonormal frame bundles of (Mi,gi)(M_{i},g_{i}), which do not generally admit a uniform sectional curvature bound such that only a C0,αC^{0,\alpha}-regularity can be derived on XX. In the following we will give a different and simple proof from the view point of local normal covers.

The key point behind is the following observation [28, Lemma 7.2] by Fukaya.

Lemma 2.2 (cf. [28, Lemma 7.2]).

Let (M,g)(M,g) be a Riemannian manifold whose sectional curvature bsecgab\leq\operatorname{sec}_{g}\leq a. Assume that there is a proper and free isometric action by GG on (M,g)(M,g). Let

(r/t)p(G)=sup{rp(g)/d(p,g(p))|geG,rp(g) is well-defined},(r/t)^{\perp}_{p}(G)=\sup\left\{\left.\|r^{\perp}_{p}(g)\|/d(p,g(p))\,\right|\,g\neq e\in G,\;\|r_{p}^{\perp}(g)\|\text{ is well-defined}\right\}, (2.1)

where rp(g)=sup{(dg(v),P(v))|vTpG(p)}\|r^{\perp}_{p}(g)\|=\sup\{\measuredangle(dg(v),P(v))\,|\,v\in T_{p}^{\perp}G(p)\} is defined when d(p,g(p))d(p,g(p)) is less than the injectivity radius at pp, and PP is the parallel transport from TpMT_{p}M to Tg(p)MT_{g(p)}M along the unique minimal geodesic.

Then the sectional curvature of the quotient M/GM/G at π(p)\pi(p) is bounded by

bsecπ(p)a+6((r/t)p(G))2.b\leq\operatorname{sec}_{\pi(p)}\leq a+6((r/t)^{\perp}_{p}(G))^{2}.
Proof.

Here the only difference from original [28, Lemma 7.2] is the rate of infinitesimal angle deviation (2.1). In [28, Lemma 7.2] the same conclusion was proved for

(r/t)p(G)=sup{rp(g)d(p,g(p))|geG,rp(g) is well-defined},(r/t)_{p}(G)=\sup\left\{\left.\frac{\|r_{p}(g)\|}{d(p,g(p))}\,\right|\,g\neq e\in G,\|r_{p}(g)\|\text{ is well-defined}\right\},

where rp(g)=sup{(dg(v),P(v))|vTpM}\|r_{p}(g)\|=\sup\{\measuredangle(dg(v),P(v))\,|\,v\in T_{p}M\} is defined when d(p,g(p))d(p,g(p)) is less than the injectivity radius at pp.

Note that by the tubular neighborhood theorem (e.g., see [35]), only the part of dgdg on the normal space TpG(p)T_{p}^{\perp}G(p) to the orbit G(p)G(p) makes effect on the horizontal directions. Lemma 2.2 follows the proof of [28, Lemma 7.2] line-by-line. ∎

Proof of Theorem 2.1.

Let us consider the smoothed metrics gi(t)g_{i}(t) on (Mi,gi)(M_{i},g_{i}), which are the solutions of Hamilton’s Ricci flow equation with the initial condition gi(0)=gig_{i}(0)=g_{i}. By [60] (cf. [56]), gi(t)g_{i}(t) also admits a uniform sectional curvature bound 1+C(n)t1+C(n)t and a uniform higher regularities (1.4).

By passing to a subsequence, we assume that (Mi,gi(t))GHXt(M_{i},g_{i}(t))\overset{GH}{\longrightarrow}X_{t} as ii\to\infty for any fixed t>0t>0. Then XtX_{t} is e2te^{2t}-bi-Lipschitz to XX, and admits a stratification S0(Xt)S1(Xt)Sk(Xt)S_{0}(X_{t})\supset S_{1}(X_{t})\supset\cdots\supset S_{k}(X_{t}) for 0kn0\leq k\leq n such that each strata Rki=Si(Xt)Si+1(Xt)R^{k-i}=S_{i}(X_{t})\setminus S_{i+1}(X_{t}) is a smooth Riemannian (ki)(k-i)-manifold with sectional curvature 1+ϰ(t|n)\geq 1+\varkappa(t\,|\,n).

Indeed, for xi(Mi,gi(t))x_{i}\in(M_{i},g_{i}(t)) that approaches xXtx\in X_{t}, let us consider the equivariant limit spaces (Yt,g(t),Gt)(Y_{t},g^{*}(t),G_{t}) of the π2\frac{\pi}{2}-ball in the tangent space TxiMiT_{x_{i}}M_{i} of xi(Mi,gi(t))x_{i}\in(M_{i},g_{i}(t)) with the pullback metric gi(t)g_{i}^{*}(t) via the exponential map:

(Bπ2(0i),gi(t),0i,Γi)iGH(Yt,g(t),yt,Gt)expπt(Bπ2(xi,gi(t)),xi)iGHBπ2(x,Xt)=Yt/Gt,\begin{CD}(B_{\frac{\pi}{2}}(0_{i}),g_{i}^{*}(t),0_{i},\Gamma_{i})@>{GH}>{i\to\infty}>(Y_{t},g^{*}(t),y_{t},G_{t})\\ @V{\exp}V{}V@V{\pi_{t}}V{}V\\ (B_{\frac{\pi}{2}}(x_{i},g_{i}(t)),x_{i})@>{GH}>{i\to\infty}>B_{\frac{\pi}{2}}(x,X_{t})=Y_{t}/G_{t},\end{CD} (2.2)

where Γi\Gamma_{i} is the pseudo-group action by the local fundamental group of Bπ2(xi,gi(t))B_{\frac{\pi}{2}}(x_{i},g_{i}(t)), and Bπ2(x,Xt)B_{\frac{\pi}{2}}(x,X_{t}) is equipped with its length metric. Then YtY_{t} is a smooth Riemannian manifold with sectional curvature |secYt|1+C(n)t|\operatorname{sec}_{Y_{t}}|\leq 1+C(n)t.

Since the pseudo-group GtG_{t} acts on YtY_{t} by isometries, by the standard theory of isometric actions on Riemannian manifolds (e.g., [3]) the orbit space Yt/GtY_{t}/G_{t} admits a standard stratification by isotropy types. So is XtX_{t}. Hence we derive (2.1.1).

For (2.1.2), it suffices to show that the sectional curvature at points in Si(Xt)S_{i}(X_{t}) ϵ\epsilon-definitely away from Si+1(Xt)S_{i+1}(X_{t}) is bounded uniformly by a constant C(n,v,D)C(n,v,D). Then (2.1.2) for the original limit space XX immediate follows from Cheeger-Gromov’s convergence theorem applied on XtX_{t} as t0t\to 0.

Let us argue by contradiction. Suppose there is a sequence of limit spaces (Xj,t,xj)(X_{j,t},x_{j}) such that xjRki(Xj,t)x_{j}\in R^{k-i}(X_{j,t}), d(xj,Si+1(Xj,t))ϵd(x_{j},S_{i+1}(X_{j,t}))\geq\epsilon, and t=tj(0,T(n)]t=t_{j}\in(0,T(n)] is arbitrary chosen, but the sectional curvature of Rki(Xj,t)R^{k-i}(X_{j,t}) at xjx_{j} is unbounded as jj\to\infty. Since XtX_{t} is an Alexandrov space with curvature C(n)\geq C(n),

maxv,wTxjsecXj,t(vw) as j+.\max_{v,w\in T_{x_{j}}}\operatorname{sec}_{X_{j,t}}(v\wedge w)\to\infty\text{ as $j\to+\infty$}.

By passing to a subsequence, let us consider the equivariant convergence of limit spaces (Yj,t,yj)(Y_{j,t},y_{j}) in (2.2) for each Xj,tX_{j,t}.

(Yj,t,gj(t),yj,Gj,t)jC1,α(Y,g,y,G)πj,tπ(Bπ2(xj,Xj,t),xj)jGH(Y/G,x).\begin{CD}(Y_{j,t},g_{j}^{*}(t),y_{j},G_{j,t})@>{C^{1,\alpha}}>{j\to\infty}>(Y,g^{*},y,G)\\ @V{\pi_{j,t}}V{}V@V{\pi_{\infty}}V{}V\\ (B_{\frac{\pi}{2}}(x_{j},X_{j,t}),x_{j})@>{GH}>{j\to\infty}>(Y/G,x).\end{CD} (2.3)

In order to apply Lemma 2.2, we first point out that the Lie group Gj,tG_{j,t} can be reduced to the case of free action.

Indeed, let Gj,0G_{j,0} be the identity component of Gj,tG_{j,t}. Instead of Gj,tG_{j,t}, we consider the actions by Gj,0G_{j,0}, which by the Heintze-Margulis lemma (e.g., [28, Lemma 4.1], cf. [33], [6]) is a niloptent Lie group. Moreover, by [28, Lemma 5.1] the isotropy group Gj,0,yjG_{j,0,y_{j}} of Gj,0G_{j,0} lies in the center of Gj,0G_{j,0}. Hence the isotropy group of Gj,0G_{j,0} in a conjugacy class is unique on Yj,tY_{j,t}. It follows that the union of all orbits of the same isotropy type Gj,0,yjG_{j,0,y_{j}}, πj,t1(Rki(Yj,t/Gj,t))\pi_{j,t}^{-1}(R^{k-i}(Y_{j,t}/G_{j,t})), is the fixed-point set of Gj,0,yjG_{j,0,y_{j}}, and hence is totally geodesic. Moreover, the isometric action Gj,tG_{j,t} on πj,t1(Rki(Yj,t/Gj,t))\pi_{j,t}^{-1}(R^{k-i}(Y_{j,t}/G_{j,t})) can be reduced to the quotient group Gj,0/Gj,0,yjG_{j,0}/G_{j,0,y_{j}} that acts effectively and freely.

By our assumption, Yj,t/Gj,0Y_{j,t}/G_{j,0} is not collapsing. For simplicity we assume that the action of Gj,0G_{j,0} itself is free.

We claim that

  1. (2.2.1)

    there is a sequence γjGj,0\gamma_{j}\in G_{j,0} such that ryj(γj)d(yj,γj(yj))\frac{\|r^{\perp}_{y_{j}}(\gamma_{j})\|}{d(y_{j},\gamma_{j}(y_{j}))}\to\infty. By a suitable choice of njn_{j}\to\infty, d(γjnj(yj),yj)0d(\gamma_{j}^{n_{j}}(y_{j}),y_{j})\to 0 and dγjnj|yj\left.d\gamma_{j}^{n_{j}}\right|_{y_{j}} maps the normal space of Gj,0(yj)G_{j,0}(y_{j}) by a uniform and definite deviation away from the parallel transformation.

  2. (2.2.2)

    the normal space of G0(y)G_{0}(y) has the same dimension as that of Gj,0(yj)G_{j,0}(y_{j}).

  3. (2.2.3)

    the limit η\eta of γjnj\gamma_{j}^{n_{j}} lies in the isotropy group G0,yG_{0,y} at yy, whose differential admits a non-trivial transformation on the normal space of G0(y)G_{0}(y).

The claims above will yield a contradiction immediately. Indeed, let ηs=limγj[snj]\eta_{s}=\lim\gamma_{j}^{[sn_{j}]} for each s[0,1]s\in[0,1] and γj\gamma_{j} in (2.2.1). Then ηs\eta_{s} forms a continuous path in G0,yG_{0,y} which by (2.2.3) acts on TyG0(y)T_{y}^{\perp}G_{0}(y) non-trivially. By (2.2.2) and the slice theorem, Y/G0Y/G_{0} has lower dimension than Yj,t/Gj,0Y_{j,t}/G_{j,0}, a contradiction to that they are assumed to be non-collapsing.

The verification of (2.2.1)-(2.2.3):

By Lemma 2.2, the unboundedness on the upper curvature at xjXj,tx_{j}\in X_{j,t} implies that (r/t)yj(Gj,0)(r/t)_{y_{j}}^{\perp}(G_{j,0}) blows up as jj\to\infty. Hence (2.2.1) holds by the definition (2.1) of (r/t)p(G)(r/t)_{p}^{\perp}(G).

Since (Yj,t,gj(t))(Y_{j,t},g_{j}^{*}(t)) converges to (Y,g)(Y,g^{*}) in the C1,αC^{1,\alpha}-topology, let us identify Bπ4(yj,gj(t))B_{\frac{\pi}{4}}(y_{j},g_{j}^{*}(t)) with Bπ4(y,g)B_{\frac{\pi}{4}}(y,g^{*}) via a suitable diffeomorphism. Then the isometric actions by one-parameter subgroups of Gj,0G_{j,0} C1C^{1}-converges to that of G0G_{0} on Bπ4(y,g(t))B_{\frac{\pi}{4}}(y,g^{*}(t)), which implies dimG0(y)dimGj,0(yj)\dim G_{0}(y)\geq\dim G_{j,0}(y_{j}).

Because Yj,t/Gj,0Y_{j,t}/G_{j,0} is assumped to be non-collapsing, the orbit Gj,0(yj)G_{j,0}(y_{j}) and G0(y)G_{0}(y) must have same dimension, which implies (2.2.2). Moreover, the normal space of Gj,0(yj)G_{j,0}(y_{j}) converges to that of G0(y)G_{0}(y), and the limit of γjnj\gamma_{j}^{n_{j}} in (2.2.1) is a nontrivial element η\eta in G0,yG_{0,y} with a non-trivial deviation on the normal space of G0(y)G_{0}(y), i.e., (2.2.3). ∎

We point out that, though the manifolds (Mi,gi)(M_{i},g_{i}) with |Ric(Mi,gi)|n1|\operatorname{Ric}_{(M_{i},g_{i})}|\leq n-1 and (δ,r)(\delta,r)-Reifenberg local covering geometry still can be smoothed to gi(t)g_{i}(t) via Ricci flow by Theorem 1.8, the proof of Theorem 2.1 fails to work for their limit spaces, due to that the sectional curvature of gi(t)g_{i}(t) generally blows up as t0t\to 0. Therefore, instead of Cheeger-Gromov’s convergence theorem, we have to construct a harmonic coordinates directly in the next two sections.

3. Construction of the C1,αC^{1,\alpha}-harmonic coordinate chart in quotient spaces

In this section we prove the technical theorem 0.11.

We first prove an adapted harmonic coordinate chart descends to a chart on the base manifold that is almost harmonic.

Lemma 3.1.

Let the assumptions be as in Theorem 0.11. Assume that

HessyjC0,α(Br(p))τ,j=1,,m.\|\operatorname{Hess}y^{j}\|_{C^{0,\alpha}(B_{r}(p))}\leq\tau,\qquad j=1,\dots,m. (3.1)

Then the adapted coordinate chart (y1,,yn):Br(p)n(y^{1},\dots,y^{n})\mathrel{\mathop{\mathchar 58\relax}}B_{r}(p)\to\mathbb{R}^{n} at pp descends to a coordinate chart (z1,,zm):B2r/3(p¯)m(z^{1},\dots,z^{m})\mathrel{\mathop{\mathchar 58\relax}}B_{2r/3}(\bar{p})\to\mathbb{R}^{m} at p¯=π(p)\bar{p}=\pi(p) such that yj=zjπy^{j}=z^{j}\circ\pi, and the metric tensor hh on B2r/3(p¯)XB_{2r/3}(\bar{p})\subset X expressed in hst=h(¯zs,¯zt)h_{st}=h(\bar{\nabla}z^{s},\bar{\nabla}z^{t}) satisfy

  1. (3.1.1)

    eQδsthsteQδste^{-Q}\delta_{st}\leq h_{st}\leq e^{Q}\delta_{st},

  2. (3.1.2)

    (2r/3)1+α¯zjhstC0,α(B2r/3(p¯))eQ(2r/3)^{1+\alpha}\left\|\bar{\nabla}z^{j}h_{st}\right\|_{C^{0,\alpha}\left(B_{2r/3}(\bar{p})\right)}\leq e^{Q}, and

  3. (3.1.3)

    zjz^{j} (j=1,,mj=1,\dots,m) is almost harmonic in the sense that

    HesshzjC0,α(B2r/3(p¯))Ψ(τ|n,α,Q),\left\|\operatorname{Hess}_{h}z^{j}\right\|_{C^{0,\alpha}\left(B_{2r/3}(\bar{p})\right)}\leq\Psi(\tau\,|\,n,\alpha,Q),

where ¯zj\bar{\nabla}z^{j} is the gradient of zjz^{j} with respect to hh, the C0,αC^{0,\alpha}-norm is taken in the coordinate chart (z1,,zm)(z^{1},\dots,z^{m}), and Ψ(τ|n,α,Q)\Psi(\tau\,|\,n,\alpha,Q) is a function depending on τ,n,α,Q\tau,n,\alpha,Q such that Ψ(τ|n,α,Q)0\Psi(\tau\,|\,n,\alpha,Q)\to 0 as τ0\tau\to 0 with fixed n,α,Qn,\alpha,Q.

Proof.

Because yjy^{j} (j=1,,mj=1,\dots,m) takes the same value on each fiber, zj=yjπ1z^{j}=y^{j}\circ\pi^{-1} is well-defined.

Let gst=g(ys,yt)g_{st}=g(\nabla y^{s},\nabla y^{t}), where \nabla is the Levi-Civita connection on (Y,g)(Y,g). From the definition of C1,αC^{1,\alpha}-harmonic coordinate chart, one has

{eQδstgsteQδst;r1+αjgstC0,α(Br(p))eQ,\begin{cases}e^{-Q}\delta_{st}\leq g_{st}\leq e^{Q}\delta_{st};\\ r^{1+\alpha}\|\partial_{j}g_{st}\|_{C^{0,\alpha}(B_{r}(p))}\leq e^{Q},\end{cases} (3.2)

where j=yj\partial_{j}=\nabla y^{j} and the C0,αC^{0,\alpha}-norm is taken in the coordinates {yj}\{y^{j}\}.

Since π\pi is a Riemannian submersion, the gradient yj\nabla y^{j} is a horizontal vector field on YY and

π(yj)=¯zj,j=1,,m,\pi_{*}(\nabla y^{j})=\bar{\nabla}z^{j},\quad j=1,\dots,m, (3.3)

where ¯\bar{\nabla} denotes the Levi-Civita connection on (X,h)(X,h), and π\pi_{*} is the tangent map of π\pi. Define hst=h(¯zs,¯zt)h_{st}=h(\bar{\nabla}z^{s},\bar{\nabla}z^{t}). Then

hst=gst,s,t=1,,m.h_{st}=g_{st},\quad s,t=1,\dots,m. (3.4)

It follows from (3.2) and (3.4) that hsth_{st} satisfies (3.1.1), and thus (z1,,zm):B2r/3(p¯)m(z^{1},\dots,z^{m})\mathrel{\mathop{\mathchar 58\relax}}B_{2r/3}(\bar{p})\to\mathbb{R}^{m} is a coordinate chart at p¯\bar{p}.

What remains is to verify the estimates (3.1.2) and (3.1.3). First, by (3.3) and (3.4), we derive that for each j=1,,m,j=1,\dots,m,

¯zj(hst)(q¯)\displaystyle\bar{\nabla}z^{j}(h_{st})(\bar{q}) =π(yj)(h(¯zs,¯zt))(π(q))=yj(g(ys,yt))(q)\displaystyle=\pi_{*}(\nabla y^{j})(h(\bar{\nabla}z^{s},\bar{\nabla}z^{t}))(\pi(q))=\nabla y^{j}(g(\nabla y^{s},\nabla y^{t}))(q) (3.5)
=Hess ys(yj,yt)(q)+Hess yt(ys,yj)(q)\displaystyle=\text{Hess }y^{s}(\nabla y^{j},\nabla y^{t})(q)+\text{Hess }y^{t}(\nabla y^{s},\nabla y^{j})(q)

for any q¯B2r/3(p¯)\bar{q}\in B_{2r/3}(\bar{p}) and qπ1(q¯)q\in\pi^{-1}(\bar{q}). Then (3.1) together with (3.5) yields (3.1.2).

Secondly, by (3.3), we have

ykyl=¯¯zk¯zl~+A(yk,yl),for any k,l=1,,m\nabla_{\nabla y^{k}}\nabla y^{l}=\widetilde{\bar{\nabla}_{\bar{\nabla}z^{k}}\bar{\nabla}z^{l}}+A(\nabla y^{k},\nabla y^{l}),\qquad\text{for any $k,l=1,\dots,m$} (3.6)

where A(,)=[,]A(\cdot,\cdot)=[\cdot,\cdot]^{\top} denotes the horizontal integral tensor of Riemannian submersion π\pi that takes values tangent to the fibers, and for any smooth vector field ZZ on XX, Z~\tilde{Z} denotes its horizontal lifting on YY. Combing (3.3) with (3.6) and by the fact that π\pi is Riemannian submersion, we have

Hesszj(¯zk,¯zl)=Hessyj(yk,yl)for any j,k,l=1,,m.\operatorname{Hess}z^{j}(\bar{\nabla}z^{k},\bar{\nabla}z^{l})=\operatorname{Hess}y^{j}(\nabla y^{k},\nabla y^{l})\qquad\text{for any $j,k,l=1,\dots,m$.} (3.7)

Now (3.7) together with (3.1) yields (3.1.3). This complete the proof of Lemma 3.1. ∎

Remark 3.2.

In the proof of Lemma 3.1, it follows from (3.4)-(3.5) that the curvature tensor of (X,h)(X,h) and its covariant derivatives of any order satisfy the same regularity as (Y,g)(Y,g) up to a definite ratio depending on nn and QQ. We will apply the fact in proving (0.1.1.a-c) in Theorem 0.1.

Next, let us prove Theorem 0.11 by solving the Dirichlet problem with the boundary condition zjz^{j}.

Proof of Theorem 0.11.

Let φ:=(z1,,zm):B2r/3(p¯)m\varphi\mathrel{\mathop{\mathchar 58\relax}}=(z^{1},\dots,z^{m})\mathrel{\mathop{\mathchar 58\relax}}B_{2r/3}(\bar{p})\to\mathbb{R}^{m} be a coordinate chart at p¯\bar{p} provided by Lemma 3.1. Let 0m0\in\mathbb{R}^{m} be the origin. By a shift in value, we assume that φ(p¯)=0\varphi(\bar{p})=0. Let Ωm\Omega\subseteq\mathbb{R}^{m} denote the image set of φ\varphi and φ1:Ω(B2r/3(p¯),h)\varphi^{-1}\mathrel{\mathop{\mathchar 58\relax}}\Omega\to(B_{2r/3}(\bar{p}),h) the inverse map of φ\varphi. Let us pullback the metric hh on XX to Ω\Omega by φ1\varphi^{-1}, where we still denote (φ1)h(\varphi^{-1})^{*}h on Ω\Omega by hh, and identify (B2r/3(p¯),h)(B_{2r/3}(\bar{p}),h) with (Ω,h)(\Omega,h) by φ1\varphi^{-1}.

From (3.1.1) and by 0<Q1020<Q\leq 10^{-2}, we have the Euclidean ball B3r/5m(0)ΩB_{3r/5}^{m}(0)\subseteq\Omega. Let xjx^{j} be the solution of the following Dirichlet problem:

{Δhxj=0,inB3r/5m(0);xj=zj,onB3r/5m(0).\begin{cases}\Delta_{h}x^{j}=0,&\text{in}\ B_{3r/5}^{m}(0);\\ x^{j}=z^{j},&\text{on}\ \partial B_{3r/5}^{m}(0).\end{cases}

Note that (3.1.1) yields that Br/2(p¯)φ1(B3r/5m(0))B_{r/2}(\bar{p})\subseteq\varphi^{-1}(B_{3r/5}^{m}(0)). We claim that

(x1,,xm):φ1(B3r/5m(0))m(x^{1},\dots,x^{m})\mathrel{\mathop{\mathchar 58\relax}}\varphi^{-1}(B_{3r/5}^{m}(0))\to\mathbb{R}^{m} (3.8)

yields the desired harmonic coordinate chart in Theorem 0.11, i.e., the following C1,αC^{1,\alpha}-estimates hold:

{e2Qδsth¯ste2Qδst;(r/2)1+αxjh¯stC0,α(Br/2(p¯))e2Q,\begin{cases}e^{-2Q}\delta_{st}\leq\bar{h}_{st}\leq e^{2Q}\delta_{st};\\ (r/2)^{1+\alpha}\left\|\frac{\partial}{\partial x^{j}}\bar{h}_{st}\right\|_{C^{0,\alpha}(B_{r/2}(\bar{p}))}\leq e^{2Q},\end{cases} (3.9)

where h¯st=h(xs,xt)\bar{h}_{st}=h(\frac{\partial}{\partial x^{s}},\frac{\partial}{\partial x^{t}}) and C0,αC^{0,\alpha}-norm is taken in the coordinates {xj}\{x^{j}\}.

Indeed, let us consider the functions wj=xjzjw^{j}=x^{j}-z^{j}, which satisfy the following equation:

{Δhwj=Δhzj,inB3r/5m(0);wj=0,onB3r/5m(0).\begin{cases}\Delta_{h}w^{j}=\Delta_{h}z^{j},&\text{in}\ B_{3r/5}^{m}(0);\\ w^{j}=0,&\text{on}\ \partial B_{3r/5}^{m}(0).\end{cases}

By (3.1.3),

ΔhzjC0,α(B3r/5m(0))Ψ(τ|n,α,Q),\left\|\Delta_{h}z^{j}\right\|_{C^{0,\alpha}\left(B_{3r/5}^{m}(0)\right)}\leq\Psi(\tau\,|\,n,\alpha,Q), (3.10)

where the C0,αC^{0,\alpha}-norm is taken in the coordinates {zj}\{z^{j}\}. By (3.1.1)-(3.1.2) and the Schauder estimates on Euclidean balls, there exists a constant C(r)=C(n,α,Q,r)>0C(r)=C(n,\alpha,Q,r)>0 such that

wjC2,α(B3r/5m(0))C(r)ΔhzjC0,α(B3r/5m(0))C(r)Ψ(τ|n,α,Q).\displaystyle\left\|w^{j}\right\|_{C^{2,\alpha}\left(B_{3r/5}^{m}(0)\right)}\leq C(r)\left\|\Delta_{h}z^{j}\right\|_{C^{0,\alpha}\left(B_{3r/5}^{m}(0)\right)}\leq C(r)\Psi(\tau\,|\,n,\alpha,Q). (3.11)

By the definition of C2,αC^{2,\alpha}-norm, it yields

xjzjC0(B3r/5m(0))\displaystyle\left\|x^{j}-z^{j}\right\|_{C^{0}\left(B_{3r/5}^{m}(0)\right)} +k=1mxjzkδjkC0(B3r/5m(0))+k,l=1m2xjzkzlC0(B3r/5m(0))\displaystyle+\sum_{k=1}^{m}\left\|\frac{\partial x^{j}}{\partial z^{k}}-\delta_{jk}\right\|_{C^{0}\left(B_{3r/5}^{m}(0)\right)}+\sum_{k,l=1}^{m}\left\|\frac{\partial^{2}x^{j}}{\partial z^{k}\partial z^{l}}\right\|_{C^{0}\left(B_{3r/5}^{m}(0)\right)} (3.12)
+k,l=1m2xjzkzlCα(B3r/5m(0))C(r)Ψ(τ|n,α,Q).\displaystyle+\sum_{k,l=1}^{m}\left\|\frac{\partial^{2}x^{j}}{\partial z^{k}\partial z^{l}}\right\|_{C^{\alpha}\left(B_{3r/5}^{m}(0)\right)}\leq C(r)\Psi(\tau\,|\,n,\alpha,Q).

By taking τ\tau sufficiently small, (3.12) implies that (3.8) is a harmonic coordinate chart.

What remains is to verify (3.9). Let B=(xjzk)B=(\frac{\partial x^{j}}{\partial z^{k}}) denote the coordinate translation matrix. Then B1=(zkxj)B^{-1}=(\frac{\partial z^{k}}{\partial x^{j}}) is the inverse matrix of BB. Therefore

zkxj=(polynomial in{xszt})/det(B),\frac{\partial z^{k}}{\partial x^{j}}=\left(\text{polynomial in}\left\{\frac{\partial x^{s}}{\partial z^{t}}\right\}\right)/\operatorname{det}(B), (3.13)

where det(B)\operatorname{det}(B) is the determinant of matrix BB. Note that

h¯ij=h(xi,xj)=h(zkxizk,zlxjzl)=zkxizlxjhkl.\displaystyle\bar{h}_{ij}=h\left(\frac{\partial}{\partial x^{i}},\frac{\partial}{\partial x^{j}}\right)=h\left(\frac{\partial z^{k}}{\partial x^{i}}\frac{\partial}{\partial z^{k}},\frac{\partial z^{l}}{\partial x^{j}}\frac{\partial}{\partial z^{l}}\right)=\frac{\partial z^{k}}{\partial x^{i}}\frac{\partial z^{l}}{\partial x^{j}}h_{kl}. (3.14)

Thus, (3.1.1)-(3.1.2) together with (3.12)-(3.14) yield the C1,αC^{1,\alpha}-estimates (3.9) in the coordinates {xj}\{x^{j}\}. ∎

4. C1,αC^{1,\alpha}-regularity on the limit space

From this section we begin to prove Theorem 0.1. We first show that any limit space X𝒳n,r,vm(δ,ρ)X\in\mathcal{X}_{n,r,v}^{m}(\delta,\rho) is a C1,αC^{1,\alpha}-Riemannian manifold for any 0<δδ(n)0<\delta\leq\delta(n) and 0<α<10<\alpha<1. Without loss of generality, we assume that r=ρr=\rho.

Let (Mi,gi)GHX(M_{i},g_{i})\overset{GH}{\longrightarrow}X, where (Mi,gi)(M_{i},g_{i}) are Riemannian nn-manifolds with |RicMi|n1|\operatorname{Ric}_{M_{i}}|\leq n-1 and (ρ,v)(\rho,v)-local covering geometry. By Lemma 1.12, we assume directly that (Mi,gi)(M_{i},g_{i}) is of (δ(n),ρ)(\delta(n),\rho)-Reifenberg local covering geometry, where δ(n)\delta(n) is the constant in (0.1).

As already pointed out in the introduction, it suffices to consider the equivariant Gromov-Hausdorff convergence:

(B^(xi,ρ2,ρ),x^i,Γi)iGH(Y,x^,G)πiπ(Bρ2(xi,gi),xi)iGH(Bρ2(x,X),x),\begin{CD}(\widehat{B}(x_{i},\frac{\rho}{2},\rho),\hat{x}_{i},\Gamma_{i})@>{GH}>{i\to\infty}>(Y,\hat{x},G)\\ @V{\pi_{i}}V{}V@V{\pi_{\infty}}V{}V\\ (B_{\frac{\rho}{2}}(x_{i},g_{i}),x_{i})@>{GH}>{i\to\infty}>(B_{\frac{\rho}{2}}(x,X),x),\end{CD} (4.1)

where the open balls and the normal covers are endowed with their length metrics, Γi\Gamma_{i} is the deck transformation by Γ12,ρ(xi)\Gamma_{\frac{1}{2},\rho}(x_{i}), and GG is the limit group of Γi\Gamma_{i}.

Let g^i=πigi\hat{g}_{i}=\pi_{i}^{*}g_{i} be the pullback Riemannian metric on the normal cover. By Remark 1.11, B^(xi,ρ2,ρ)\widehat{B}(x_{i},\frac{\rho}{2},\rho) admits a uniform C1,αC^{1,\alpha}-harmonic radius at points definitely away from the boundary. Then by Anderson’s convergence theorem 1.6, the interior YY^{\circ} of YY is a smooth manifold, where g^i\hat{g}_{i} gives rise to a C1,αC^{1,\alpha}-Riemannian metric tensor h^\hat{h} on YY^{\circ}.

Note that Y/GY/G is the limit of (Bρ2(xi,gi),xi)(B_{\frac{\rho}{2}}(x_{i},g_{i}),x_{i}) with its length metric. By the fact that (Bρ4(x,Y/G),dY/G)(B_{\frac{\rho}{4}}(x,Y/G),d_{Y/G}) endowed with the restricted metric is isometric to (Bρ4(x,X),dX)(B_{\frac{\rho}{4}}(x,X),d_{X}), and the same holds for any open ball centered at an interior point zY/Gz\in Y/G whose radius <ρ4d(x,z)2<\frac{\rho}{4}-\frac{d(x,z)}{2}, we derive that the interior of Y/GY/G is isometric to Bρ2(x,X)B_{\frac{\rho}{2}}(x,X) equipped with the length metric.

Proposition 4.1.

There is δ=δ(n)>0\delta=\delta(n)>0 such that

  1. (4.1.1)

    π\pi_{\infty} in (4.1) is a CC^{\infty}-smooth submersion that is Riemannian between C1,αC^{1,\alpha}-Riemannian metrics.

  2. (4.1.2)

    XX is a CC^{\infty}-smooth manifold (X,h)(X,h) with a C1,αC^{1,\alpha}-Riemannian metric.

We first verify that XX is regular, which is a direct corollary of the following observation.

Lemma 4.2.

Let GG be a closed subgroup of the isometry group of n\mathbb{R}^{n}. Then any tangent cone of the quotient space n/G\mathbb{R}^{n}/G is either isometric to an Euclidean space, or definitely ϵ(n)\epsilon(n)-away from any Euclidean space in the pointed Gromov-Hausdorff distance.

Proof.

Let G(o)G(o) be the orbit of GG at the origin oo of n\mathbb{R}^{n}, and GoO(n)G_{o}\subset O(n) the isotropy group at oo. Let To=mT_{o}^{\perp}=\mathbb{R}^{m} be the normal space of G(o)G(o) at oo, and o¯\bar{o} the quotient point of oo in n/G\mathbb{R}^{n}/G. By the standard theory of isometric group actions (e.g. see [35, Proposition 1.8]), the tangent cone To¯T_{\bar{o}} at o¯n/G\bar{o}\in\mathbb{R}^{n}/G is isometric to To/GoT_{o}^{\perp}/G_{o}. Hence it is an Euclidean space if and only if the action of GoG_{o} on ToT_{o}^{\perp} is trivial.

In the following we assume GoG_{o} acts on ToT_{o}^{\perp} non-trivially.

Case 1. the action of GoG_{o} on ToT^{\perp}_{o} is discrete. Then To/GoT_{o}^{\perp}/G_{o} is an Euclidean cone C(Σ)C(\Sigma) over Σ=𝕊m1/Go\Sigma=\mathbb{S}^{m-1}/G_{o}, whose volume is no more than half of 𝕊m1\mathbb{S}^{m-1}. Hence To¯T_{\bar{o}} is definitely away from m\mathbb{R}^{m}.

Case 2. the identity component of GoG_{o} acts on ToT^{\perp}_{o} non-trivially. Since the orbit of GoG_{o} on ToT^{\perp}_{o} must contain a great circle, the radius of 𝕊m1/Go\mathbb{S}^{m-1}/G_{o} is no more than π2\frac{\pi}{2}. Hence To¯T_{\bar{o}} is also definitely away from any Euclidean space. ∎

Lemma 4.3.

There is δ1(n)>0\delta_{1}(n)>0 such that for any δδ1(n)\delta\leq\delta_{1}(n), if XX satisfies the (δ,ρ)(\delta,\rho)-Reifenberg condition, then it is regular.

Proof.

By definition (0.1), the unit ball in any tangent cone TxT_{x} of XX is δ\delta-close to the unit Euclidean ball. At the same time, by (4.1) TxT_{x} is a quotient space of n\mathbb{R}^{n} by an isometric group action. Lemma 4.3 follows from Lemma 4.2 immediately. ∎

Next, we prove the key lemma in this section. Let us fix the pointed Gromov-Hausdorff approximation α^i:(B^(xi,ρ2,ρ),x^i,Γi)(Y,x^,G)\hat{\alpha}_{i}\mathrel{\mathop{\mathchar 58\relax}}(\widehat{B}(x_{i},\frac{\rho}{2},\rho),\hat{x}_{i},\Gamma_{i})\to(Y,\hat{x},G) in (4.1).

Lemma 4.4.

For any δ<δ1(n)\delta<\delta_{1}(n), the action of GG on YY in (4.1) is smooth and proper, such that the open ball Bρ2(x,Y/G)B_{\frac{\rho}{2}}(x,Y/G) is a C1,αC^{1,\alpha}-Riemannian manifold.

Proof.

Let us apply the Ricci flow on (Mi,gi)(M_{i},g_{i}). By Theorem 1.8, the solution gi(t)g_{i}(t) of Ricci flow (1.3) with initial condion gi(0)=gig_{i}(0)=g_{i} exists for 0<tT(n,ρ)0<t\leq T(n,\rho), a positive constant depending only on nn and ρ\rho, such that (1.4) holds for gi(t)g_{i}(t).

For any fixed t(0,T(n,ρ)]t\in(0,T(n,\rho)], let us view g^i(t)\hat{g}_{i}(t) as a sequence of metrics on B^(xi,ρ2,ρ)\widehat{B}(x_{i},\frac{\rho}{2},\rho) in (4.1). Then by the Arzela-Ascoli Theorem under Gromov-Hausdorff convergence, the underlying distance functions d^i,t\hat{d}_{i,t} subconverges to a distance function d^,t\hat{d}_{\infty,t}, which is e2te^{2t}-bi-Lipschitz to the underlying distance of h^\hat{h}. By (1.4) and Anderson’s C1,αC^{1,\alpha}-convergence theorem, up to a subsequence g^i(t)\hat{g}_{i}(t) uniformly converges to a smooth metric tensor h^(t)\hat{h}(t) on (Y,h^)(Y,\hat{h}) with the same regularity as gi(t)g_{i}(t) in (1.4). Thus, we have

(U^i,d^i,t,x^i,Γi)iα^i(Yt,h^(t),x^,Gt)πiπt,(Bρ2(xi,gi),di,t,xi)iGHYt/Gt,\begin{CD}(\widehat{U}_{i},\hat{d}_{i,t},\hat{x}_{i},\Gamma_{i})@>{\hat{\alpha}_{i}}>{i\to\infty}>(Y_{t},\hat{h}(t),\hat{x},G_{t})\\ @V{\pi_{i}}V{}V@V{\pi_{t,\infty}}V{}V\\ (B_{\frac{\rho}{2}}(x_{i},g_{i}),d_{i,t},x_{i})@>{GH}>{i\to\infty}>Y_{t}/G_{t},\end{CD} (4.2)

where U^i=B^(xi,ρ2,ρ)\widehat{U}_{i}=\widehat{B}(x_{i},\frac{\rho}{2},\rho), YtY_{t} is a smooth Riemannian manifold, and GtG_{t} is the limit group action of Γi\Gamma_{i}. (Furthermore, if α^i\hat{\alpha}_{i} is replaced by a local diffeomorphism ψi\psi_{i} that realizes the C1,αC^{1,\alpha}-convergence of g^i\hat{g}_{i} to h^\hat{h}, then it can be seen that ψig^i(t)\psi_{i}^{*}\hat{g}_{i}(t) CC^{\infty}-converges to h^(t)\hat{h}(t). We do not need this fact here.)

The key point here is that the action of Γi\Gamma_{i} on (U^i,g^i(t))(\widehat{U}_{i},\hat{g}_{i}(t)) and (U^i,g^i)(\widehat{U}_{i},\hat{g}_{i}) is the same as deck transformation of the normal cover πi:(U^i,x^i)(Bρ2(xi,gi),xi)\pi_{i}\mathrel{\mathop{\mathchar 58\relax}}(\widehat{U}_{i},\hat{x}_{i})\to(B_{\frac{\rho}{2}}(x_{i},g_{i}),x_{i}). And at the same time, g^i(t)\hat{g}_{i}(t) and g^i\hat{g}_{i}, and hence h^(t)\hat{h}(t) and h^\hat{h}, are two metric tensors on the same smooth manifolds U^i\widehat{U}_{i} and YY respectively, which by (1.4) are e2te^{2t}-bi-Lipschitz equivalent.

It follows from the Arzela-Ascoli Theorem that the limit action GtG_{t} of Γi\Gamma_{i} under gi(t)g_{i}(t) is also the same as GG under gig_{i}. Since the action of GtG_{t} is smooth and proper on (Y,h^(t))(Y,\hat{h}(t)), so is GG on (Y,h^)(Y,\hat{h}). Thus, we derive that πt,\pi_{t,\infty} coincides with π\pi_{\infty}, and Y/GY/G can be identified to Yt/GtY_{t}/G_{t} as topological spaces.

In order to show that Bρ2(x,Y/G)B_{\frac{\rho}{2}}(x,Y/G) is a C1,αC^{1,\alpha}-Riemannian manifold, we first prove that Yt/GtY_{t}/G_{t} is a smooth Riemannian manifold.

Indeed, the tangent cone of Y/GY/G is the same as that in XX, which by the (δ,ρ)(\delta,\rho)-Reifenberg condition, is δ\delta-close to be Euclidean on the unit ball at the vertex point. By the e2te^{2t}-bi-Lipschitz equivalence between Yt/GtY_{t}/G_{t} and Y/GY/G, the tangent cone of Yt/GtY_{t}/G_{t} is (e2t1+δ)(e^{2t}-1+\delta)-close to be Euclidean on the unit ball. By Lemma 4.2, for e2t1+δ<δ1(n)e^{2t}-1+\delta<\delta_{1}(n), Yt/GtY_{t}/G_{t} is regular.

Then by the standard theory of isometric actions on Riemannian manifolds (e.g., [3]), the slice representation of the isotropy group GpG_{p} at any interior point pYtp\in Y_{t} is trivial, hence the orbit Gt(p)G_{t}(p) is principle. By the principal orbit theorem (e.g. see [35, §1]), it follows that Yt/GtY_{t}/G_{t} is a smooth Riemannian manifold, and πt,\pi_{t,\infty} is a smooth Riemannian submersion.

Let us endow Bρ2(x,Y/G)B_{\frac{\rho}{2}}(x,Y/G) with the smooth structure on Yt/GtY_{t}/G_{t}. Because π:(Y,h^)Bρ2(x,Y/G)\pi_{\infty}\mathrel{\mathop{\mathchar 58\relax}}(Y,\hat{h})\to B_{\frac{\rho}{2}}(x_{\infty},Y/G) coincides with πt,\pi_{t,\infty}, π\pi_{\infty} is also a submersion. By the fact that GG acts on YY isometrically, the Riemannian metric tensor h^\hat{h} on YY induces a quotient Riemannian metric hh on Yt/GtY_{t}/G_{t}.

Furthermore, by the implicit function theorem, an adapted coordinate chart can always be constructed around preimages of a point in Bρ2(x,Y/G)B_{\frac{\rho}{2}}(x,Y/G). By the proof of Lemma 3.1, the adapted coordinate charts on YY descend to CC^{\infty}-admissible local coordinate charts on Bρ2(x,Y/G)B_{\frac{\rho}{2}}(x,Y/G), where the quotient metric tensor hh is C1,αC^{1,\alpha}. ∎

Proof of Proposition 4.1.

Let δ1(n)\delta_{1}(n) be that in Lemma 4.3 and δ=12δ1(n)\delta=\frac{1}{2}\delta_{1}(n). Then (4.1.1) has been proved in Lemma 4.4. In the following we prove (4.1.2). That is, XX admits a CC^{\infty}-smooth differentiable structure, such that the metric tensor hh induced locally from π\pi_{\infty} in (4.1) is C1,αC^{1,\alpha}.

It suffices to show the local charts induced from π\pi_{\infty} in (4.1) are CC^{\infty}-admissible with each other, where the metric tensors coincide with each other by pulling back.

Indeed, let Bρ2(x1,i,gi)B_{\frac{\rho}{2}}(x_{1,i},g_{i}) and Bρ2(x2,i,gi)B_{\frac{\rho}{2}}(x_{2,i},g_{i}) be two open balls, whose intersection Wi(ρ2)=Bρ2(x1,i,gi)Bρ2(x2,i,gi)W_{i}(\frac{\rho}{2})=B_{\frac{\rho}{2}}(x_{1,i},g_{i})\cap B_{\frac{\rho}{2}}(x_{2,i},g_{i}) is non-empty. Then the identity map ıj,i\imath_{j,i} from (Wi(ρ2),dWi(ρ2))(W_{i}(\frac{\rho}{2}),d_{W_{i}(\frac{\rho}{2})}) with its length metric to (Wi(ρ2),dBρ2(xj,i,gi))(W_{i}(\frac{\rho}{2}),d_{B_{\frac{\rho}{2}}(x_{j,i},g_{i})}) with the restricted metric is 11-Lipschitz and locally isometric. By the precompactness Theorem 1.10, the normal cover πi,W:W^i(ρ2,ρ)(Wi(ρ2),dWi(ρ2))\pi_{i,W}\mathrel{\mathop{\mathchar 58\relax}}\widehat{W}_{i}(\frac{\rho}{2},\rho)\to(W_{i}(\frac{\rho}{2}),d_{W_{i}(\frac{\rho}{2})}) sub-converges to π,W:W^(ρ2,ρ)W(ρ2)\pi_{\infty,W}\mathrel{\mathop{\mathchar 58\relax}}\widehat{W}_{\infty}(\frac{\rho}{2},\rho)\to W_{\infty}(\frac{\rho}{2}) as ii\to\infty. And the identity map ıj,i\imath_{j,i} sub-converges to ıj,:W(ρ2)Wj,(ρ2)Yj/Gj=limiBρ2(xj,i,gi)\imath_{j,\infty}\mathrel{\mathop{\mathchar 58\relax}}W_{\infty}(\frac{\rho}{2})\to W_{j,\infty}(\frac{\rho}{2})\subset Y_{j}/G_{j}=\lim_{i\to\infty}B_{\frac{\rho}{2}}(x_{j,i},g_{i}), which is also locally isometric.

Furthermore, by (1.5), πj,i,W:W^i(ρ2,ρ)πj,i1(Wi(ρ2))B^(xj,i,ρ2,ρ)\pi_{j,i,W}\mathrel{\mathop{\mathchar 58\relax}}\widehat{W}_{i}(\frac{\rho}{2},\rho)\to\pi_{j,i}^{-1}(W_{i}(\frac{\rho}{2}))\subset\widehat{B}(x_{j,i},\frac{\rho}{2},\rho) is a normal cover for j=1,2j=1,2. Let πj,:YjYj/Gj\pi_{j,\infty}\mathrel{\mathop{\mathchar 58\relax}}Y_{j}\to Y_{j}/G_{j} be the limit submersion of πj,i:B^(xi,ρ2,ρ)Bρ2(xi,gi)\pi_{j,i}\mathrel{\mathop{\mathchar 58\relax}}\widehat{B}(x_{i},\frac{\rho}{2},\rho)\to B_{\frac{\rho}{2}}(x_{i},g_{i}) as in (4.1). Then we derive the following commutative diagram

W^(ρ2,ρ)\textstyle{\widehat{W}_{\infty}(\frac{\rho}{2},\rho)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π1,,W\scriptstyle{\pi_{1,\infty,W}}π2,,W\scriptstyle{\pi_{2,\infty,W}}π,W\scriptstyle{\pi_{\infty,W}}π1,1(W1,(ρ2))\textstyle{\pi_{1,\infty}^{-1}(W_{1,\infty}(\frac{\rho}{2}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π1,\scriptstyle{\pi_{1,\infty}}π2,1(W2,(ρ2))\textstyle{\pi_{2,\infty}^{-1}(W_{2,\infty}(\frac{\rho}{2}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π2,\scriptstyle{\pi_{2,\infty}}W1,(ρ2)\textstyle{W_{1,\infty}(\frac{\rho}{2})}W(ρ2)\textstyle{W_{\infty}(\frac{\rho}{2})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ı2,\scriptstyle{\imath_{2,\infty}}ı1,\scriptstyle{\imath_{1,\infty}}W2,(ρ2)\textstyle{W_{2,\infty}(\frac{\rho}{2})}

where the maps ıj,\imath_{j,\infty} are locally isometric.

By the proof of Lemma 4.4, W(ρ2)W_{\infty}(\frac{\rho}{2}) is also a smooth manifold and π,W\pi_{\infty,W} is a smooth submersion. By the commutative diagram above, local coordinate charts on Wj,(ρ2)W_{j,\infty}(\frac{\rho}{2}) from Yj/GjY_{j}/G_{j} coincide with those on W(ρ2)W_{\infty}(\frac{\rho}{2}) descending from W^(ρ2,ρ)\widehat{W}_{\infty}(\frac{\rho}{2},\rho). Hence they are CC^{\infty}-admissible with each other. Moreover, the metric tensors on Wj,(ρ2)W_{j,\infty}(\frac{\rho}{2}) are also the same up to the diffeomorphism ıj,\imath_{j,\infty}. ∎

Remark 4.5.

Let (Mi,gi)GHX𝒳n,r,vm(δ(n),ρ)(M_{i},g_{i})\overset{GH}{\longrightarrow}X\in\mathcal{X}_{n,r,v}^{m}(\delta(n),\rho), where (Mi,gi)(M_{i},g_{i}) are Riemannian nn-manifolds with (r,v)(r,v)-local covering geometry, and δ=δ(n)\delta=\delta(n) is the minimum of that in (0.1) and in Proposition 4.1. By Theorem 1.8 (Dai-Wei-Ye [23], cf. [17]), the Ricci flow solution gi(t)g_{i}(t) with initial metric gi(0)=gig_{i}(0)=g_{i} admits a uniform positive existence time T(n,r,v,ρ)T(n,r,v,\rho), and the higher-ordered regularities (1.4). Assume that (Mi,gi(t))(M_{i},g_{i}(t)) sub-converges to XtX_{t}. We point out that, by the proof of Lemma 4.4, for not only tt satisfying e2t1+δ(n)<δ1(n)e^{2t}-1+\delta(n)<\delta_{1}(n), but also all t(0,T(n,r,v,ρ)]t\in(0,T(n,r,v,\rho)], XtX_{t} is regular. Hence, by the proof of Proposition 4.1, XtX_{t} is a smooth Riemannian manifold for any t(0,T(n,r,v,ρ)]t\in(0,T(n,r,v,\rho)].

Indeed, this can be seen from the fact that YtY_{t} shares the same isotropy group as Yt0Y_{t_{0}} for e2t01+δ(n)<δ1(n)e^{2t_{0}}-1+\delta(n)<\delta_{1}(n), or an open and closed argument on t0=sup{t:Xt is regular}t_{0}=\sup\{t\mathrel{\mathop{\mathchar 58\relax}}\text{$X_{t}$ is regular}\}, due to that e2t1=δ1(n)2e^{2t}-1=\frac{\delta_{1}(n)}{2} implies that the tangent cone of (X,h(t))(X,h(t)) is isometric to m\mathbb{R}^{m}, and hence t0t_{0} is extended further to cover 2t2t.

5. Adapted harmonic coordinates on the local covers of balls

Let the assumptions be as in the beginning of Section 4. In this section we construct adapted harmonic coordinates on B^(xi,ρ2,ρ)\widehat{B}(x_{i},\frac{\rho}{2},\rho) and on its limit space (Y,x^)(Y,\hat{x}) in the graph (4.1) respectively. Let h^\hat{h} be the Riemannian metric tensor on YY and hh its quotient metric tensor on XX.

Proposition 5.1.

There are constants δ(n)>0\delta(n)>0 and r1(n,r,v,ρ,α,Q)>0r_{1}(n,r,v,\rho,\alpha,Q)>0 such that for 0<δδ(n)0<\delta\leq\delta(n) and Rmin{δ1/2r1,δ1/4ρ}R\leq\min\{\delta^{-1/2}r_{1},\delta^{-1/4}\rho\}, there exists an adapted harmonic coordinate chart (f^1,,f^m,f^m+1,,f^n)(\hat{f}^{1},\dots,\hat{f}^{m},\hat{f}^{m+1},\dots,\hat{f}^{n}) defined on BR(x^,h^δ=δ1h^)B_{R}(\hat{x}_{\infty},\hat{h}_{\delta}=\delta^{-1}\hat{h}) such that f^k\hat{f}^{k} (k=1,,mk=1,\dots,m) descends to a smooth function fkf^{k} in BR(x,hδ)B_{R}(x_{\infty},h_{\delta}), and admits the following regularities:

{eQδklh^δ,kl=h^δ(f^k,f^l)eQδkl;R1+αh^δ,klf^jC0,α(BR(x^,h^δ))Ψ(δn,r,v,ρ,α,Q)\begin{cases}e^{-Q}\delta_{kl}\leq\hat{h}_{\delta,kl}=\hat{h}_{\delta}(\frac{\partial}{\partial\hat{f}^{k}},\frac{\partial}{\partial\hat{f}^{l}})\leq e^{Q}\delta_{kl};\\ R^{1+\alpha}\left\|\frac{\partial\hat{h}_{\delta,kl}}{\partial\hat{f}^{j}}\right\|_{C^{0,\alpha}(B_{R}(\hat{x},\hat{h}_{\delta}))}\leq\Psi(\delta\mid n,r,v,\rho,\alpha,Q)\end{cases} (5.1)

and

Hessf^jC0,α(BR(x^,h^δ))Ψ(δn,r,v,ρ,α,Q),j=1,,n,\|\operatorname{Hess}\hat{f}^{j}\|_{C^{0,\alpha}(B_{R}(\hat{x},\hat{h}_{\delta}))}\leq\Psi(\delta\mid n,r,v,\rho,\alpha,Q),\ \ j=1,\dots,n, (5.2)

where the C0,αC^{0,\alpha}-norm is taken in the coordinates {f^j}\{\hat{f}^{j}\}.

Assuming Proposition 5.1, we first prove Theorem 0.1.

Proof of Theorem 0.1.

Let δ=δ(n)\delta=\delta(n) be that in Proposition 5.1, and X𝒳n,r,vm(δ,ρ)X\in\mathcal{X}_{n,r,v}^{m}(\delta,\rho) a compact Ricci-limit space of Riemannian nn-manifolds (Mi,gi)(M_{i},g_{i}) with |RicMi|n1|\operatorname{Ric}_{M_{i}}|\leq n-1 and (r,v)(r,v)-local covering geometry, such that any point xXx\in X is (δ,ρ)(\delta,\rho)-Reifenberg. By Lemma 1.12, (Mi,gi)(M_{i},g_{i}) is of (δ,r)(\delta,r^{\prime})-Reifenberg local covering geometry for r=r(n,r,v,ρ)>0r^{\prime}=r^{\prime}(n,r,v,\rho)>0.

Let r1=r1(n,r,α,Q)r_{1}=r_{1}(n,r^{\prime},\alpha,Q) be as in Proposition 5.1, and let r0=min{r1,δ1/4ρ}r_{0}=\min\{r_{1},\delta^{1/4}\rho\}. By Proposition 4.1, X𝒳n,r,vm(δ,ρ)X\in\mathcal{X}_{n,r,v}^{m}(\delta,\rho) is a CC^{\infty}-smooth manifold with a C1,αC^{1,\alpha}-Riemannian metric tensor hh. Proposition 5.1 together with Theorem 0.11 implies the harmonic radius of (X,h)(X,h) is no less than r0r_{0}. The first part of (0.1.1) is complete.

For (0.1.2), let us assume that (X,h)𝒳n,r,vm(δ,ρ,D)(X,h)\in\mathcal{X}_{n,r,v}^{m}(\delta,\rho,D). Since (X,h)(X,h) is a Ricci-limit space, there are harmonic coordinate charts that covers (X,h)(X,h) whose number admits a uniform bound N(n,r0,D)N(n,r_{0},D), as well as the multiplicity of their intersections. By a standard argument, e.g., [4, Lemma 2.1], or [5, arguments below Theorem 0.2], Xn,r,vm(δ,ρ,D)X_{n,r,v}^{m}(\delta,\rho,D) is compact in the C1,αC^{1,\alpha}-topology.

What remains is to show (0.1.1.a-c).

Let (Mi,gi)GHX(M_{i},g_{i})\overset{GH}{\longrightarrow}X, where (Mi,gi)(M_{i},g_{i}) are Riemannian nn-manifolds with |RicMi|n1|\operatorname{Ric}_{M_{i}}|\leq n-1 and (δ,r)(\delta,r^{\prime})-Reifenberg local covering geometry. Let us consider the Ricci flow solution gi(t)g_{i}(t) with initial metric gi(0)=gig_{i}(0)=g_{i}, which exists for t(0,T(n,r)]t\in(0,T(n,r^{\prime})] by Theorem 1.8 (Dai-Wei-Ye [23], cf. [17]) and satisfies the higher-ordered regularities (1.4). By Remark 4.5, for all t(0,T(n,r))t\in(0,T(n,r^{\prime})) the limit space XtX_{t} of (Mi,gi(t))(M_{i},g_{i}(t)) is a smooth Riemannian manifold (Xt,h(t))(X_{t},h(t)), which is e2te^{2t}-bi-Lipschitz equivalent to (X,h)(X,h).

By Remark 4.5 again, there is ρ1(n)>0\rho_{1}(n)>0 such that (Xt,h(t))(X_{t},h(t)) lies in 𝒳n,r,vm(δ,ρ1,D1)\mathcal{X}_{n,r,v}^{m}(\delta,\rho_{1},D_{1}) for D1=eT(n,r)DD_{1}=e^{T(n,r^{\prime})}D. Hence it admits a uniform harmonic radius for all t[0,T(n,r)]t\in[0,T(n,r^{\prime})]. By the C1,αC^{1,\alpha}-precompactness, there is a diffeomorphism φt\varphi_{t} from XX to XtX_{t} such that the pullback metric φth(t)\varphi_{t}^{*}h(t) converges to hh in the C1,αC^{1,\alpha}-norm.

Let g^i(t)=πigi(t)\hat{g}_{i}(t)=\pi_{i}^{*}g_{i}(t) be the pullback metric on the normal cover B^(xi,r2,r)\widehat{B}(x_{i},\frac{r^{\prime}}{2},r^{\prime}) of Br2(xi,gi)B_{\frac{r^{\prime}}{2}}(x_{i},g_{i}) in the graph (4.2). Then the regularities (1.4) pass to the limit metric h^(t)\hat{h}(t) on YtY_{t}. By the proof of Lemma 3.1, the relations (3.4)-(3.5) between the quotient metric h(t)h(t) and h^(t)\hat{h}(t) imply that h(t)h(t) satisfies the same regularities (1.4) up to a definite ratio depending on nn and QQ. ∎

5.1. Preparation

Let us make some preparation for the proof of Proposition 5.1.

Lemma 5.2.

Let φ=(y1,,yn):Br(p)(M,h)Ωn\varphi=(y^{1},\dots,y^{n})\mathrel{\mathop{\mathchar 58\relax}}B_{r}(p)\subset(M,h)\to\Omega\subseteq\mathbb{R}^{n} be a harmonic coordinate chart at pp with φ(p)=0\varphi(p)=0 such that

eQδsthst=h(ys,yt)eQδste^{-Q}\delta_{st}\leq h_{st}=h(\frac{\partial}{\partial y^{s}},\frac{\partial}{\partial y^{t}})\leq e^{Q}\delta_{st} (5.3)

and

r1+αhstyjC0,α(Ω)eQ,r^{1+\alpha}\left\|\frac{\partial h_{st}}{\partial y^{j}}\right\|_{C^{0,\alpha}\left(\Omega\right)}\leq e^{Q}, (5.4)

where the C0,αC^{0,\alpha}-norm is taken in the coordinates {yj}\{y^{j}\}. Then after blowing up hh by λ2\lambda^{2}, where λ\lambda\to\infty, the harmonic coordinates φλ=(λyj):Bλr(p,hλ)λΩ\varphi_{\lambda}=(\lambda y^{j})\mathrel{\mathop{\mathchar 58\relax}}B_{\lambda r}(p,h_{\lambda})\to\lambda\Omega C2,αC^{2,\alpha}-converge to a Cartesian coordinate system (x1,,xn)(x^{1},\dots,x^{n}) in n\mathbb{R}^{n}, and hλ=λ2hh_{\lambda}=\lambda^{2}h C1,αC^{1,\alpha}-converge to an Euclidean metric gng_{\mathbb{R}^{n}}. That is, for 0<Rλ1/2r0<R\leq\lambda^{1/2}r

yλjφλ1xjC2,α(BRn(0))Ψ(λ1|r,α,Q,n)\left\|y_{\lambda}^{j}\circ\varphi_{\lambda}^{-1}-x^{j}\right\|_{C^{2,\alpha}\left(B_{R}^{n}(0)\right)}\leq\Psi(\lambda^{-1}\,|\,r,\alpha,Q,n) (5.5)

and

hλ,stφλ1δstC1,α(BRn(0))Ψ(λ1|r,α,Q,n),\left\|h_{\lambda,st}\circ\varphi_{\lambda}^{-1}-\delta_{st}\right\|_{C^{1,\alpha}\left(B_{R}^{n}(0)\right)}\leq\Psi(\lambda^{-1}\,|\,r,\alpha,Q,n), (5.6)

where Ψ(λ1|r,α,Q,n)0\Psi(\lambda^{-1}\,|\,r,\alpha,Q,n)\to 0 as λ10\lambda^{-1}\to 0 with the other variables fixed.

Proof.

This is an elementary fact. Let us identify (Br(p,h),h)(B_{r}(p,h),h) and (Ω,h=(φ1)h)(\Omega,h=(\varphi^{-1})^{*}h) with the pullback metric via the inverse of φ\varphi. Then each yjy^{j} is a harmonic function on (Ω,h)(\Omega,h). By a linear transformation, we can assume that hst(0)=δsth_{st}(0)=\delta_{st}. Put yλj=λyjy_{\lambda}^{j}=\lambda y^{j} and hλ,st=hλ(yλs,yλt)h_{\lambda,st}=h_{\lambda}(\frac{\partial}{\partial y_{\lambda}^{s}},\frac{\partial}{\partial y_{\lambda}^{t}}). Then hλ,st=hsth_{\lambda,st}=h_{st}, and by the rescaling property of the C1,αC^{1,\alpha}-harmonic radius in Remark 1.2,

λ1+αr1+αhλ,styλjC0,α(Ω)eQ,\lambda^{1+\alpha}\cdot r^{1+\alpha}\left\|\frac{\partial h_{\lambda,st}}{\partial y_{\lambda}^{j}}\right\|_{C^{0,\alpha}\left(\Omega\right)}\leq e^{Q}, (5.7)

where the C0,αC^{0,\alpha}-norm is taken in coordinates {yλj}\{y_{\lambda}^{j}\}.

Therefore, by (5.7) and the standard Schauder interior estimates, (yλ1,,yλn):(Ω,hλ)n(y_{\lambda}^{1},\dots,y_{\lambda}^{n})\mathrel{\mathop{\mathchar 58\relax}}(\Omega,h_{\lambda})\to\mathbb{R}^{n} C2,αC^{2,\alpha}-converge to a standard Cartesian coordinate system in n\mathbb{R}^{n}, and the metric hλh_{\lambda} C1,αC^{1,\alpha}-converge to the standard Euclidean metric. ∎

Let us assume that (X,h)(X,h) is (δ,ρ)(\delta,\rho)-Reifenberg. Let xXx\in X and xiMix_{i}\in M_{i} that converges to xx. Let us consider the Gromov-Hausdorff convergence associated to the normal cover πi:(U^i,g^i,x^i)(Bρ2(xi,gi),xi)\pi_{i}\mathrel{\mathop{\mathchar 58\relax}}(\widehat{U}_{i},\hat{g}_{i},\hat{x}_{i})\to(B_{\frac{\rho}{2}}(x_{i},g_{i}),x_{i}) as in the diagram (4.1).

By Lemma 1.12 and [4], there exists δ(n)>0\delta(n)>0 such that for given any 0<α<10<\alpha<1, Q>0Q>0 and 0δδ(n)0\leq\delta\leq\delta(n), the C1,αC^{1,\alpha}-harmonic radius of (Bρ4(x^i,g^i),g^i)(B_{\frac{\rho}{4}}(\hat{x}_{i},\hat{g}_{i}),\hat{g}_{i}) are uniformly bounded below by some constant r1=r1(n,r,v,ρ,α,Q)>0r_{1}=r_{1}(n,r,v,\rho,\alpha,Q)>0. Assume that x^i\hat{x}_{i} converges to x^Y\hat{x}\in Y. By the continuity of C1,αC^{1,\alpha}-harmonic radius in Proposition 1.5, so is Bρ4(x^,h^)YB_{\frac{\rho}{4}}(\hat{x},\hat{h})\subset Y.

After blowing up by δ1\delta^{-1}, it follows from the definition of (δ,ρ)(\delta,\rho)-Reifenberg and Lemma 5.2 that

(Bδ1/2ρ/2(x^,h^δ),h^δ)δ0C1,α(n,gn)πδπ0(Bδ1/2ρ/2(x,hδ),hδ)δ0GH(m,gm)\begin{CD}(B_{\delta^{-1/2}\rho/2}(\hat{x},\hat{h}_{\delta}),\hat{h}_{\delta})@>{C^{1,\alpha}}>{\delta\to 0}>(\mathbb{R}^{n},g_{\mathbb{R}^{n}})\\ @V{\pi_{\delta}}V{}V@V{}V{\pi_{0}}V\\ (B_{\delta^{-1/2}\rho/2}(x,h_{\delta}),h_{\delta})@>{GH}>{\delta\to 0}>(\mathbb{R}^{m},g_{\mathbb{R}^{m}})\end{CD} (5.8)

where πδ\pi_{\delta} is the restriction of π\pi_{\infty} in (4.1) on the local balls, hδ=δ1hh_{\delta}=\delta^{-1}h, and h^δ=πδ(hδ)\hat{h}_{\delta}=\pi_{\delta}^{*}(h_{\delta}). After passing to a subsequence, as δ0\delta\to 0 πδ\pi_{\delta} converge to a submetry π0\pi_{0}, which is the canonical projection from n\mathbb{R}^{n} to m\mathbb{R}^{m}.

Up to a suitable diffeomorphism, we view Br1(x^i,g^i)B_{r_{1}}(\hat{x}_{i},\hat{g}_{i}) as a fixed domain in n\mathbb{R}^{n} with metric g^i\hat{g}_{i}. Let φi:Br1(x^i,g^i)n\varphi_{i}\mathrel{\mathop{\mathchar 58\relax}}B_{r_{1}}(\hat{x}_{i},\hat{g}_{i})\to\mathbb{R}^{n} be a sequence of harmonic coordinates C2,αC^{2,\alpha}-converges to a limit harmonic coordinates φ:Br1(x^,h^)n\varphi\mathrel{\mathop{\mathchar 58\relax}}B_{r_{1}}(\hat{x},\hat{h})\to\mathbb{R}^{n} as ii\to\infty. Let g^i,δ=δ1g^i\hat{g}_{i,\delta}=\delta^{-1}\hat{g}_{i}. By Lemma 5.2 again, each

φi,δ=δ1/2φi:Bδ1/2r1(x^i,g^i,δ)n\varphi_{i,\delta}=\delta^{-1/2}\varphi_{i}\mathrel{\mathop{\mathchar 58\relax}}B_{\delta^{-1/2}r_{1}}(\hat{x}_{i},\hat{g}_{i,\delta})\to\mathbb{R}^{n} (5.9)

is uniformly C2,αC^{2,\alpha}-close to a Cartesian coordinates (x^i1,,x^in)(\hat{x}_{i}^{1},\dots,\hat{x}_{i}^{n}) on n\mathbb{R}^{n} as δ0\delta\to 0. By passing to a subsequence of {i}\{i\}, we may assume that (x^i1,,x^in)(\hat{x}_{i}^{1},\dots,\hat{x}_{i}^{n}) converges to a Cartesian coordinates (x^1,,x^n)(\hat{x}^{1},\dots,\hat{x}^{n}) with the following regularities: For any fixed 0<R<δ1/2r10<R<\delta^{-1/2}r_{1},

φi,δjx^ijC2,α(BRn(0))Ψ(δ|r1,α,Q,n)\left\|\varphi_{i,\delta}^{j}-\hat{x}_{i}^{j}\right\|_{C^{2,\alpha}\left(B_{R}^{n}(0)\right)}\leq\Psi(\delta\,|\,r_{1},\alpha,Q,n) (5.10)

and

g^i,δ,stδstC1,α(BRn(0))Ψ(δ|r1,α,Q,n),\left\|\hat{g}_{i,\delta,st}^{*}-\delta_{st}\right\|_{C^{1,\alpha}\left(B_{R}^{n}(0)\right)}\leq\Psi(\delta\,|\,r_{1},\alpha,Q,n), (5.11)

where (x^i1,,x^in)(\hat{x}_{i}^{1},\dots,\hat{x}_{i}^{n}) is the limit Cartesian coordinates system of φi,δ\varphi_{i,\delta} as δ0\delta\to 0, g^i,δ,st\hat{g}_{i,\delta,st}^{*} is the metric matrix of g^i,δ\hat{g}_{i,\delta}^{*} in coordinates φi,δ\varphi_{i,\delta}, and the norm is taken in the Euclidean coordinates {x^ij}\{\hat{x}_{i}^{j}\} .

Furthermore, up to composing an orthonormal transformation on (x^1,,x^n)(\hat{x}^{1},\dots,\hat{x}^{n}) (also on φi,δ\varphi_{i,\delta}), we assume that the limit projection π0:nm\pi_{0}\mathrel{\mathop{\mathchar 58\relax}}\mathbb{R}^{n}\to\mathbb{R}^{m} in the diagram (5.8)

π0(x^1,,x^m,,x^n)=(x^1,,x^m),\pi_{0}(\hat{x}^{1},\dots,\hat{x}^{m},\dots,\hat{x}^{n})=(\hat{x}^{1},\dots,\hat{x}^{m}), (5.12)

gives rise to a Cartesian coordiates on m\mathbb{R}^{m}.

5.2. Construction of the adapted harmonic coordinates

Based on the preparation above, let us begin the construction of the adapted harmonic coordinates on U^i\widehat{U}_{i} and their limit YY.

By the definition of (δ,ρ)(\delta,\rho)-Reifenberg, for all sufficiently large ii we have the Gromov-Hausdorff distance

dGH(Bδ1/2ρ(xi,δ1gi),Bδ1/2ρm(0))2δ1/2.d_{GH}(B_{\delta^{-1/2}\rho}(x_{i},\delta^{-1}g_{i}),B_{\delta^{-1/2}\rho}^{m}(0))\leq 2\delta^{1/2}.

Let αi:Bδ1/2ρ(xi,δ1gi)Bδ1/2ρm(0)m\alpha_{i}\mathrel{\mathop{\mathchar 58\relax}}B_{\delta^{-1/2}\rho}(x_{i},\delta^{-1}g_{i})\to B_{\delta^{-1/2}\rho}^{m}(0)\subset\mathbb{R}^{m} be an 2δ1/22\delta^{1/2}-Gromov-Hausdorff approximation. Let {e1,,em}\{e_{1},\dots,e_{m}\} be the orthonormal basis at the origin of m\mathbb{R}^{m} associated to the Cartesian coordinates (x^1,,x^m)(\hat{x}^{1},\dots,\hat{x}^{m}) in (5.12).

Let gi,δ=δ1gig_{i,\delta}=\delta^{-1}g_{i}. Then for each j=1,,mj=1,\dots,m, there exists pi,jBδ1/2ρ(xi,gi,δ)p_{i,j}\in B_{\delta^{-1/2}\rho}(x_{i},g_{i,\delta}) pairwise 2δ1/22\delta^{1/2}-close to δ1/2ρej\delta^{-1/2}\rho e_{j} for all large ii. Let

bij()=dgi,δ(pij,)dgi,δ(pij,xi),b_{i}^{j}(\cdot)=d_{g_{i,\delta}}(p_{i}^{j},\cdot)-d_{g_{i,\delta}}(p_{i}^{j},x_{i}), (5.13)

where dgi,δd_{g_{i,\delta}} denotes the distance induced by gi,δg_{i,\delta} on MiM_{i}. Let fijf_{i}^{j} be the solution of the following Dirichlet problem for Rδ1/4ρR\leq\delta^{-1/4}\rho,

{Δgi,δfij=0,inB4R(xi,gi,δ);fij=bij,onB4R(xi,gi,δ),\begin{cases}\Delta_{g_{i,\delta}}f_{i}^{j}=0,&\text{in}\ B_{4R}(x_{i},g_{i,\delta});\\ f_{i}^{j}=b_{i}^{j},&\text{on}\ \partial B_{4R}(x_{i},g_{i,\delta}),\end{cases} (5.14)

where Δgi,δ\Delta_{g_{i,\delta}} is the Laplace-Beltrami operator associated with metric gi,δg_{i,\delta}. Ric(Mi,gi,δ)(n1)δ1/2\operatorname{Ric}(M_{i},g_{i,\delta})\geq-(n-1)\delta^{1/2}, (fi1,,fim):BR(xi)m(f_{i}^{1},\dots,f_{i}^{m})\mathrel{\mathop{\mathchar 58\relax}}B_{R}(x_{i})\to\mathbb{R}^{m} forms a δ\delta-splitting map in the sense of Cheeger-Colding [10].

By [10, Theorem 6.68] (or [21, Lemma 1.23], or the quantitative maximum principles [9, §8] together with Abresch-Gromoll’s excess estimate [2]), there is δ(n)>0\delta(n)>0 such that for any 0<δδ(n)0<\delta\leq\delta(n) and Rδ1/4ρR\leq\delta^{-1/4}\rho, the following C0C^{0}-estimate holds for all sufficiently large ii:

|fijbij|Ψ(δ|n,ρ),inB2R(pi,gi,δ).\left|f_{i}^{j}-b_{i}^{j}\right|\leq\Psi(\delta\,|\,n,\rho),\quad\text{in}\ B_{2R}(p_{i},g_{i,\delta}). (5.15)

As πi:(U^i,g^i,δ)Bδ1/2ρ/2(xi,gi,δ)\pi_{i}\mathrel{\mathop{\mathchar 58\relax}}(\widehat{U}_{i},\hat{g}_{i,\delta})\to B_{\delta^{-1/2}\rho/2}(x_{i},g_{i,\delta}) is locally isometric, f^ij=fijπi\hat{f}_{i}^{j}=f_{i}^{j}\circ\pi_{i} is also harmonic, i.e.,

Δg^i,δf^ij=0,inπi1(B2R(pi,gi,δ)).\Delta_{\hat{g}_{i,\delta}}\hat{f}_{i}^{j}=0,\ \text{in}\ \pi_{i}^{-1}(B_{2R}(p_{i},g_{i,\delta})). (5.16)

In the following we show that (f^i1,,f^im,φi,δm+1,,φi,δn)(\hat{f}_{i}^{1},\dots,\hat{f}_{i}^{m},\varphi_{i,\delta}^{m+1},\dots,\varphi_{i,\delta}^{n}) still forms a harmonic coordinates. Since (Bδ1/2ρ/2(x^i,g^i,δ),g^i,δ)C1,α(V,h^δ)(Y,h^δ=δ1h^)\left(B_{\delta^{-1/2}\rho/2}(\hat{x}_{i},\hat{g}_{i,\delta}),\hat{g}_{i,\delta}\right)\overset{C^{1,\alpha}}{\longrightarrow}(V,\hat{h}_{\delta})\subset(Y,\hat{h}_{\delta}=\delta^{-1}\hat{h}) as ii\to\infty, there exists a sequence of diffeomorphisms ψi:(V,h^δ)Bδ1/2ρ/4(x^i,g^i,δ),\psi_{i}\mathrel{\mathop{\mathchar 58\relax}}(V,\hat{h}_{\delta})\to B_{\delta^{-1/2}\rho/4}(\hat{x}_{i},\hat{g}_{i,\delta}), with ψi(x^)=x^i\psi_{i}(\hat{x})=\hat{x}_{i}, such that the pullback metrics g^i,δ=ψig^i,δ\hat{g}_{i,\delta}^{*}=\psi_{i}^{*}\hat{g}_{i,\delta} (sub-)converge to h^δ\hat{h}_{\delta} in the C1,αC^{1,\alpha}-norm. From now on, in the following subsections let us identify (Bδ1/2ρ/4(x^i,g^i,δ),g^i,δ)(B_{\delta^{-1/2}\rho/4}(\hat{x}_{i},\hat{g}_{i,\delta}),\hat{g}_{i,\delta}) with (V,g^i,δ)(V,\hat{g}_{i,\delta}^{*}) via ψi\psi_{i}.

Lemma 5.3.

The map (f^i1,,f^im,φi,δm+1,,φi,δn):BR(x^,g^i,δ)n(-\hat{f}_{i}^{1},\dots,-\hat{f}_{i}^{m},\varphi_{i,\delta}^{m+1},\dots,\varphi_{i,\delta}^{n})\mathrel{\mathop{\mathchar 58\relax}}B_{R}(\hat{x},\hat{g}_{i,\delta}^{*})\to\mathbb{R}^{n} still forms a harmonic coordinates that converges to the Cartesian coordinates (x^1,,x^n)(\hat{x}^{1},\dots,\hat{x}^{n}) given by (5.12) in the C2,αC^{2,\alpha}-norm as ii\to\infty and δ0\delta\to 0.

Proof.

It suffices to show that f^ij\hat{f}_{i}^{j} is C2,αC^{2,\alpha}-close to φi,δj-\varphi_{i,\delta}^{j} for every j=1,,mj=1,\dots,m.

By (5.10), φi,δ\varphi_{i,\delta} is close to the Cartesian coordinates (x^ij)(\hat{x}_{i}^{j}). Moreover, by the assumption under (5.9), (x^ij)(\hat{x}_{i}^{j}) converges to the Cartesian coordinates (x^j)j=1n(\hat{x}^{j})_{j=1}^{n}, which projects to the Cartesian coordinates (x^j)j=1m(\hat{x}^{j})_{j=1}^{m}.

At the same time, let bijb_{i}^{j} be the Buseman-typed function defined in (5.13) in the construction of fijf_{i}^{j}. Then by the choice of pijp_{i}^{j}, for any fixed R>0R>0,

bij+x^jαiC0(BR(xi,gi,δ))Ψ(i1,δ|R,ρ)\left\|b_{i}^{j}+\hat{x}^{j}\circ\alpha_{i}\right\|_{C^{0}(B_{R}(x_{i},g_{i,\delta}))}\leq\Psi(i^{-1},\delta\,|\,R,\rho) (5.17)

where αi:Bδ1/2ρ(xi,gi,δ)Bδ1/2ρm(0)m\alpha_{i}\mathrel{\mathop{\mathchar 58\relax}}B_{\delta^{-1/2}\rho}(x_{i},g_{i,\delta})\to B_{\delta^{-1/2}\rho}^{m}(0)\subset\mathbb{R}^{m} is the Gromov-Hausdorff approximation.

By the construction above, φi,δ\varphi_{i,\delta} converges to (x^1,,x^n)(\hat{x}^{1},\dots,\hat{x}^{n}) as ii\to\infty and δ0\delta\to 0. Moreover, by (5.15) the difference between fijf_{i}^{j} and bijb_{i}^{j} goes to zero as δ0\delta\to 0 and ii\to\infty. Let (x1,,xm)(x^{1},\dots,x^{m}) be the Cartesian coordinates on m\mathbb{R}^{m} in (5.8) such that x^j=xjπ0\hat{x}^{j}=x^{j}\circ\pi_{0}. Then, by (5.12) and the triangle inequality below

f^ij+φi,δjfijπibijπi+bijπi+xjαiπi+xjαiπi+x^j+x^j+φi,δj,\|\hat{f}_{i}^{j}+\varphi_{i,\delta}^{j}\|\leq\|f_{i}^{j}\circ\pi_{i}-b_{i}^{j}\circ\pi_{i}\|+\|b_{i}^{j}\circ\pi_{i}+x^{j}\circ\alpha_{i}\circ\pi_{i}\|+\|-x^{j}\circ\alpha_{i}\circ\pi_{i}+\hat{x}^{j}\|+\|-\hat{x}^{j}+\varphi_{i,\delta}^{j}\|,

we derive

φi,δj+f^ijC0(B2R(x^,h^δ))Ψ(i1,δ|n,r,v,ρ,α,Q),for eachj=1,,m.\left\|\varphi_{i,\delta}^{j}+\hat{f}_{i}^{j}\right\|_{C^{0}\left(B_{2R}(\hat{x},\hat{h}_{\delta})\right)}\leq\Psi(i^{-1},\delta\,|\,n,r,v,\rho,\alpha,Q),\ \text{for each}\ j=1,\dots,m.

Since φi,δj+f^ij\varphi_{i,\delta}^{j}+\hat{f}_{i}^{j} is also harmonic, by the Schauder interior estimates,

φi,δj+f^ijC2,α(BR(x^,g^i,δ))\displaystyle\left\|\varphi_{i,\delta}^{j}+\hat{f}_{i}^{j}\right\|_{C^{2,\alpha}\left(B_{R}(\hat{x},\hat{g}_{i,\delta}^{*})\right)} C5(n,r,v,ρ,α,Q)φi,δj+f^ijC0(BR(x^,g^i,δ))\displaystyle\leq C_{5}(n,r,v,\rho,\alpha,Q)\left\|\varphi_{i,\delta}^{j}+\hat{f}_{i}^{j}\right\|_{C^{0}\left(B_{R}(\hat{x},\hat{g}_{i,\delta}^{*})\right)} (5.18)
Ψ(i1,δ|n,r,v,ρ,α,Q),\displaystyle\leq\Psi(i^{-1},\delta\,|\,n,r,v,\rho,\alpha,Q),

where the C2,αC^{2,\alpha}-norm is taken in coordinates φi,δ\varphi_{i,\delta}.

Now together with (5.10) and (5.11), (5.18) implies that

(y^i1,,y^in):=(f^i1,,f^im,φi,δm+1,,φi,δn):BRn(0,g^i,δ)n(\hat{y}_{i}^{1},\dots,\hat{y}_{i}^{n})\mathrel{\mathop{\mathchar 58\relax}}=(-\hat{f}_{i}^{1},\dots,-\hat{f}_{i}^{m},\varphi_{i,\delta}^{m+1},\dots,\varphi_{i,\delta}^{n})\mathrel{\mathop{\mathchar 58\relax}}B_{R}^{n}(0,\hat{g}_{i,\delta}^{*})\to\mathbb{R}^{n} (5.19)

still forms a harmonic coordinates that converges to the Cartesian coordinates (x^1,,x^n)(\hat{x}^{1},\dots,\hat{x}^{n}) in the C2,αC^{2,\alpha}-norm as ii\to\infty and δ0\delta\to 0, which satisfies

y^ijx^jC2,α(BRn(0,g^i,δ))Ψ(i1,δ|n,r,v,ρ,α,Q)\left\|\hat{y}_{i}^{j}-\hat{x}^{j}\right\|_{C^{2,\alpha}\left(B_{R}^{n}(0,\hat{g}_{i,\delta}^{*})\right)}\leq\Psi(i^{-1},\delta\,|\,n,r,v,\rho,\alpha,Q) (5.20)

and

g^i,δ,stδstC1,α(BRn(0,g^i,δ))Ψ(i1,δ|n,r,v,ρ,α,Q).\left\|\hat{g}_{i,\delta,st}^{*}-\delta_{st}\right\|_{C^{1,\alpha}\left(B_{R}^{n}(0,\hat{g}_{i,\delta}^{*})\right)}\leq\Psi(i^{-1},\delta\,|\,n,r,v,\rho,\alpha,Q). (5.21)

Now we are ready to prove Proposition 5.1.

Proof of Proposition 5.1.

Let (y^i1,,y^in)=(f^i1,,f^im,φi,δm+1,,φi,δn):BR(x^,g^i,δ)n(\hat{y}_{i}^{1},\dots,\hat{y}_{i}^{n})=(\hat{f}_{i}^{1},\dots,\hat{f}_{i}^{m},\varphi_{i,\delta}^{m+1},\dots,\varphi_{i,\delta}^{n})\mathrel{\mathop{\mathchar 58\relax}}B_{R}(\hat{x},\hat{g}_{i,\delta}^{*})\to\mathbb{R}^{n} be the harmonic coordinate chart constructed in Lemma 5.3. Since g^i,δ\hat{g}_{i,\delta} C1,αC^{1,\alpha}-converges to h^\hat{h} as ii\to\infty, by (5.20), (y^i1,,y^in)(\hat{y}_{i}^{1},\dots,\hat{y}_{i}^{n}) C2,αC^{2,\alpha}-converges to a limit harmonic coordinates (y^1,,y^n):BR(x^,h^δ)n(\hat{y}^{1},\dots,\hat{y}^{n})\mathrel{\mathop{\mathchar 58\relax}}B_{R}(\hat{x},\hat{h}_{\delta})\to\mathbb{R}^{n} as ii\to\infty.

For each j=1,,mj=1,\dots,m, by the construction of y^ij=f^ij=fijπi\hat{y}_{i}^{j}=\hat{f}_{i}^{j}=f_{i}^{j}\circ\pi_{i}, its limits y^j\hat{y}^{j} takes the same value along every π\pi_{\infty}-fiber, and thus it naturally descends to a C2,αC^{2,\alpha}-smooth function fjf^{j} on BR(x,hδ)δ1/2XB_{R}(x,h_{\delta})\subset\delta^{-1/2}X.

What remains is to verify (5.1) and (5.2). First, by (5.21), it is clear that the first inequality in (5.1) holds. Note that together with (5.2), the first inequality implies the second in (5.1). It suffices to verify (5.2).

Secondly, by the C2,αC^{2,\alpha}-convergence of coordinate functions and (5.20), it is clear that (5.2) holds.

Now the proof of Proposition 5.1 is complete. ∎

6. Harmonic radius estimate in terms of the volume

This section is devoted to the proof of Theorem 0.3.

Proof of Theorem 0.3.

Let δ=δ(n)\delta=\delta(n) be the constant in Theorem 0.1, and X𝒵n,δ,rm(δ)X\in\mathcal{Z}_{n,\delta,r}^{m}(\delta). For (0.3.1), let us first prove that XX is a C1,αC^{1,\alpha}-Riemannian manifold with a positive harmonic radius.

Indeed, let (Mi,gi)GHX(M_{i},g_{i})\overset{GH}{\longrightarrow}X, where (Mi,gi)(M_{i},g_{i}) have |RicMi|(n1)|\operatorname{Ric}_{M_{i}}|\leq(n-1) and (δ,r)(\delta,r)-Reifenberg local covering geometry. For any xXx\in X, let us consider the equivariant convergence of normal covers of r2\frac{r}{2}-balls in (4.1). Since every tangent cone TxT_{x} at xx is δ\delta-close to m\mathbb{R}^{m}, and the proofs of Propositions 4.1 and 5.1 still go through for XX. That is, XX is a smooth manifold and for any xXx\in X, the δ\delta-splitting map on (Mi,gi)(M_{i},g_{i}) defined by (5.14) from the Buseman functions by the closeness between TxT_{x} and m\mathbb{R}^{m} gives rise to an adapted harmonic coordinate chart that descends from YY in (4.1), which satisfies the uniform regularities (5.1) and (5.2). Then by Theorem 0.11, XX admits a C1,αC^{1,\alpha}-harmonic coordinate around any point xXx\in X. The continuity of C1,αC^{1,\alpha}-harmonic radius at points in XX yields a positive lower bound of the harmonic radius of XX.

Next, we show that the C1,αC^{1,\alpha}-harmonic radius at a point xXx\in X satisfying

Vol(BR(x,X))w>0\operatorname{Vol}(B_{R}(x,X))\geq w>0 (6.1)

admits a uniform bound r0(n,r,w,R)>0\geq r_{0}(n,r,w,R)>0.

Let us argue by contradiction. Assume that there is a sequence of spaces (Xj,hj)𝒵n,δ,rm(δ)(X_{j},h_{j})\in\mathcal{Z}_{n,\delta,r}^{m}(\delta), each of which contains a point xjXjx_{j}\in X_{j} satisfying (6.1), but the C1,αC^{1,\alpha}-harmonic radius rh(xj,hj)0r_{h}(x_{j},h_{j})\to 0. By passing to a subsequence we assume that (Xj,xj)GH(X,x)(X_{j},x_{j})\overset{GH}{\longrightarrow}(X,x).

Let (Mj,i,gj,i)GH(Xj,hj)(M_{j,i},g_{j,i})\overset{GH}{\longrightarrow}(X_{j},h_{j}), where (Mj,i,gj,i)(M_{j,i},g_{j,i}) has bounded Ricci curvature and (δ,r)(\delta,r)-Reifenberg local covering geometry. By Theorem 1.8, there is t0=t0(n,r)>0t_{0}=t_{0}(n,r)>0 such that any (Mj,i,gj,i)(M_{j,i},g_{j,i}) admits a smoothed metric gj,i(t0)g_{j,i}(t_{0}) with uniformly higher regularities (1.4). By passing to a diagonal subsequence, we assume that

(Mj,i,gj,i(t0))GHXj,t0as i for any fixed j, and Xj,t0GHXt0.(M_{j,i},g_{j,i}(t_{0}))\overset{GH}{\longrightarrow}X_{j,t_{0}}\quad\text{as $i\to\infty$ for any fixed $j$, and }\quad X_{j,t_{0}}\overset{GH}{\longrightarrow}X_{t_{0}}.

Since (Xj,hj)(X_{j},h_{j}) is regular and Xj,t0X_{j,t_{0}} is e2t0e^{2t_{0}}-bi-Lipschitz to (Xj,hj)(X_{j},h_{j}), by Lemma 4.2, each Xj,t0X_{j,t_{0}} is also regular. Hence by Proposition 4.1 Xj,t0X_{j,t_{0}} is a smooth Riemannian manifold (Xj,hj(t0))(X_{j},h_{j}(t_{0})). Moreover, by O’Neill’s formula applied on the Riemannian submersion πt0,\pi_{t_{0},\infty} in (4.2), the sectional curvature of hj(t0)h_{j}(t_{0}) is bounded below uniformly by C(n,r)t01/2C(n,r)t_{0}^{-1/2}.

Since the volume of B2R(xj,hj(t0))B_{2R}(x_{j},h_{j}(t_{0})) is bounded below by e2mt0we^{-2mt_{0}}w, their limit Xt0X_{t_{0}} is an Alexandrov space with curvature C(n,r)t01/2\geq C(n,r)t_{0}^{-1/2} of Hausdorff dimension mm.

By the proof of Theorem 2.1, the sectional curvature of (Xj,hj(t0))(X_{j},h_{j}(t_{0})) is also bounded uniformly from above. By [16, Theorem 4.7] the injectivity radius of (Xj,hj(t0))(X_{j},h_{j}(t_{0})) is bounded below by i0(n,r,w,R)>0i_{0}(n,r,w,R)>0. By Cheeger-Gromov’s C1,αC^{1,\alpha}-precompactness, Xt0X_{t_{0}} is regular. Then by the e2t0e^{2t_{0}}-bi-Lipschitz equivalence between Xt0X_{t_{0}} and the original limit XX and Lemma 4.2 again, XX is also regular.

Therefore, by the first part of (0.3.1) XX is a C1,αC^{1,\alpha}-Riemannian manifold that admits a positive harmonic radius r>0r_{\infty}>0. In particular, for any pXp\in X, (X,ϵi1h,p)(X,\epsilon_{i}^{-1}h,p) is ϰ(ϵi)\varkappa(\epsilon_{i})-close to (TpX,o)=(m,0)(T_{p}X,o)=(\mathbb{R}^{m},0). Then (Xj,ϵi1hj,xj)(X_{j},\epsilon_{i}^{-1}h_{j},x_{j}) is also ϰ(ϵi)\varkappa(\epsilon_{i})-close to an Euclidean space (m,0)(\mathbb{R}^{m},0) for fixed ii and any large jj. Then harmonic coordinate charts of a definite radius can be constructed at xj(Xj,hj)x_{j}\in(X_{j},h_{j}) by the proof of Theorem 0.1. It contradicts to that rj0r_{j}\to 0.

Now Lemma 5.2 and Theorem 0.8 together imply (0.3.2). ∎

For limit spaces under bounded local covering geometry, we have the following C1,αC^{1,\alpha}-regularity that depends on the space itself.

Let 𝒳n,r,vm(δ)\mathcal{X}_{n,r,v}^{m}(\delta) be the set consisting of all compact Ricci-limit spaces of Riemannian nn-manifolds with |RicM|n1|\operatorname{Ric}_{M}|\leq n-1 and (r,v)(r,v)-local covering geometry, such that each element X𝒳n,r,vm(δ)X\in\mathcal{X}_{n,r,v}^{m}(\delta) is δ\delta-almost regular.

Corollary 6.1.

Let (Mi,gi)GHX𝒳n,r,vm(δ)(M_{i},g_{i})\overset{GH}{\longrightarrow}X\in\mathcal{X}_{n,r,v}^{m}(\delta), where (Mi,gi)(M_{i},g_{i}) are Riemannian nn-manifolds with |RicMi|n1|\operatorname{Ric}_{M_{i}}|\leq n-1 and (r,v)(r,v)-local covering geometry. Then the followings hold for δ=δ2(n)\delta=\delta_{2}(n).

  1. (6.1.1)

    For any sufficient large ii and any xiMix_{i}\in M_{i}, the preimages of xix_{i} in the universal cover of Br(xi)B_{r}(x_{i}) admit a uniform C1,αC^{1,\alpha}-harmonic radius r0(X)>0\geq r_{0}(X)>0.

  2. (6.1.2)

    Any element X𝒳n,r,vm(δ)X\in\mathcal{X}_{n,r,v}^{m}(\delta) is regular in the sense that for any xXx\in X, any tangent cone TxXT_{x}X at xx is isometric to m\mathbb{R}^{m}.

Proof.

Let δ2(n)=ϵ(n)/2\delta_{2}(n)=\epsilon(n)/2, where ϵ(n)\epsilon(n) is the constant that satisfies both Lemma 1.12 and Lemma 4.2.

Suppose that xiMix_{i}\in M_{i} converges to xXx\in X, and (sjX,x)(s_{j}X,x) converges to a tangent cone (Tx,o)(T_{x},o) that is δ2(n)\delta_{2}(n)-close to m\mathbb{R}^{m}, where sjs_{j}\to\infty. Let us consider the equivariant pointed Gromov-Hausdorff convergence

(Bsj(x~ij,sj2g~ij),x~ij,Γi)GH(Y,x~,G)πiπ(Bsj(xij,sj2gij),xij)GH(Tx,o)\begin{CD}(B_{s_{j}}(\tilde{x}_{i_{j}},s_{j}^{2}\tilde{g}_{i_{j}}),\tilde{x}_{i_{j}},\Gamma_{i})@>{GH}>{}>(Y,\tilde{x},G)\\ @V{\pi_{i}}V{}V@V{}V{\pi}V\\ (B_{s_{j}}(x_{i_{j}},s_{j}^{2}g_{i_{j}}),x_{i_{j}})@>{GH}>{}>(T_{x},o)\end{CD}

By Lemma 1.12, for any large jj, x~ijBsj(x~ij,sj2g~ij)\tilde{x}_{i_{j}}\in B_{s_{j}}(\tilde{x}_{i_{j}},s_{j}^{2}\tilde{g}_{i_{j}}) is a (δ(n),ρ(n,v))(\delta(n),\rho(n,v))-Reifenberg point, where δ(n)\delta(n) is the constant in (0.1). By |RicMi|n1|\operatorname{Ric}_{M_{i}}|\leq n-1 and the arguments in [4], the C1,αC^{1,\alpha}-harmonic radius rh(x~ij,sj2g~ij)r0(n,v)>0r_{h}(\tilde{x}_{i_{j}},s_{j}^{2}\tilde{g}_{i_{j}})\geq r_{0}(n,v)>0. By the C1,αC^{1,\alpha}-precompactness, x~\tilde{x} is a regular point. Then by Lemma 4.2, TxT_{x} is isometric to m\mathbb{R}^{m}. So we derive (6.1.2).

For (6.1.1), let us consider the equivariant convergence of the normal covers of Br2(xi,gi)B_{\frac{r}{2}}(x_{i},g_{i}) in (4.1). By Lemma 4.2, every point in the limit space Y/GY/G of Br2(xi,gi)B_{\frac{r}{2}}(x_{i},g_{i}) is regular. Then by the proof of Lemma 1.12, the limit space of the normal cover B^(xi,r2,r)\widehat{B}(x_{i},\frac{r}{2},r) is also regular, and thus x^iB^(xi,r2,r)\hat{x}_{i}\in\widehat{B}(x_{i},\frac{r}{2},r) is (δ(n),ρ(X))(\delta(n),\rho(X))-Reifenberg. Now (6.1.1) follows. ∎

7. Construction of canonical fibration under bounded Ricci curvature

In this section, we prove Theorem 0.8.

Let us first observe that Theorem 0.8 can be reduced to the case that ρ=1\rho=1, where ρ\rho is the radius appearing in the Reifenberg condition and bounded local covering geometry.

Indeed, let (M,g)(M,g) be a closed Riemannian nn-manifold with bounded Ricci curvature and (δ(n),ρ)(\delta(n),\rho)-Reifenberg local covering geometry, and (X,h)(X,h) a closed C1,αC^{1,\alpha}-Riemannian mm-manifold in 𝒴nm(δ(n),ρ)\mathcal{Y}_{n}^{m}(\delta(n),\rho) for 0<α<10<\alpha<1 and 0<ρ10<\rho\leq 1, such that

dGH((M,g),(X,h))=ϵρ.d_{GH}((M,g),(X,h))=\epsilon\cdot\rho.

If ρ<1\rho<1, up to a rescaling ρ2\rho^{-2} on the metric gg and hh, (M,ρ2g)(M,\rho^{-2}g) is of (δ(n),1)(\delta(n),1)-Reifenberg local covering geometry and (X,ρ2h)𝒳nm(δ(n),1)(X,\rho^{-2}h)\in\mathcal{X}_{n}^{m}(\delta(n),1) with

dGH((M,ρ2g),(X,ρ2h))=ϵ.d_{GH}((M,\rho^{-2}g),(X,\rho^{-2}h))=\epsilon.

By viewing gg and hh as the rescaled metrics ρ2g\rho^{-2}g and ρ2h\rho^{-2}h respectively, we have ρ=1\rho=1 and

dGH((M,g),(X,h))=ϵ,d_{GH}((M,g),(X,h))=\epsilon,

such that all rescaling invariant constants as in (0.6.1)-(0.6.3) remain the same.

We first prove a slightly weaker version of Theorem 0.8. The center of mass technique below was applied earlier in [49] and [39] under the similar settings. For the difference see Remark 0.10.

Proposition 7.1.

Let (M,g)(M,g) be a closed Riemannian nn-manifold with |RicM|n1|\operatorname{Ric}_{M}|\leq n-1 and (δ(n),1)(\delta(n),1)-Reifenberg local covering geometry, and (X,h)𝒴nm(δ(n),1)(X,h)\in\mathcal{Y}_{n}^{m}(\delta(n),1). If dGH(M,X)ϵ<ϵ(n)d_{GH}(M,X)\leq\epsilon<\epsilon(n), then there is a C2,αC^{2,\alpha}-smooth fibration f:MXf\mathrel{\mathop{\mathchar 58\relax}}M\to X that satisfies

  1. (7.1.1)

    f:(M,g)(X,h)f\mathrel{\mathop{\mathchar 58\relax}}(M,g)\to(X,h) is a ϰ(ϵ|n)\varkappa(\epsilon\,|\,n)-almost Riemannian submersion, where after fixing nn, ϰ(ϵ|n)0\varkappa(\epsilon\,|\,n)\to 0 as ϵ0\epsilon\to 0;

  2. (7.1.2)

    the fiber Fy=f1(y)F_{y}=f^{-1}(y) has intrinsic diameter diamg(Fy)C(n)ϵ\operatorname{diam}_{g}(F_{y})\leq C(n)\epsilon for any yXy\in X;

  3. (7.1.3)

    the second fundamental form of ff satisfies |2f|g,hC(n)|\nabla^{2}f|_{g,h}\leq C(n);

  4. (7.1.4)

    FyF_{y} is diffeomorphic to an infra-nilmanifold.

Proof of Proposition 7.1.

Let α:(M,g)(X,h)\alpha\mathrel{\mathop{\mathchar 58\relax}}(M,g)\to(X,h) be an ϵ1\epsilon_{1}-Gromov-Hausdorff approximation (for simplicity, ϵ1\epsilon_{1}-GHA) with ϵ12dGH(M,X)\epsilon_{1}\leq 2d_{GH}(M,X). Without loss of generality, we assume ϵ1=ϵ\epsilon_{1}=\epsilon.

Step 1. First, we prove that for τ=τ(n,1)\tau=\tau(n,1), 0<ϵτ20<\epsilon\leq\tau^{2}, and any pMp\in M, there is a local fibration fp:Bτ100(p,g)(X,h)f_{p}\mathrel{\mathop{\mathchar 58\relax}}B_{\tau 100}(p,g)\to(X,h), which is a Ψ(ϵ|n)\Psi(\epsilon\,|\,n)-almost Riemannian submersion with a uniform second derivative control

|2fp|g,hC(n),|\nabla^{2}f_{p}|_{g,h}\leq C(n), (7.1)

and fpf_{p} satisfies

fpαC0(Bτ100(p,g),(X,h))τΨ(ϵ/τ|n).\|f_{p}-\alpha\|_{C^{0}(B_{\tau 100}(p,g),(X,h))}\leq\tau\cdot\Psi(\epsilon/\tau\,|\,n). (7.2)

Indeed, by the assumptions (M,g)(M,g) is of (δ(n),1)(\delta(n),1)-Reifenberg local covering geometry. Let us consider the equivariant closeness of the normal cover in the diagram (7.3):

(B^(p,12,1),g^,p^,Γ)ϵcloseC1,α(Y,h^,y^,G)ππ(B12(p,g),p)ϵcloseαp(B12(y,h),y)Y/G,\begin{CD}(\widehat{B}(p,\frac{1}{2},1),\hat{g},\hat{p},\Gamma)@>{C^{1,\alpha}}>{\epsilon-\text{close}}>(Y,\hat{h},\hat{y},G)\\ @V{\pi}V{}V@V{\pi_{\infty}}V{}V\\ (B_{\frac{1}{2}}(p,g),p)@>{\alpha_{p}}>{\epsilon-\text{close}}>(B_{\frac{1}{2}}(y,h),y)\subset Y/G,\end{CD} (7.3)

where αp\alpha_{p} is a Ψ(ϵ|n)\Psi(\epsilon\,|\,n)-GHA, whose restriction on B14(p,g)B_{\frac{1}{4}}(p,g) coincides with α\alpha.

As the arguments before Lemma 5.3 in Section 5.2, in the following we identify B14(p^,g^)B_{\frac{1}{4}}(\hat{p},\hat{g}) with an open domain VYV\subset Y with the the pullback metric g^=ψg^\hat{g}^{*}=\psi^{*}\hat{g} via a diffeomorphism ψ:(V,h^)YB14(p^,g^)\psi\mathrel{\mathop{\mathchar 58\relax}}(V,\hat{h})\subset Y\to B_{\frac{1}{4}}(\hat{p},\hat{g}), such that ψ(y^)=p^\psi(\hat{y})=\hat{p} and g^\hat{g}^{*} is C1,αC^{1,\alpha}-close to h^\hat{h}.

Let hτ=τ2hh_{\tau}={\tau}^{-2}h and gτ=τ2gg_{\tau}=\tau^{-2}g. Since by Theorem 0.1 the C1,αC^{1,\alpha}-harmonic radius of (X,h)(X,h) is no less than r0=r0(n,1,α,Q)>0r_{0}=r_{0}(n,1,\alpha,Q)>0, the rescaled (X,hτ)(X,h_{\tau}) has C1,αC^{1,\alpha}-harmonic radius at least τ1r0\tau^{-1}r_{0}. By Lemma 5.2, Bτ1/8r0(α(p),hτ)B_{\tau^{-1/8}r_{0}}(\alpha(p),h_{\tau}) is C(r0,Q,n)τ1/8C(r_{0},Q,n)\tau^{1/8}-close to an Euclidean ball Bτ1/8r0m(0)RmB^{m}_{\tau^{-1/8}r_{0}}(0)\subset R^{m}.

Let us consider the harmonic δ\delta-splitting map constructed in (5.14),

up,τ=(u1,,um):B200(p,gτ)m.u_{p,\tau}=(u^{1},\dots,u^{m})\mathrel{\mathop{\mathchar 58\relax}}B_{200}(p,g_{\tau})\to\mathbb{R}^{m}. (7.4)

Since the Ricci curvature of gτg_{\tau} satisfies Ricgτ(n1)τ2\operatorname{Ric}_{g_{\tau}}\geq-(n-1)\tau^{2}. By the triangle inequality, Bτ1/8r0(p,gτ)B_{\tau^{-1/8}r_{0}}(p,g_{\tau}) is (C(r0,Q,n)τ1/8+ϵ/τ)(C(r_{0},Q,n)\tau^{1/8}+\epsilon/\tau)-close to Bτ1/8r0m(0)B^{m}_{\tau^{-1/8}r_{0}}(0). Then there is τ=τ(r0,Q,n)\tau=\tau(r_{0},Q,n) such that for 0<ϵτ20<\epsilon\leq\tau^{2}, (5.15) holds for uju^{j} and the Buseman functions bjb^{j} as in (5.14).

In the following we fix τ=τ(r0,Q,n)\tau=\tau(r_{0},Q,n). By the proof of Proposition 5.1 applied on (7.3), the lifted harmonic functions u^p,τ\hat{u}_{p,\tau} on B^(p,1/2,1)\widehat{B}(p,1/2,1) together with other harmonic functions form a C1,αC^{1,\alpha}-harmonic coordinate chart H^p^,τ:B100(p^,g^τ)n\widehat{H}_{\hat{p},\tau}\mathrel{\mathop{\mathchar 58\relax}}B_{100}(\hat{p},\hat{g}^{*}_{\tau})\to\mathbb{R}^{n}, whose coordinate functions admit a uniform C2,αC^{2,\alpha}-norm bound C(n)C(n).

By the C2,αC^{2,\alpha^{\prime}}-compactness for 0<α<α0<\alpha^{\prime}<\alpha, H^p^,τ\widehat{H}_{\hat{p},\tau} is Ψ(ϵ/τ|n)\Psi(\epsilon/\tau\,|\,n)-C2,αC^{2,\alpha^{\prime}}-close to an adapted harmonic coordinate chart H^,τ:B100(p^,g^τ)n\widehat{H}_{\infty,\tau}\mathrel{\mathop{\mathchar 58\relax}}B_{100}(\hat{p},\hat{g}^{*}_{\tau})\to\mathbb{R}^{n} for π\pi_{\infty}, which by Lemma 3.1 descends to an almost harmonic coordinate chart H,τ=(u1,,um)H_{\infty,\tau}=(u_{\infty}^{1},\dots,u_{\infty}^{m}) on B100(α(p),hτ)B_{100}(\alpha(p),h_{\tau}), whose coordinate functions are C2,αC^{2,\alpha^{\prime}}.

By the almost commutative diagram (7.3),

|up,τH,τα|B100(p,gτ)Ψ(ϵ/τ|n),\left|u_{p,\tau}-H_{\infty,\tau}\circ\alpha\right|_{B_{100}(p,g_{\tau})}\leq\Psi(\epsilon/\tau\,|\,n), (7.5)

and since u^p,τ=up,τπ\hat{u}_{p,\tau}=u_{p,\tau}\circ\pi,

|up,τπH,τπ|C2,α(B100(p^,g^τ))Ψ(ϵ/τ|n)\left|u_{p,\tau}\circ\pi-H_{\infty,\tau}\circ\pi_{\infty}\right|_{C^{2,\alpha^{\prime}}(B_{100}(\hat{p},\hat{g}^{*}_{\tau}))}\leq\Psi(\epsilon/\tau\,|\,n) (7.6)

Let us define fp=H,τ1up,τ:B100(p,gτ)(X,hτ)f_{p}=H_{\infty,\tau}^{-1}\circ u_{p,\tau}\mathrel{\mathop{\mathchar 58\relax}}B_{100}(p,g_{\tau})\to(X,h_{\tau}). Then (7.5) implies (7.2).

By (7.6), fpπ:(V,g^)Xf_{p}\circ\pi\mathrel{\mathop{\mathchar 58\relax}}(V,\hat{g}^{*})\to X is C2,αC^{2,\alpha^{\prime}}-close to π:(V,h^)X\pi_{\infty}\mathrel{\mathop{\mathchar 58\relax}}(V,\hat{h})\to X, which is a Riemannian submersion. Hence fpf_{p} is an ϰ(ϵ|n)\varkappa(\epsilon\,|\,n)-Riemannian submersion.

By |2u^p,τ|B100(p^,g^τ)=|2up,τ|B100(p,gτ)\left|\nabla^{2}\hat{u}_{p,\tau}\right|_{B_{100}(\hat{p},\hat{g}^{*}_{\tau})}=\left|\nabla^{2}u_{p,\tau}\right|_{B_{100}(p,g_{\tau})} and the uniform bound on H,τH_{\infty,\tau} up to the 2nd covariant derivative in Lemma 3.1, fpf_{p} admits a uniformly bounded Hessian. After rescaling back, we derive (7.1).

Step 2. Secondly, in order to obtain local fibrations that can be glued together, all operations below are done with respect to the fixed ϵ\epsilon-GHA α:(M,g)(X,h)\alpha\mathrel{\mathop{\mathchar 58\relax}}(M,g)\to(X,h).

Let us consider the nearby metric h(t0)h(t_{0}) provided by Theorem 0.1 for fixed t0=12min{T(n,1),ln2}t_{0}=\frac{1}{2}\min\{T(n,1),\ln 2\}. For 0<δt00<\delta\leq\sqrt{t_{0}}, let hδ,t0=δ2h(t0)h_{\delta,t_{0}}=\delta^{-2}h(t_{0}) and hδ=δ2hh_{\delta}=\delta^{-2}h. By (0.1.1) for ρ=1\rho=1, |sec(X,hδ,t0)|C1(n)t01/2δ2C1(n)δ|\sec(X,h_{\delta,t_{0}})|\leq C_{1}(n)t_{0}^{-1/2}\delta^{2}\leq C_{1}(n)\delta, and the C1,αC^{1,\alpha}-harmonic radius of (X,hδ,t0)(X,h_{\delta,t_{0}}) is at least δ1r0(n)>0\delta^{-1}r_{0}(n)>0. It follows that the injectivity radius of (X,hδ,t0)(X,h_{\delta,t_{0}}) admits a lower bound δ1/2i0(n)>0\delta^{-1/2}i_{0}(n)>0. Hence the convexity radius of (X,hδ,t0)(X,h_{\delta,t_{0}}) is no less than 12δ1/2i0(n)\frac{1}{2}\delta^{-1/2}i_{0}(n). Without loss of generality, we assume that δ1/2i0(n)100\delta^{-1/2}i_{0}(n)\geq 100.

In the following we view the rescaled metric gδ=(τ/δ)2gτg_{\delta}=(\tau/\delta)^{2}g_{\tau} as a rescaling of gτg_{\tau}, where τ=τ(n,1)\tau=\tau(n,1) is provided by Step 1. Let {pj}j=1N\{p_{j}\}_{j=1}^{N} be a 11-net in (M,gδ)(M,g_{\delta}) with NN depends only on the dimension nn of MM. For each pjp_{j}, let fj:B100(pj,gτ)(X,hτ)f_{j}\mathrel{\mathop{\mathchar 58\relax}}B_{100}(p_{j},g_{\tau})\to(X,h_{\tau}) be the local fibrations constructed in Step 1. Since hδ,t0h_{\delta,t_{0}} is e2t0e^{2t_{0}}-bi-Lipschitz to hδh_{\delta}, (7.2) implies that

fjαC0(B100τ/δ(pj,gδ),(X,hδ,t0))(τ/δ)Ψ(ϵ/τ|n)\|f_{j}-\alpha\|_{C^{0}(B_{100\tau/\delta}(p_{j},g_{\delta}),(X,h_{\delta,t_{0}}))}\leq(\tau/\delta)\cdot\Psi(\epsilon/\tau\,|\,n) (7.7)

with respect to hδ,t0h_{\delta,t_{0}}.

By taking δ=τΨ(ϵ/τ|n)0\delta=\tau\sqrt{\Psi(\epsilon/\tau\,|\,n)}\to 0 such that (τ/δ)Ψ(ϵ/τ|n)0(\tau/\delta)\cdot\Psi(\epsilon/\tau\,|\,n)\to 0, we will glue such local fibrations fjf_{j} together with respect to the smoothed metric hδ,t0h_{\delta,t_{0}} to a ϰ(ϵ,δ|n)\varkappa(\epsilon,\delta\,|\,n)-Riemannian submersion f:(M,gδ)(X,hδ)f\mathrel{\mathop{\mathchar 58\relax}}(M,g_{\delta})\to(X,h_{\delta}) with respect to the rescaled original metric hδh_{\delta}.

Let ϕ:11\phi\mathrel{\mathop{\mathchar 58\relax}}\mathbb{R}^{1}\to\mathbb{R}^{1} be a smooth cut-off function such that ϕ|[0,10]1\phi|_{[0,10]}\equiv 1, ϕ|[20,)0\phi|_{[20,\infty)}\equiv 0, and |ϕ|,|ϕ′′|10|\phi^{\prime}|,|\phi^{\prime\prime}|\leq 10. Let ϕj(x)=ϕ(dhδ,t0(fj(pj),fj(x)))\phi_{j}(x)=\phi(d_{h_{\delta,t_{0}}}(f_{j}(p_{j}),f_{j}(x))). Let us consider the energy function

E:M×X,E(x,y)=12jϕj(x)dhδ,t0(fj(x),y)2.E\mathrel{\mathop{\mathchar 58\relax}}M\times X\to\mathbb{R},\qquad E(x,y)=\frac{1}{2}\sum_{j}\phi_{j}(x)d_{h_{\delta,t_{0}}}(f_{j}(x),y)^{2}.

By the construction of ϕj(x)\phi_{j}(x) and (7.7), for Ψ(ϵ/τ|n)<1\Psi(\epsilon/\tau\,|\,n)<1, E(x,)E(x,\cdot) is a strictly convex function in B10(αδ(x),hδ,t0)B_{10}(\alpha_{\delta}(x),h_{\delta,t_{0}}), and it takes a unique minimum point, cm(x)cm(x), that is Ψ(ϵ/τ)\sqrt{\Psi(\epsilon/\tau)}-close to αδ(x)\alpha_{\delta}(x) measured in hδ,t0h_{\delta,t_{0}}. We define

f:(M,g)(X,h),f(x)=cm(x)f\mathrel{\mathop{\mathchar 58\relax}}(M,g)\to(X,h),\qquad f(x)=cm(x)

Step 3. What remains is to verify (7.1.1)-(7.1.4).

Let us prove (7.1.1)-(7.1.3) first. By its definition ff is determined by the equations

F(x,y)=yE=jϕj(x)(12d(fj(x),)2)=0,F(x,y)=\partial_{y}E=\sum_{j}\phi_{j}(x)\cdot\nabla\left(\frac{1}{2}d(f_{j}(x),\cdot)^{2}\right)=0,

where yE\partial_{y}E is the gradient of EE with respect to yy. Note that, in the normal coordinates at y=f(x)y=f(x), F(x,y)F(x,y) can be written in the form

F(x,y)=(jϕj(x)fj1(x),,jϕj(x)fjm(x)),F(x,y)=\left(-\sum_{j}\phi_{j}(x)f_{j}^{1}(x),\dots,-\sum_{j}\phi_{j}(x)f_{j}^{m}(x)\right),

where fj(x)=(fj1(x),,fjm(x))f_{j}(x)=(f_{j}^{1}(x),\dots,f_{j}^{m}(x)) is the position vector of fj(x)f_{j}(x) in the normal coordinates of yy. Then the differential of ff is determined by df=(yF)1xFdf=-(\partial_{y}F)^{-1}\circ\partial_{x}F, where

xF=j[dϕjfj(x)+ϕj(x)dfj].-\partial_{x}F=\sum_{j}\left[d\phi_{j}\cdot f_{j}(x)+\phi_{j}(x)df_{j}\right].

By the sectional curvature bound C(n,1)t01/2C(n,1)t_{0}^{-1/2} of h(t0)h(t_{0}) provided by (0.1.1.b), 212rfj(x)2\nabla^{2}\frac{1}{2}r_{f_{j}(x)}^{2} is Ψ1(δ|n,t0)\Psi_{1}(\delta\,|\,n,t_{0})-close to the identity matrix EE, which is the Hessian of squared Euclidean distance. At the same time, by the Bishop-Gromov’s relative volume comparison, the count of jj with non-vanishing ϕj(x)\phi_{j}(x) can be chosen at most C2(n)C_{2}(n). Hence (yF)1(\partial_{y}F)^{-1} is also Ψ2(δ|n,t0)\Psi_{2}(\delta\,|\,n,t_{0})-close to (jϕj(x))1E\left(\sum_{j}\phi_{j}(x)\right)^{-1}\cdot E. It follows that for all sufficient small δ\delta,

|df(yF)1jϕj(x)dfj|2(jϕj(x))1|jdϕj(x)fj(x)|.\left|df-(\partial_{y}F)^{-1}\sum_{j}\phi_{j}(x)df_{j}\right|\leq 2\left(\sum_{j}\phi_{j}(x)\right)^{-1}\left|\sum_{j}d\phi_{j}(x)\cdot f_{j}(x)\right|. (7.8)

Since by (7.7) and the choice of δ\delta, fj(x)f_{j}(x) is Ψ(ϵ/τ|n)\sqrt{\Psi(\epsilon/\tau\,|\,n)}-close to f(x)f(x) whenever fj(x)f_{j}(x) is well-defined,

|fj(x)|Ψ(ϵ/τ|n),for xB40(pj,gδ).|f_{j}(x)|\leq\sqrt{\Psi(\epsilon/\tau\,|\,n)},\quad\text{for $x\in B_{40}(p_{j},g_{\delta})$.} (7.9)

Combing (7.8) and (7.7) together, we derive

|dfjϕj(x)kϕk(x)dfj|C3(n)(Ψ2(δ|n,t0)+Ψ(ϵ/τ|n)).\left|df-\sum_{j}\frac{\phi_{j}(x)}{\sum_{k}\phi_{k}(x)}df_{j}\right|\leq C_{3}(n)\left(\Psi_{2}(\delta\,|\,n,t_{0})+\sqrt{\Psi(\epsilon/\tau\,|\,n)}\right). (7.10)

In order to show that ff is ϰ(ϵ|n)\varkappa(\epsilon\,|\,n)-almost Riemannian submersion, it suffices to show that the local fibrations fjf_{j} nearby are C1C^{1}-close to each other.

Indeed, by the definition of fjf_{j} in Step 1, Hj,,τfj=upj,τH_{j,\infty,\tau}\circ f_{j}=u_{p_{j},\tau} is a harmonic map, which by (7.7) is Ψ(ϵ/τ|n)\sqrt{\Psi(\epsilon/\tau\,|\,n)}-close to each other up to a transformation in the intersection of their domains. In particular, for j1j2j_{1}\neq j_{2}

|upj1,τHj1,,τHj2,,τ1upj2,τ|B40(pj1,gδ)B40(pj2,gδ)Ψ(ϵ/τ|n).\left|u_{p_{j_{1}},\tau}-H_{j_{1},\infty,\tau}\circ H_{j_{2},\infty,\tau}^{-1}\circ u_{p_{j_{2}},\tau}\right|_{B_{40}(p_{j_{1}},g_{\delta})\cap B_{40}(p_{j_{2}},g_{\delta})}\leq\sqrt{\Psi(\epsilon/\tau\,|\,n)}. (7.11)

At the same time, by the construction of Hj,,τH_{j,\infty,\tau} (see also Proposition 5.1 and Lemma 3.1), Hj1,,τHj2,,τ1H_{j_{1},\infty,\tau}\circ H_{j_{2},\infty,\tau}^{-1} on B40(pj1,gδ)B40(pj2,gδ)B_{40}(p_{j_{1}},g_{\delta})\cap B_{40}(p_{j_{2}},g_{\delta}) is C(r0,Q,n)δC(r_{0},Q,n)\delta-C1,αC^{1,\alpha} close to a constant isometric transformation Aj1j2A_{j_{1}j_{2}} on m\mathbb{R}^{m}.

Now by Cheng-Yau’s gradient estimate [20] for the component harmonic functions of

upj1,τAj1j2upj2,τ+2max|upj1,τAj1j2upj2,τ|B30(x,gδ)B40(pj1,gδ)B40(pj2,gδ)u_{p_{j_{1}},\tau}-A_{j_{1}j_{2}}u_{p_{j_{2}},\tau}+2\max\left|u_{p_{j_{1}},\tau}-A_{j_{1}j_{2}}u_{p_{j_{2}},\tau}\right|_{B_{30}(x,g_{\delta})\subset B_{40}(p_{j_{1}},g_{\delta})\cap B_{40}(p_{j_{2}},g_{\delta})}

in the context of uniform lower Ricci curvature bound, it follows that, for ηj1(fj1(x))0\eta_{j_{1}}(f_{j_{1}}(x))\neq 0 and ηj2(fj2(x))0\eta_{j_{2}}(f_{j_{2}}(x))\neq 0,

|dfj1dfj2|(x)C4(n)Ψ(ϵ/τ|n).\left|df_{j_{1}}-df_{j_{2}}\right|(x)\leq C_{4}(n)\sqrt{\Psi(\epsilon/\tau\,|\,n)}. (7.12)

(Note that the support of ηj\eta_{j} lies in B20(pj,gδ)B_{20}(p_{j},g_{\delta}).)

Since fj:B40(pj,gδ)(X,hδ)f_{j}\mathrel{\mathop{\mathchar 58\relax}}B_{40}(p_{j},g_{\delta})\to(X,h_{\delta}) is a ϰ(ϵ|n)\varkappa(\epsilon\,|\,n)-Riemannian submersion, by C1C^{1}-closeness (7.12) for fjf_{j}, the average of dfjdf_{j} in (7.10) is a ϰ(ϵ|n)\varkappa(\epsilon\,|\,n)-Riemannian submersion. Then (7.10) implies (7.1.1).

The uniform bound on the 2nd fundamental form of ff in (7.1.3) follows from further calculations on the 2nd derivatives of implicit function and the uniform bounds (0.1.1.b-c) on Rm\operatorname{Rm} and Rm\nabla\operatorname{Rm} of h(t0)h(t_{0}) .

Indeed, the 2nd fundamental form of ff with respect to the Euclidean metric gyg_{y} in the normal coordinates from TyXT_{y}X can be expressed in matrix by

d2f=(yF)1(x2F+y(xF)df)(yF)1(x(yF)df+y2F(df)2),d^{2}f=-(\partial_{y}F)^{-1}\cdot\left(\partial_{x}^{2}F+\partial_{y}(\partial_{x}F)\cdot df\right)-(\partial_{y}F)^{-1}\cdot\left(\partial_{x}(\partial_{y}F)\cdot df+\partial_{y}^{2}F\cdot(df)^{2}\right), (7.13)

where x2F\partial_{x}^{2}F consists of the Hessian of each components of FF with respect to xx such that

x2F=j2ϕj(x)fj(x)+2dϕj(x)dfj(x)+ϕj(x)2fj(x).\partial_{x}^{2}F=\sum_{j}\nabla^{2}\phi_{j}(x)\cdot f_{j}(x)+2d\phi_{j}(x)\otimes df_{j}(x)+\phi_{j}(x)\cdot\nabla^{2}f_{j}(x). (7.14)

For the last term of (7.13), we note that yF=y2E\partial_{y}F=\partial^{2}_{y}E is a combination of 212rfj(x)2\nabla^{2}\frac{1}{2}r_{f_{j}(x)}^{2}, which is Ψ(δ|n,t0)\Psi(\delta\,|\,n,t_{0})-close to the identity matrix EE. So is its inverse (yF)1(\partial_{y}F)^{-1}. For x(yF)\partial_{x}(\partial_{y}F) and β=12rfj(x)2\beta=\frac{1}{2}r^{2}_{f_{j}(x)}, we have

dfj(X)(2β(Y,Y))\displaystyle\nabla_{df_{j}(X)}\left(\nabla^{2}\beta(Y,Y)\right) =dfj(X)(YYβ)dfj(X)((YY)β)\displaystyle=df_{j}(X)\left(YY\beta\right)-df_{j}(X)\left((\nabla_{Y}Y)\beta\right)
=YYdfj(X),β(YY)dfj(X)β.\displaystyle=YY\left<df_{j}(X),\nabla\beta\right>-(\nabla_{Y}Y)df_{j}(X)\beta.

Observe that, dfj(X),β=Xkfjtxkyshts(t0)\left<df_{j}(X),\nabla\beta\right>=X^{k}\frac{\partial f_{j}^{t}}{\partial x^{k}}y^{s}h_{ts}(t_{0}) in the normal coordinates (y1,,ym)(y^{1},\dots,y^{m}) at fj(x)f_{j}(x). It follows that x(y2E)\partial_{x}(\partial_{y}^{2}E) admits a uniform upper norm bound C(n,t0)C(n,t_{0}) with respect to gg and h(t0)h(t_{0}). So are y(xF)\partial_{y}(\partial_{x}F) and y2F\partial_{y}^{2}F.

At the same time, by (7.9) and (7.1) for δ=τΨ(ϵ/τ|n)\delta=\tau\sqrt{\Psi(\epsilon/\tau\,|\,n)}, each term in (7.14) is bounded by δC(n,t0)\delta\cdot C(n,t_{0}). It follows that |2f|gδ,hδ,t0C(n)Ψ(ϵ/τ|n)|\nabla^{2}f|_{g_{\delta},h_{\delta,t_{0}}}\leq C(n)\sqrt{\Psi(\epsilon/\tau\,|\,n)} with respect to the rescaled metric gδg_{\delta} and hδ,t0h_{\delta,t_{0}}.

After rescaling back, we derive the bound of 2nd fundamental form in (7.1.3) with respect to h(t0)h(t_{0}). By the C1,αC^{1,\alpha}-compactness in Theorem 0.1, it can be seen that after replacing h(t0)h(t_{0}) with the original metric hh, (7.1.3) still holds.

The inequality (7.1.2) follows from the same argument in [12, proof of (2.6.1) of the fibration theorem 2.6]. Indeed, up to a rescaling, let us assume that both the harmonic radius of (X,h)(X,h) and injectivity radius of (X,h(t0))(X,h(t_{0})) is 1\geq 1, where e2t01/2e^{2t_{0}}\leq 1/2. Suppose for some y(X,h)y\in(X,h), the intrinsic diameter

diamg(Fy)μϵ,\operatorname{diam}_{g}(F_{y})\leq\mu\epsilon,

where dGH(M,X)ϵd_{GH}(M,X)\leq\epsilon. By the second fundamental form bound in (7.1.3), the extrinsic diameter of other fibers over B1/4(y,h(t0))B_{1/4}(y,h(t_{0})) is no less than C6(n)μϵC_{6}(n)\mu\epsilon. By (7.1.1), at least C7(n)ϵmμC_{7}(n)\epsilon^{-m}\mu many of ϵ\epsilon-balls are required to cover f1(B1/4(y,h(t0)))f1(B1/2(y,h))f^{-1}(B_{1/4}(y,h(t_{0})))\subset f^{-1}(B_{1/2}(y,h)). However, by the existence of ϵ\epsilon-GHA α:(M,g)(X,h)\alpha\mathrel{\mathop{\mathchar 58\relax}}(M,g)\to(X,h), which is Ψ(ϵ|n)\Psi(\epsilon\,|\,n)-close to ff, and the harmonic coordinates on B1(y,h)B_{1}(y,h), at most C7(n)ϵmC_{7}(n)\epsilon^{-m} such balls are required. Hence μC8(n)\mu\leq C_{8}(n).

The same argument in [50, Step 3 of the proof of Proposition 6.6] yields (7.1.4). Here we give a simple proof by the regularities (7.1.2) and (7.1.3).

Recall that by Theorem 1.8, gg can be smoothed to g(t0)g(t_{0}) by the Ricci flow [23], whose sectional curvature is uniformly bounded by C(n)t01/2C(n)t_{0}^{-1/2}. By the C1,αC^{1,\alpha}-compactness, lifted to the universal cover of 11-balls on MM, the C1,αC^{1,\alpha}-norm between g~(t0)\tilde{g}(t_{0}) and g~\tilde{g} is uniformly bounded by C(t0,n)C(t_{0},n). So is the Levi-Civita connections of g(t0)g(t_{0}) and gg. By the expression of the second fundamental form in terms of Christoffel symbols, it is easy to see that Fy=f1(y)F_{y}=f^{-1}(y) with respect to g(t0)g(t_{0}) still satisfies (7.1.3). Hence, FyF_{y} with the induced metric by g(t0)g(t_{0}) admits a uniform sectional curvature bound C(t0,n)C(t_{0},n) with a small diameter C(n)ϵC(n)\epsilon. The Gromov’s almost flat theorem [33] (cf. also [59, 58]) implies (7.1.4)(\ref{local-fibration-to-X4}). ∎

Proof of Theorem 0.8.

The only difference from Proposition 7.1 is the CC^{\infty}-smoothness of the fibration f:MNf\mathrel{\mathop{\mathchar 58\relax}}M\to N. Note that the C2,αC^{2,\alpha^{\prime}}-regularity of ff in Proposition 7.1 is due to the limit coordinate chart H,τH_{\infty,\tau} on B100(α(p),hτ)B_{100}(\alpha(p),h_{\tau}) in the definition of fpf_{p}; see the paragraph below (7.6).

Let g(t)g(t) be the metric smoothed by Theorem 1.8, and h(t)h(t) be the metric on XX by the proof of Theorem 0.1. Let Ht,,τH_{t,\infty,\tau} be the corresponding limit coordinate chart on XX with respect to τ2g(t)\tau^{-2}g(t) and τ2h(t)\tau^{-2}h(t). Then Ht,,τH_{t,\infty,\tau} is CC^{\infty}-smooth and C2,αC^{2,\alpha}-converges to H,τH_{\infty,\tau} as t0t\to 0.

By replacing H,τH_{\infty,\tau} by Ht,,τH_{t,\infty,\tau} in the proof of Proposition 7.1, we derive Theorem 0.8. ∎

Remark 7.2.

We point out that the local fibration fp,τf_{p,\tau} with fixed τ=τ(n,ρ)\tau=\tau(n,\rho) are modeled on the Riemannian submersion π\pi_{\infty} in (7.3) is necessary in guarantee that both (7.1.1) and (7.1.3) hold at the same time.

If, instead of a rescaling of fp,τf_{p,\tau}, one picks up for each δ>0\delta>0 a local “δ\delta-splitting map” modeled on the Euclidean space as that in (5.14), then by Cheeger-Colding’s well-known L2L^{2}-estimates on their gradients and Hessians [11, 10], it also yields a Ψ(δ,ϵ/δ|n)\Psi(\delta,\epsilon/\delta\,|\,n)-almost Riemannian submersion fp,δ:B100(p,gδ)(X,hδ)f_{p,\delta}\mathrel{\mathop{\mathchar 58\relax}}B_{100}(p,g_{\delta})\to(X,h_{\delta}) for 0<ϵδ2(n)0<\epsilon\leq\delta^{2}(n) (cf. [50, Proposition 6.6]), such that

fp,δαC0(B100(p,gδ),(X,hδ))Ψ(δ,ϵ/δ|n),\|f_{p,\delta}-\alpha\|_{C^{0}(B_{100}(p,g_{\delta}),(X,h_{\delta}))}\leq\Psi(\delta,\epsilon/\delta\,|\,n), (7.15)

and

|2fp,δ|gδ,hδC(n).|\nabla^{2}f_{p,\delta}|_{g_{\delta},h_{\delta}}\leq C(n). (7.16)

But one immediately encounters the following issues:

  1. (7.2.1)

    to guarantee the local fibrations and the global fibration glued together are ϰ(ϵ)\varkappa(\epsilon)-almost Riemannian submersions, δ\delta has to approach 0;

  2. (7.2.2)

    after the local fibrations are re-defined locally for each δ\delta, the 2nd derivative (7.16) after rescaling back blows up as δ0\delta\to 0.

Without a solution of the above, only a rescaling invariant 2nd derivative control on a ϰ(ϵ|n)\varkappa(\epsilon\,|\,n)-almost Riemannian submersion can be derive, such as

|2f|g,h(x)diamg(Ff(x))C(n)ϵ1/2.\left|\nabla^{2}f\right|_{g,h}(x)\cdot\operatorname{diam}_{g}(F_{f(x)})\leq C(n)\epsilon^{1/2}. (7.17)
Remark 7.3.

There are several well-known methods (e.g., Hamilton’s Ricci flow [36] applied in [23] or Perelman’s pseudo-locality [52] (cf. [40]), embedding to Hilbert space by PDEs [54] (cf. [1])), by which a collapsed manifold (M,g)(M,g) with |RicM|n1|\operatorname{Ric}_{M}|\leq n-1 and (δ,r)(\delta,r)-local covering geometry can be smoothed to a nearby metric gtg_{t} that admits a uniform bounded sectional curvature depending on tt. Hence by Theorem 0.6 a fibration ftf_{t} exists such that (0.6.1-4) hold with respect to gtg_{t}.

In order to remove the perturbing error that arises from tt, such that the fibration ft:(M,g)(X,h)f_{t}\mathrel{\mathop{\mathchar 58\relax}}(M,g)\to(X,h) remains to be a ϰ(ϵ)\varkappa(\epsilon)-Gromov-Hausdorff approximation as ϵ0\epsilon\to 0, tt must go to zero. Moreover, since the sectional curvature of gtg_{t} generally blows up as t0t\to 0, some explicit curvature control on gtg_{t} (e.g., |secgt|αt|\operatorname{sec}_{g_{t}}|\leq\frac{\alpha}{t} in [40, Theorem 1.6] by Perelman’s pseudo-locality [52]) and an arbitrary small distance distortion (e.g., αt\leq\alpha\sqrt{t} in [40, Lemma 1.11]) on a definite scale are required. For details, see the proof of [40, Theorem B].

Under the same setting of Theorem 0.8, we construct fibrations in [44] satisfying (0.6.1-2) and (0.6.4) via the smoothed metric g(t)g(t) by the Ricci flow method and by suitably choosing the flow time with respect to dGH(M,X)d_{GH}(M,X), among which (7.17) holds as the best regularity for the 2nd order derivative.

Though (7.17) is strictly weaker than (0.6.3), we prove in [44] that all such fibrations are equivalent to each others as diffeomorphic types. Moreover, they are stable under Lipschitz perturbation on the metric gg. It justifies the regularity (7.17) is also suitable in describing the topology of full-rank collapsing phenomena under bounded Ricci curvature.

References

  • [1] U. Abresch. Über das glatten Riemannisher metriken. PhD thesis, Habilitationsschrift, Reinischen Friedrich-Willhelms-Universität Bonn, 1988.
  • [2] U. Abresch and D. Gromoll. On complete manifolds with nonnegative Ricci curvature. J. Amer. Math. Soc., 3(2):355–374, 1990.
  • [3] M. M. Alexandrino and R. G. Bettiol. Lie groups and geometric aspects of isometric actions, volume 8. Cham: Springer, 2015.
  • [4] M. T. Anderson. Convergence and rigidity of manifolds under Ricci curvature bounds. Invent. Math., 102(1):429–445, 1990.
  • [5] M. T. Anderson and J. Cheeger. Cα{C}^{\alpha}-compactness for manifolds with Ricci curvature and injectivity radius bounded below. J. Differential Geom., 35(2):265–281, 1992.
  • [6] P. Buser and H. Karcher. Gromov’s almost flat manifolds. Astérisque, 81. Société mathématique de France, 1981.
  • [7] E. Calabi and P. Hartman. On the smoothness of isometries. Duke Math. J., 37(4):741–750, 1970.
  • [8] J. Cheeger. Finiteness theorems for Riemannian manifolds. Amer. J. Math., 92(1):61–74, 1970.
  • [9] J. Cheeger. Degeneration of Riemannian metrics under Ricci curvature bounds. Accademia nazionale dei Lincei, Scuola Normale superiore, 2001. Classe di Science.
  • [10] J. Cheeger and T. H. Colding. Lower bounds on Ricci curvature and the almost rigidity of warped products. Ann. of Math., 144(1):189–237, 1996.
  • [11] J. Cheeger and T. H. Colding. On the structure of spaces with Ricci curvature bounded below. I. J. Differential Geom., 46(3):406–480, 1997.
  • [12] J. Cheeger, K. Fukaya, and M. Gromov. Nilpotent structures and invariant metrics on collapsed manifolds. J. Amer. Math. Soc., 5(2):327–372, 1992.
  • [13] J. Cheeger and M. Gromov. On the characteristic numbers of complete manifolds of bounded curvature and finite volume. In Differential Geometry and Complex Analysis, pages 115–154. Springer, Berlin, Heidelberg, 1985.
  • [14] J. Cheeger and M. Gromov. Collapsing Riemannian manifolds while keeping their curvature bounded. I. J. Differential Geom., 23(3):309–346, 1986.
  • [15] J. Cheeger and M. Gromov. Collapsing Riemannian manifolds while keeping their curvature bounded. II. J. Differential Geom., 32(1):269–298, 1990.
  • [16] J. Cheeger, M. Gromov, and M. Taylor. Finite propagation speed, kernel estimates for functions of the Laplace operator, and the geometry of complete Riemannian manifolds. J. Differential Geom., 17(1):15–53, 1982.
  • [17] L. Chen, X. Rong, and S. Xu. Quantitative volume space form rigidity under lower Ricci curvature bound II. Trans. Amer. Math. Soc., 370(6):4509–4523, 2018.
  • [18] L. Chen, X. Rong, and S. Xu. Quantitative volume space form rigidity under lower Ricci curvature bound I. J. Differential Geom., 113(2):227–272, 2019.
  • [19] S. Y. Cheng, P. Li, and S. T. Yau. On the upper estimate of the heat kernel of a complete Riemannian manifold. Amer. J. Math., 103(5):1021–1063, 1981.
  • [20] S. Y. Cheng and S. T. Yau. Differential equations on Riemannian manifolds and their geometric applications. Comm. Pure Appl. Math., 28(3):333–354, 1975.
  • [21] T. H. Colding. Ricci curvature and volume convergence. Ann. of Math., 145(3):477–501, 1997.
  • [22] T. H. Colding and A. Naber. Sharp Hölder continuity of tangent cones for spaces with a lower Ricci curvature bound and applications. Ann. of Math., 176, 2012.
  • [23] X. Dai, G. Wei, and R. Ye. Smoothing Riemannian metrics with Ricci curvature bounds. Manuscripta Math., 90(1):49–61, 1996.
  • [24] D. Deturck and J. Kazdan. Some regularity theorems in Riemannian geometry. Ann. Sci. Ec. Norm. Sup., 14(3):249–260, 1981.
  • [25] T. Fujioka. Fibration theorems for collapsing Alexandrov spaces. PhD thesis, Kyoto University, 2021.
  • [26] K. Fukaya. On a compactification of the set of Riemannian manifolds with bounded curvatures and diameters. In Curvature and Topology of Riemannian manifold, Lecture Notes in Math., Vol. 1201, pages 89–107. Springer, Berlin, Heidelberg, 1986.
  • [27] K. Fukaya. Collapsing Riemannian manifolds to ones with lower dimensions. J. Differential Geom., 25:139–156, 1987.
  • [28] K. Fukaya. A boundary of the set of the Riemannian manifolds with bounded curvatures and diameters. J. Differential Geom., 28:1–21, 1988.
  • [29] K. Fukaya. Collapsing Riemannian manifolds to ones with lower dimension. II. J. Math. Soc. Japan, 41(2):333–356, 1989.
  • [30] F. Galaz-Garcia, M. Kell, A. Mondino, and G. Sosa. On quotients of spaces with Ricci curvature bounded below. J. Funct. Anal., 275(6):1368–1446, 2018.
  • [31] L. Z. Gao. Convergence of Riemannian manifolds; Ricci and Ln/2{L}^{n/2}-curvature pinching. J. Differential Geom., 32(2):349–381, 1990.
  • [32] R. E. Green and H. Wu. Lipschitz convergence of Riemannian manifolds. Pacific J. Math., 131(1):119–141, 1988.
  • [33] M. Gromov. Almost flat manifolds. J. Differential Geom., 13(2):231–241, 1978.
  • [34] M. Gromov. Structures mètriques pour les variètès Riemanniennes. (French) [Metric structures for Riemann manifolds] Edited by J. Lafontaine and P. Pansu. Textes Mathématiques [Mathematical Texts], 1. CEDIC, Paris, 1981.
  • [35] K. Grove. Geometry of, and via, symmetries. University Lecture Series-American Mathematical Society, 27:31–53, 2002.
  • [36] Richard S. Hamilton. Three-manifolds with positive Ricci curvature. J. Differential Geom., 17(2):255–306, 1982.
  • [37] P. Hartman and A. Wintner. On the problem of geodesics in the small. Amer. J. Math., 73(1):132–148, 1951.
  • [38] E. Hebey and M. Herzlich. Harmonic coordinates, harmonic radius and convergence of Riemannian manifolds. Rend. Mat. Appl, 17(4):569–605, 1997.
  • [39] H. Huang. Fibrations and stability for compact group actions on manifolds with local bounded Ricci covering geometry. Front. Math. China, 15(1):69–89, 2020.
  • [40] H. Huang, L. Kong, X. Rong, and S. Xu. Collapsed manifolds with Ricci bounded covering geometry. Trans. Amer. Math. Soc., 373(11):8039–8057, 2020.
  • [41] S. Huang, X. Rong, and B. Wang. Collapsing geometry with Ricci curvature bounded below and Ricci flow smoothing. Symmetry, Integrability and Geometry: Methods and Applications, 16:123, 2020.
  • [42] S. Huang and B. Wang. Rigidity of the first Betti number via Ricci flow smoothing. arXiv preprint arXiv:2004.09762[math.DG], 2020.
  • [43] S. Huang and B. Wang. Ricci flow smoothing for locally collapsing manifolds. Calc. Var. Partial Differential Equations, 61(2):1–32, 2022.
  • [44] Z. Jiang, L. Kong, and S. Xu. Canonical nilpotent structure under bounded Ricci curvature and local covering geometry. Preprint, 2022.
  • [45] J. Jost and H. Karcher. Geometrische methoden zur gewinnung von a-priori-schranken fr harmonische abbildungen. Manuscripta Math., 40(1):27–77, 1982.
  • [46] V. Kapovitch and B. Wilking. Structure of fundamental groups of manifolds with Ricci curvature bounded below. arXiv preprint arXiv:1105.5955, 2011.
  • [47] A. Kasue. A convergence theorem for Riemannian manifolds and some applications. Nagoya Math. J., 114:21–51, 1989.
  • [48] J. Lott. Some geometric properties of the Bakry-Émery Ricci tensor. Comment. Math. Helv., 78(4):865–883, 2003.
  • [49] A. Naber and G. Tian. Geometric structures of collapsing Riemannian manifolds II. J. Reine Angew. Math., 744:103–132, 2018.
  • [50] A. Naber and R. Zhang. Topology and ϵ\epsilon-regularity theorems on collapsed manifolds with Ricci curvature bounds. Geom. Topol., 20(5):2575–2664, 2016.
  • [51] J. Pan and G. Wei. Examples of Ricci limit spaces with non-integer Hausdorff dimension. arXiv preprint arXiv:2106.03967, 2021.
  • [52] G. Perelman. The entropy formula for Ricci flow and its geometric applications. arXiv preprint arXiv:math/0211159 [math.DG], 2002.
  • [53] S. Peters. Convergence of Riemannian manifolds. Compositio Math., 62(1):3–16, 1987.
  • [54] P. Petersen, G. Wei, and R. Ye. Controlled geometry via smoothing. Comment. Math. Helv., 74(3):345–363, 1999.
  • [55] C. Pro and F. Wilhelm. Riemannian submersions need not preserve positive Ricci curvature. Proc. Amer. Math. Soc., 142(7):2529–2535, 2014.
  • [56] X. Rong. On the fundamental groups of manifolds of positive sectional curvature. Ann. of Math., 143(2):397–411, 1996.
  • [57] X. Rong. Manifolds of Ricci curvature and local rewinding volume bounded below (in Chinese). Sci. Sin. Math., 48(6):791–806, 2018.
  • [58] X. Rong. A new proof of the Gromov’s theorem on almost flat manifolds. preprint, 2022.
  • [59] E. A. Ruh. Almost flat manifolds. J. Differential Geom., 17(1):1–14, 1982.
  • [60] W. X. Shi. Deforming the metric on complete Riemannian manifolds. J. Differential Geom., 30(1):223–301, 1989.
  • [61] W. X. Shi. Ricci deformation of the metric on complete noncompact Riemannian manifolds. J. Differential Geom., 30(2):303–394, 1989.
  • [62] C. Sormani and G. Wei. Universal covers for Hausdorff limits of noncompact spaces. Trans. Amer. Math. Soc., 356(3):1233–1270, 2004.
  • [63] S. Xu. Precompactness of local Ricci bounded covering geometry. Preprint, 2022.