Convergence of Ricci-limit spaces under bounded Ricci curvature and local covering geometry I
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 -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 limit spaces.
2010 Mathematics Subject Classification:
53C23, 53C21, 53C20, 53240. 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 of all closed Riemannian -manifolds of sectional curvature , volume and diameter is precompact in the -topology (cf. also [53, 32, 47]). This -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 (resp. ) consisting of all closed Riemannian -manifolds of Ricci curvature (resp. ), injectivity radius and diameter is precompact in the -topology (resp. -topology). A weaker replacement of the injectivity radius is -Reifenberg condition [11] (cf. [40]) for , i.e., around each point in such a manifold the Gromov-Hausdorff distance
(0.1) |
where is the -ball at the origin in , is a constant depending only on , and is a small positive constant. A point satisfying (0.1) is called a -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 -Reifenberg condition.
By the convergence theorems above, a limit space of manifolds in (resp. ) in the Gromov-Hausdorff topology is always a smooth manifold with a -Riemannian metric (resp. -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 with normalized curvature bound is called -collapsed, if the volume of any unit ball at is less than . As the counterpart of Cheeger-Gromov’s convergence theorem, -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 -manifold with is said to have -local covering geometry, if for any , the local rewinding volume of , , where is a preimage point of in the (incomplete) Riemannian universal cover . Moreover, is said to have -Reifenberg local covering geometry, if for any , is a -Reifenberg point. By Cheeger-Fukaya-Gromov [12, Theorem 1.3], for any , there is such that any collapsed complete manifold with sectional curvature bound admits -Reifenberg local covering geometry.
In this paper, we prove the -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 be the set consisting of all compact Ricci-limit spaces of Riemannian -manifolds with and -local covering geometry, such that the dimension of is in the sense of Colding-Naber [22], and any point in is -Reifenberg.
Theorem 0.1 (bounded local covering geometry).
Given and positive integers , there are constants such that the followings hold.
-
(0.1.1)
Any is a -Riemannian manifold with a positive -harmonic radius for any . And for each , there is a nearby Riemannian metric on such that
-
(a)
, where as with other parameters fixed,
-
(b)
the sectional curvature of satisfies ,
-
(c)
the -derivative of curvature tensor
where the constants and depend only on the given parameters.
-
(a)
-
(0.1.2)
The subset is compact in the -topology. In particular, contains only finitely many diffeomorphism types.
Recently Naber and Zhang [50] proved the -regularity theorem for locally full-rank collapsed Riemannian manifolds with (lower) bounded Ricci curvature, which says that there are uniform constants such that for any complete Riemannan -manifold with and , if an open ball is -Gromov-Hausdorff close to a -ball in a lower dimensional Euclidean space , then the subgroup generated by loops at of length in the fundamental group has the full rank if and only if the preimages of in the universal cover of admit a uniform -harmonic radius . Recall that by the generalized Margulis lemma [46], the local fundamental group contains a nilpotent subgroup of finite index , where the rank of is defined to be that of , which is no more than (see e.g., [50, Theorem 2.27]).
Let be the set consisting of all compact Ricci-limit spaces of locally full-rank collapsed manifolds with , such that the dimension of is , and any point in is -Reifenberg. By Naber-Zhang’s -regularity, for some and .
Corollary 0.2 (locally full-rank collapsed limit spaces).
Since all closed manifolds in are contained in for some , Theorem 0.1 generalizes Anderson’s -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 -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 -Hausdorff measure (resp. the volume).
Let be the set consisting of all compact Ricci-limit spaces of Riemannian -manifolds with and -Reifenberg local covering geometry, such that each element is -almost regular in the sense that for any , any tangent cone at is -close to in the pointed Gromov-Hausdorff topolgy.
Theorem 0.3 (Reifenberg-bounded local covering geometry).
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 -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 -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 can be smoothed to metrics whose sectional curvature is bounded two-sided uniformly by , where , and -Hausdorff measure ; for details see Section 2, and also compare [28, Theorem 0.9] and [28, Corollary 0.11], which states is a smooth manifold with a continuous metric tensor inducing a 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 -convergence, one has to construct -harmonic coordinates directly on a limit space. This is the new ingredient in Theorem 0.1.
For those non-collapsed -manifolds with and -Reifenberg local covering geometry, the -precompactness of non-collapsed has been proved in [17, Theorem E].
It should be pointed out that, though a limit space in Theorems 0.1-0.3 admits the sythetic 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 is collapsed as in Theorem 0.3. On the other hand, for the original -Hausdorff measure on , we do not know whether the volume comparison is satisfied; compare [55].
Remark 0.5.
It is well known by [54] that once admits a positive -harmonic radius , it can be smoothed to a new metric , whose sectional curvature . The smoothed metric in Theorem 0.1 has a better order (0.1.1.b-c), which arises from the Ricci flow solutions on such manifolds, where is their limit metric under Gromov-Hausdorff topology.
Indeed, by [40] (see also Lemma 1.12 below), for any the manifolds in its converging sequence with bounded Ricci curvature and -local covering geometry satisfy -Reifenberg local covering geometry for some . Then by Dai-Wei-Ye [23] (see also Theorem 1.8 below), the solution of Ricci flow equation with initial value exists in for some constant , and satisfies (0.1.1.a-c) on the local universal cover for . Hence any limit space in admits a nearby metric that locally is a quotient orbit space of manifolds with bounded sectional curvature .
By elementary facts on Riemannian submersions (e.g., see the proof of Lemma 3.1 below), the curvature condition on naturally passes to the quotient metric in a harmonic coordinate chart (see the diagram (0.2) below). Note that, though the lower curvature bound can be always passed to 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 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 , , there are constants and such that the following holds.
Let and be a closed Riemannian -manifold and -manifold respectively, whose sectional curvature and injectivity radius satisfy
If with , then there is a -smooth fibration such that
-
(0.6.1)
is a -almost Riemannian submersion, i.e., for any vector perpendicular to an -fiber, where after fixing , as .
-
(0.6.2)
The intrinsic diameter of any -fiber over satisfies
-
(0.6.3)
The second fundamental form is bounded by
-
(0.6.4)
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 and , called -regular in [12], i.e., the curvature tensor satisfies for all integer . It is well-known that the dependence of 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 be the constant in Theorem 0.1.
Theorem 0.8.
Given and positive integers , there exist constants such that the following holds.
Theorem 0.8 also holds for closed Riemannian -manifolds with and -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 -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 -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 -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 is an Euclidean -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 -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 -estimates [10, 11] on the -splitting map. The subtle point is the balance between (0.6.1) and (0.6.3). For example, if a fiber bundle is constructed with respect to a smoothed metric by the earlier known methods, then is a -almost Riemannian submersion depending also on . In order to derive (0.6.1), has to approach , while and hence generally blows up as . Similar issue also happens in applying Cheeger-Colding’s -estimates. Instead, we apply the -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, has only -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 offered by Theorem 0.1, which admits a convex radius depends on , such that (0.6.1)-(0.6.4) are proved for the original metric with fixed .
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 -norm estimates of the metric tensor in these coordinates [8, 45]. The main ingredients are as follows.
Let be a sequence of Riemannian -manifolds with and -local covering geometry that converges to in the Gromov-Hausdorff topology. By Lemma 1.12 and [4], for , the -harmonic radius of in the universal cover of admits a uniform lower bound . For simplicity we assume .
According to the precompactness principle for domains with boundary [63] (see Theorem 1.9 below), there is a normal cover of equipped with its length metric such that by passing to a subsequence, it converges equivariantly in the pointed Gromov-Hausdorff topology:
(0.2) |
where is the deck-transformation of , is the limit group of , and the quotient is locally isometric to the limit ball on . Moreover, by the definition of (see Remark 1.11 below), it still admits a uniform -harmonic radius lower bound. Hence is a -Riemannian manifold.
In order to present the idea shortly, we first assume that is a smooth Riemannian manifold. Since the tangent cone of is the quotient space of , which is either isometric to or definitely away from , the -Reifenberg condition implies that is regular. By the standard theory of isometric actions on Riemannian manifolds (e.g. [35, §1]), is also a Riemannian manifold. Then we construct a harmonic coordinate on in the following two steps.
Step 1. Following Cheeger-Colding [10, 11], we construct a harmonic -splitting map . We lift to a harmonic -splitting map on , and then by appending other harmonic functions, we complete it to a harmonic coordinate chart .
Step 2. By taking limit of , we get a harmonic coordinate chart on such that takes the same value on each -orbit. Hence descends to a smooth map on . For simplicity such harmonic coordinate chart is called to be adapted for a submersion , i.e., each takes the same value along every -fiber.
By the -precompactness on via harmonic coordinates, admits a small Hessian up to a definite rescaling on the metric. By the technical result below, gives rise to a harmonic coordinate chart of definite size on .
Theorem 0.11.
Given any and integers , there is a constant such that the following holds.
Let be a Riemannian submersion from a Riemannian -manifold (may not complete) to a Riemannian -manifold. Suppose that there is an adapted -harmonic coordinate chart at with --control (see (1.1.1)-(1.1.2) below). If the Hessian of each adapted coordinate function satisfies
(0.3) |
where the -norm is taken in the coordinates , then there is a -harmonic coordinate chart at with --control.
Note that the limit space of the normal covers of balls in (0.2) is only a -Riemannian manifold, and in general, a -Riemannian manifold may even admit no standard exponential map (see [37], cf. [7]). In order to guarantee the arguments above, we will show that is still a smooth submersion between -Riemannian manifolds. This can be seen by applying the Ricci flow on to obtain smooth limits and , which share the same limit group action in (0.2); see Proposition 4.1 below. Thus a harmonic coordinate chart on 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 -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 is a smooth manifold with a -Riemannian metric. In section 5, we will construct -harmonic coordinate charts on , 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. -Convergence
In this subsection, we will introduce the concepts such as the harmonic radius of a Riemannian manifold, -convergence of a sequence of Riemannian manifolds. After that, we will give the well-known Cheeger-Gromov’s and Anderson’s -convergence theorems [8], [34], [4], [5] (cf. also [53], [32], [47], [31], [38]).
Definition 1.1.
Given and . Let be a smooth -manifold with a -Riemannian metric . For any , we define the -harmonic radius at as the largest number such that on the geodesic ball of radius centered at , there is a harmonic coordinate chart such that the metric tensor admits the following --control:
-
(1.1.1)
as bilinear forms, where , for , and are the Kronecker symbols,
-
(1.1.2)
, which means
holds for and the distance associated with .
The harmonic radius of is now defined by . For simplicity we will omit and when there is no confusion.
In general, the -norms of the components, , of metric in the coordinates are defined on the Euclidean domain . For convenience, we also denote the -norm of on by .
Note that, and in (1.1.1) and (1.1.2) can be replaced equivalently with and respectively, where is the gradient of the coordinate function with respect to metric .
Remark 1.2.
By definition, for .
Now we give the concept of -convergence of a sequence of Riemannian manifolds.
Definition 1.3.
Let be a closed smooth -manifold. Let and be complete -smooth Riemannian metrics on . We say that converges to in the sense of -norm if for any , there exists a coordinate chart around , , such that -converges to as . i.e.,
(1.1) |
where the -norm of a smooth function is defined by
and .
In practice, an open cover of coordinate charts are usually fixed as the background coordinate charts for the -convergence on .
Definition 1.4.
Let and be closed smooth -manifolds with -smooth Riemannian metrics. We say that converges to in the -topology if there exists an integer such that the following holds: for each there exists -diffeomorphism such that the pullback metric converges to in the sense of -norm.
Note that, the pullback of by a diffeomorphism is crucial in Definition 1.4: even if converges to on the same manifold in the sense of -topology, it does not mean that converges to in the -norm. A counterexample can be found in [38, Remark 3 below Main theorem].
We say that a sequence of pointed complete Riemannian manifold converges in the -topology to a limit if
-
(1.4.1)
there exists an exhaustion of by open subsets such that and ;
-
(1.4.2)
there exists a sequence of -embeddings such that
uniformly on any compact subset of .
Let us view as the domain in Definition 1.1 with the pullback metric by , where is a -harmonic coordinate chart . Then the Cartesian coordinates on is harmonic with respect to , which satisfies the --control (1.1.1)-(1.1.2). In the harmonic coordinates for a metric tensor , the Ricci curvature satisfies the following equation:
where is a quadratic term in for (cf. [24]). By the standard -estimate for elliptic PDEs, the -norm of admits a uniform bound that depends on -norm of , -bound on the term and the -bound on for any . Hence a subsequence of metric tensors converges to a limit -metric in the -norm if where for any . More generally, if converges to in the -norm with respect to another fixed coordinates on , then similarly by the elliptic regularity, the harmonic coordinates of admit a uniform -bound, which implies that they converge to the harmonic coordinates of in the -norm.
Conversely, given a harmonic coordinate chart at with --control for , the Dirichlet problem associated with -nearby metric can be solved on , with boundary value for each . And the Schauder estimates give almost the same -control in the interior of .
Therefore, the -harmonic radius under bounded Ricci curvature is continuous in the sense of -topology, i.e., the following proposition.
Proposition 1.5 ([4],[5]).
Let be a sequence of Riemannian manifolds with , which -converges to a -Riemannian manifold . Then
The same holds for and as converges to .
Cheeger-Gromov-Anderson’s -convergence theorem says that the diffeomorphism types of the whole manifolds are also stable under the -topology, which provides a fundamental tool in this paper.
Theorem 1.6 ([4]).
The space of all closed Riemannian -manifolds such that
(1.2) |
is precompact in the -topology for any . More precisely, any sequence of -manifolds admits a subsequence that converges to a closed smooth manifold with a -Riemannian metric via -smooth diffeomorphisms in the -topology.
In particular, there are only finitely many diffeomorphism types of -manifolds satisfying (1.2).
1.2. Ricci Flows
Let be a closed Riemannian manifold. The Ricci flow was introduced by Hamilton [36] as the solution of the following degenerate parabolic PDE,
(1.3) |
The solution always exists for a short time , and if it admits a finite maximal flow time , then the curvature tensor blows up as , i.e., .
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 exists for a closed -manifold satisfying and the conjugate radius . As already pointed out by [17], the conjugate radius condition in their proof is only used to derive a uniform -harmonic coordinates for the lifted metric on for all (see [23, Remark 1]) and , which is required to apply the weak maximum principle [23, Theorem 2.1]. Since the same holds at a preimage point on the universal covering space of a -ball on when has -local covering geometry, [23, Theorem 1.1] can be reformulated into the following form.
Theorem 1.8 ([23], cf. [17, Theorem 1.5]).
Given , there exist constants and such that for any , if is a closed -manifold with and -Reifenberg local covering geometry, then the Ricci flow equation (1.3) has a unique smooth solution for satisfying
(1.4) |
where denotes the curvature tensor of , the -convariant derivative of , whose norm is measured in .
1.3. Gromov-Hausdorff precompactness for the covering spaces of open balls
Let be a sequence of complete Riemannian -manifolds with . Let be an open -ball in , and the Riemannian universal cover of . Then 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 , which particularly is suitable for the covering spaces of open balls.
Let be a component of the preimage in the Riemannian universal cover . Then is a normal -cover of , where
We endow with a base point in the preimage of and its length metric induced from .
Theorem 1.9 ([63]).
For any , let be the set consisting of the Riemannian normal covers of all open balls in complete Riemannian -manifolds with . Then is precompact in the pointed Gromov-Hausdorff topology.
More generally, let be a non-empty intersection of open -balls in , and be the Riemannian universal cover of . For , we define to be a component of the preimage of in the Riemannian universal cover .
Theorem 1.10 ([63]).
For any , let be the set consisting of the normal covers endowed with length metric of in complete Riemannian -manifolds with . Then is precompact in the pointed Gromov-Hausdorff topology.
Note that for each , a component of in the universal cover is a normal cover of , such that covers . By the definition of , we derive that
is a normal cover of . | (1.5) |
This fact will be applied in Section 4.
Remark 1.11.
Since is a component of the preimage in the Riemannian universal cover , the -Reifenberg local covering geometry condition is naturally passed to . That is, if for any , is a -Reifenberg point, then is -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 and real numbers , there are constants such that the following holds.
Let be a Riemannian -manifold with and
where is a preimage point of in the universal cover of . If
then is a -Reifenberg point, i.e.,
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 with , , , but is not a -Reifenberg point for any fixed for all sufficient large .
By passing to a subseqence, let us consider the equivariant pointed Gromov-Hausdorff convergence as in (0.2)
Since , a standard blowing-up argument implies that the quotient of any tangent cone modular the infinitesimal actions induced by is , where the lines on can be lifted onto . At the same time, by the fact that is a Ricci-limit space of a non-collapsing sequence, is a metric cone. Hence splits to , and acts transitively on . Hence is .
In particular, there is such that is a -Reifenberg point. Then by Colding’s volume convergence ([21]), is -Reifenberg point for large, a contradiction. ∎
2. -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 be a sequence of closed Riemannian -manifolds with such that . Then admits a stratification for (some of them may be the same) such that
-
(2.1.1)
the Hausdorff dimension of is equal to .
-
(2.1.2)
is a smooth -manifold with a -Riemannian metric.
Moreover, if the -Hausdorff measure and the diameter of , then for any point that is -away from , the -harmonic radius at is no less than , where , .
Note that the original proof of [28, Theorem 0.9] depends on orthonormal frame bundles of , which do not generally admit a uniform sectional curvature bound such that only a -regularity can be derived on . 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 be a Riemannian manifold whose sectional curvature . Assume that there is a proper and free isometric action by on . Let
(2.1) |
where is defined when is less than the injectivity radius at , and is the parallel transport from to along the unique minimal geodesic.
Then the sectional curvature of the quotient at is bounded by
Proof.
Proof of Theorem 2.1.
Let us consider the smoothed metrics on , which are the solutions of Hamilton’s Ricci flow equation with the initial condition . By [60] (cf. [56]), also admits a uniform sectional curvature bound and a uniform higher regularities (1.4).
By passing to a subsequence, we assume that as for any fixed . Then is -bi-Lipschitz to , and admits a stratification for such that each strata is a smooth Riemannian -manifold with sectional curvature .
Indeed, for that approaches , let us consider the equivariant limit spaces of the -ball in the tangent space of with the pullback metric via the exponential map:
(2.2) |
where is the pseudo-group action by the local fundamental group of , and is equipped with its length metric. Then is a smooth Riemannian manifold with sectional curvature .
Since the pseudo-group acts on by isometries, by the standard theory of isometric actions on Riemannian manifolds (e.g., [3]) the orbit space admits a standard stratification by isotropy types. So is . Hence we derive (2.1.1).
For (2.1.2), it suffices to show that the sectional curvature at points in -definitely away from is bounded uniformly by a constant . Then (2.1.2) for the original limit space immediate follows from Cheeger-Gromov’s convergence theorem applied on as .
Let us argue by contradiction. Suppose there is a sequence of limit spaces such that , , and is arbitrary chosen, but the sectional curvature of at is unbounded as . Since is an Alexandrov space with curvature ,
By passing to a subsequence, let us consider the equivariant convergence of limit spaces in (2.2) for each .
(2.3) |
In order to apply Lemma 2.2, we first point out that the Lie group can be reduced to the case of free action.
Indeed, let be the identity component of . Instead of , we consider the actions by , 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 of lies in the center of . Hence the isotropy group of in a conjugacy class is unique on . It follows that the union of all orbits of the same isotropy type , , is the fixed-point set of , and hence is totally geodesic. Moreover, the isometric action on can be reduced to the quotient group that acts effectively and freely.
By our assumption, is not collapsing. For simplicity we assume that the action of itself is free.
We claim that
-
(2.2.1)
there is a sequence such that . By a suitable choice of , and maps the normal space of by a uniform and definite deviation away from the parallel transformation.
-
(2.2.2)
the normal space of has the same dimension as that of .
-
(2.2.3)
the limit of lies in the isotropy group at , whose differential admits a non-trivial transformation on the normal space of .
The claims above will yield a contradiction immediately. Indeed, let for each and in (2.2.1). Then forms a continuous path in which by (2.2.3) acts on non-trivially. By (2.2.2) and the slice theorem, has lower dimension than , a contradiction to that they are assumed to be non-collapsing.
By Lemma 2.2, the unboundedness on the upper curvature at implies that blows up as . Hence (2.2.1) holds by the definition (2.1) of .
Since converges to in the -topology, let us identify with via a suitable diffeomorphism. Then the isometric actions by one-parameter subgroups of -converges to that of on , which implies .
We point out that, though the manifolds with and -Reifenberg local covering geometry still can be smoothed to 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 generally blows up as . 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 -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
(3.1) |
Then the adapted coordinate chart at descends to a coordinate chart at such that , and the metric tensor on expressed in satisfy
-
(3.1.1)
,
-
(3.1.2)
, and
-
(3.1.3)
() is almost harmonic in the sense that
where is the gradient of with respect to , the -norm is taken in the coordinate chart , and is a function depending on such that as with fixed .
Proof.
Because () takes the same value on each fiber, is well-defined.
Let , where is the Levi-Civita connection on . From the definition of -harmonic coordinate chart, one has
(3.2) |
where and the -norm is taken in the coordinates .
Since is a Riemannian submersion, the gradient is a horizontal vector field on and
(3.3) |
where denotes the Levi-Civita connection on , and is the tangent map of . Define . Then
(3.4) |
It follows from (3.2) and (3.4) that satisfies (3.1.1), and thus is a coordinate chart at .
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
(3.5) | ||||
for any and . Then (3.1) together with (3.5) yields (3.1.2).
Secondly, by (3.3), we have
(3.6) |
where denotes the horizontal integral tensor of Riemannian submersion that takes values tangent to the fibers, and for any smooth vector field on , denotes its horizontal lifting on . Combing (3.3) with (3.6) and by the fact that is Riemannian submersion, we have
(3.7) |
Now (3.7) together with (3.1) yields (3.1.3). This complete the proof of Lemma 3.1. ∎
Remark 3.2.
Next, let us prove Theorem 0.11 by solving the Dirichlet problem with the boundary condition .
Proof of Theorem 0.11.
Let be a coordinate chart at provided by Lemma 3.1. Let be the origin. By a shift in value, we assume that . Let denote the image set of and the inverse map of . Let us pullback the metric on to by , where we still denote on by , and identify with by .
From (3.1.1) and by , we have the Euclidean ball . Let be the solution of the following Dirichlet problem:
Note that (3.1.1) yields that . We claim that
(3.8) |
yields the desired harmonic coordinate chart in Theorem 0.11, i.e., the following -estimates hold:
(3.9) |
where and -norm is taken in the coordinates .
4. -regularity on the limit space
From this section we begin to prove Theorem 0.1. We first show that any limit space is a -Riemannian manifold for any and . Without loss of generality, we assume that .
Let , where are Riemannian -manifolds with and -local covering geometry. By Lemma 1.12, we assume directly that is of -Reifenberg local covering geometry, where is the constant in (0.1).
As already pointed out in the introduction, it suffices to consider the equivariant Gromov-Hausdorff convergence:
(4.1) |
where the open balls and the normal covers are endowed with their length metrics, is the deck transformation by , and is the limit group of .
Let be the pullback Riemannian metric on the normal cover. By Remark 1.11, admits a uniform -harmonic radius at points definitely away from the boundary. Then by Anderson’s convergence theorem 1.6, the interior of is a smooth manifold, where gives rise to a -Riemannian metric tensor on .
Note that is the limit of with its length metric. By the fact that endowed with the restricted metric is isometric to , and the same holds for any open ball centered at an interior point whose radius , we derive that the interior of is isometric to equipped with the length metric.
Proposition 4.1.
There is such that
-
(4.1.1)
in (4.1) is a -smooth submersion that is Riemannian between -Riemannian metrics.
-
(4.1.2)
is a -smooth manifold with a -Riemannian metric.
We first verify that is regular, which is a direct corollary of the following observation.
Lemma 4.2.
Let be a closed subgroup of the isometry group of . Then any tangent cone of the quotient space is either isometric to an Euclidean space, or definitely -away from any Euclidean space in the pointed Gromov-Hausdorff distance.
Proof.
Let be the orbit of at the origin of , and the isotropy group at . Let be the normal space of at , and the quotient point of in . By the standard theory of isometric group actions (e.g. see [35, Proposition 1.8]), the tangent cone at is isometric to . Hence it is an Euclidean space if and only if the action of on is trivial.
In the following we assume acts on non-trivially.
Case 1. the action of on is discrete. Then is an Euclidean cone over , whose volume is no more than half of . Hence is definitely away from .
Case 2. the identity component of acts on non-trivially. Since the orbit of on must contain a great circle, the radius of is no more than . Hence is also definitely away from any Euclidean space. ∎
Lemma 4.3.
There is such that for any , if satisfies the -Reifenberg condition, then it is regular.
Proof.
Next, we prove the key lemma in this section. Let us fix the pointed Gromov-Hausdorff approximation in (4.1).
Lemma 4.4.
For any , the action of on in (4.1) is smooth and proper, such that the open ball is a -Riemannian manifold.
Proof.
Let us apply the Ricci flow on . By Theorem 1.8, the solution of Ricci flow (1.3) with initial condion exists for , a positive constant depending only on and , such that (1.4) holds for .
For any fixed , let us view as a sequence of metrics on in (4.1). Then by the Arzela-Ascoli Theorem under Gromov-Hausdorff convergence, the underlying distance functions subconverges to a distance function , which is -bi-Lipschitz to the underlying distance of . By (1.4) and Anderson’s -convergence theorem, up to a subsequence uniformly converges to a smooth metric tensor on with the same regularity as in (1.4). Thus, we have
(4.2) |
where , is a smooth Riemannian manifold, and is the limit group action of . (Furthermore, if is replaced by a local diffeomorphism that realizes the -convergence of to , then it can be seen that -converges to . We do not need this fact here.)
The key point here is that the action of on and is the same as deck transformation of the normal cover . And at the same time, and , and hence and , are two metric tensors on the same smooth manifolds and respectively, which by (1.4) are -bi-Lipschitz equivalent.
It follows from the Arzela-Ascoli Theorem that the limit action of under is also the same as under . Since the action of is smooth and proper on , so is on . Thus, we derive that coincides with , and can be identified to as topological spaces.
In order to show that is a -Riemannian manifold, we first prove that is a smooth Riemannian manifold.
Indeed, the tangent cone of is the same as that in , which by the -Reifenberg condition, is -close to be Euclidean on the unit ball at the vertex point. By the -bi-Lipschitz equivalence between and , the tangent cone of is -close to be Euclidean on the unit ball. By Lemma 4.2, for , is regular.
Then by the standard theory of isometric actions on Riemannian manifolds (e.g., [3]), the slice representation of the isotropy group at any interior point is trivial, hence the orbit is principle. By the principal orbit theorem (e.g. see [35, §1]), it follows that is a smooth Riemannian manifold, and is a smooth Riemannian submersion.
Let us endow with the smooth structure on . Because coincides with , is also a submersion. By the fact that acts on isometrically, the Riemannian metric tensor on induces a quotient Riemannian metric on .
Furthermore, by the implicit function theorem, an adapted coordinate chart can always be constructed around preimages of a point in . By the proof of Lemma 3.1, the adapted coordinate charts on descend to -admissible local coordinate charts on , where the quotient metric tensor is . ∎
Proof of Proposition 4.1.
Let be that in Lemma 4.3 and . Then (4.1.1) has been proved in Lemma 4.4. In the following we prove (4.1.2). That is, admits a -smooth differentiable structure, such that the metric tensor induced locally from in (4.1) is .
It suffices to show the local charts induced from in (4.1) are -admissible with each other, where the metric tensors coincide with each other by pulling back.
Indeed, let and be two open balls, whose intersection is non-empty. Then the identity map from with its length metric to with the restricted metric is -Lipschitz and locally isometric. By the precompactness Theorem 1.10, the normal cover sub-converges to as . And the identity map sub-converges to , which is also locally isometric.
Furthermore, by (1.5), is a normal cover for . Let be the limit submersion of as in (4.1). Then we derive the following commutative diagram
where the maps are locally isometric.
By the proof of Lemma 4.4, is also a smooth manifold and is a smooth submersion. By the commutative diagram above, local coordinate charts on from coincide with those on descending from . Hence they are -admissible with each other. Moreover, the metric tensors on are also the same up to the diffeomorphism . ∎
Remark 4.5.
Let , where are Riemannian -manifolds with -local covering geometry, and 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 with initial metric admits a uniform positive existence time , and the higher-ordered regularities (1.4). Assume that sub-converges to . We point out that, by the proof of Lemma 4.4, for not only satisfying , but also all , is regular. Hence, by the proof of Proposition 4.1, is a smooth Riemannian manifold for any .
Indeed, this can be seen from the fact that shares the same isotropy group as for , or an open and closed argument on , due to that implies that the tangent cone of is isometric to , and hence is extended further to cover .
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 and on its limit space in the graph (4.1) respectively. Let be the Riemannian metric tensor on and its quotient metric tensor on .
Proposition 5.1.
There are constants and such that for and , there exists an adapted harmonic coordinate chart defined on such that () descends to a smooth function in , and admits the following regularities:
(5.1) |
and
(5.2) |
where the -norm is taken in the coordinates .
Proof of Theorem 0.1.
Let be that in Proposition 5.1, and a compact Ricci-limit space of Riemannian -manifolds with and -local covering geometry, such that any point is -Reifenberg. By Lemma 1.12, is of -Reifenberg local covering geometry for .
Let be as in Proposition 5.1, and let . By Proposition 4.1, is a -smooth manifold with a -Riemannian metric tensor . Proposition 5.1 together with Theorem 0.11 implies the harmonic radius of is no less than . The first part of (0.1.1) is complete.
For (0.1.2), let us assume that . Since is a Ricci-limit space, there are harmonic coordinate charts that covers whose number admits a uniform bound , 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], is compact in the -topology.
What remains is to show (0.1.1.a-c).
Let , where are Riemannian -manifolds with and -Reifenberg local covering geometry. Let us consider the Ricci flow solution with initial metric , which exists for by Theorem 1.8 (Dai-Wei-Ye [23], cf. [17]) and satisfies the higher-ordered regularities (1.4). By Remark 4.5, for all the limit space of is a smooth Riemannian manifold , which is -bi-Lipschitz equivalent to .
By Remark 4.5 again, there is such that lies in for . Hence it admits a uniform harmonic radius for all . By the -precompactness, there is a diffeomorphism from to such that the pullback metric converges to in the -norm.
Let be the pullback metric on the normal cover of in the graph (4.2). Then the regularities (1.4) pass to the limit metric on . By the proof of Lemma 3.1, the relations (3.4)-(3.5) between the quotient metric and imply that satisfies the same regularities (1.4) up to a definite ratio depending on and . ∎
5.1. Preparation
Let us make some preparation for the proof of Proposition 5.1.
Lemma 5.2.
Let be a harmonic coordinate chart at with such that
(5.3) |
and
(5.4) |
where the -norm is taken in the coordinates . Then after blowing up by , where , the harmonic coordinates -converge to a Cartesian coordinate system in , and -converge to an Euclidean metric . That is, for
(5.5) |
and
(5.6) |
where as with the other variables fixed.
Proof.
This is an elementary fact. Let us identify and with the pullback metric via the inverse of . Then each is a harmonic function on . By a linear transformation, we can assume that . Put and . Then , and by the rescaling property of the -harmonic radius in Remark 1.2,
(5.7) |
where the -norm is taken in coordinates .
Therefore, by (5.7) and the standard Schauder interior estimates, -converge to a standard Cartesian coordinate system in , and the metric -converge to the standard Euclidean metric. ∎
Let us assume that is -Reifenberg. Let and that converges to . Let us consider the Gromov-Hausdorff convergence associated to the normal cover as in the diagram (4.1).
By Lemma 1.12 and [4], there exists such that for given any , and , the -harmonic radius of are uniformly bounded below by some constant . Assume that converges to . By the continuity of -harmonic radius in Proposition 1.5, so is .
After blowing up by , it follows from the definition of -Reifenberg and Lemma 5.2 that
(5.8) |
where is the restriction of in (4.1) on the local balls, , and . After passing to a subsequence, as converge to a submetry , which is the canonical projection from to .
Up to a suitable diffeomorphism, we view as a fixed domain in with metric . Let be a sequence of harmonic coordinates -converges to a limit harmonic coordinates as . Let . By Lemma 5.2 again, each
(5.9) |
is uniformly -close to a Cartesian coordinates on as . By passing to a subsequence of , we may assume that converges to a Cartesian coordinates with the following regularities: For any fixed ,
(5.10) |
and
(5.11) |
where is the limit Cartesian coordinates system of as , is the metric matrix of in coordinates , and the norm is taken in the Euclidean coordinates .
Furthermore, up to composing an orthonormal transformation on (also on ), we assume that the limit projection in the diagram (5.8)
(5.12) |
gives rise to a Cartesian coordiates on .
5.2. Construction of the adapted harmonic coordinates
Based on the preparation above, let us begin the construction of the adapted harmonic coordinates on and their limit .
By the definition of -Reifenberg, for all sufficiently large we have the Gromov-Hausdorff distance
Let be an -Gromov-Hausdorff approximation. Let be the orthonormal basis at the origin of associated to the Cartesian coordinates in (5.12).
Let . Then for each , there exists pairwise -close to for all large . Let
(5.13) |
where denotes the distance induced by on . Let be the solution of the following Dirichlet problem for ,
(5.14) |
where is the Laplace-Beltrami operator associated with metric . , forms a -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 such that for any and , the following -estimate holds for all sufficiently large :
(5.15) |
As is locally isometric, is also harmonic, i.e.,
(5.16) |
In the following we show that still forms a harmonic coordinates. Since as , there exists a sequence of diffeomorphisms with , such that the pullback metrics (sub-)converge to in the -norm. From now on, in the following subsections let us identify with via .
Lemma 5.3.
The map still forms a harmonic coordinates that converges to the Cartesian coordinates given by (5.12) in the -norm as and .
Proof.
It suffices to show that is -close to for every .
By (5.10), is close to the Cartesian coordinates . Moreover, by the assumption under (5.9), converges to the Cartesian coordinates , which projects to the Cartesian coordinates .
At the same time, let be the Buseman-typed function defined in (5.13) in the construction of . Then by the choice of , for any fixed ,
(5.17) |
where is the Gromov-Hausdorff approximation.
By the construction above, converges to as and . Moreover, by (5.15) the difference between and goes to zero as and . Let be the Cartesian coordinates on in (5.8) such that . Then, by (5.12) and the triangle inequality below
we derive
Since is also harmonic, by the Schauder interior estimates,
(5.18) | ||||
where the -norm is taken in coordinates .
Now together with (5.10) and (5.11), (5.18) implies that
(5.19) |
still forms a harmonic coordinates that converges to the Cartesian coordinates in the -norm as and , which satisfies
(5.20) |
and
(5.21) |
∎
Now we are ready to prove Proposition 5.1.
Proof of Proposition 5.1.
Let be the harmonic coordinate chart constructed in Lemma 5.3. Since -converges to as , by (5.20), -converges to a limit harmonic coordinates as .
For each , by the construction of , its limits takes the same value along every -fiber, and thus it naturally descends to a -smooth function on .
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).
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 be the constant in Theorem 0.1, and . For (0.3.1), let us first prove that is a -Riemannian manifold with a positive harmonic radius.
Indeed, let , where have and -Reifenberg local covering geometry. For any , let us consider the equivariant convergence of normal covers of -balls in (4.1). Since every tangent cone at is -close to , and the proofs of Propositions 4.1 and 5.1 still go through for . That is, is a smooth manifold and for any , the -splitting map on defined by (5.14) from the Buseman functions by the closeness between and gives rise to an adapted harmonic coordinate chart that descends from in (4.1), which satisfies the uniform regularities (5.1) and (5.2). Then by Theorem 0.11, admits a -harmonic coordinate around any point . The continuity of -harmonic radius at points in yields a positive lower bound of the harmonic radius of .
Next, we show that the -harmonic radius at a point satisfying
(6.1) |
admits a uniform bound .
Let us argue by contradiction. Assume that there is a sequence of spaces , each of which contains a point satisfying (6.1), but the -harmonic radius . By passing to a subsequence we assume that .
Let , where has bounded Ricci curvature and -Reifenberg local covering geometry. By Theorem 1.8, there is such that any admits a smoothed metric with uniformly higher regularities (1.4). By passing to a diagonal subsequence, we assume that
Since is regular and is -bi-Lipschitz to , by Lemma 4.2, each is also regular. Hence by Proposition 4.1 is a smooth Riemannian manifold . Moreover, by O’Neill’s formula applied on the Riemannian submersion in (4.2), the sectional curvature of is bounded below uniformly by .
Since the volume of is bounded below by , their limit is an Alexandrov space with curvature of Hausdorff dimension .
By the proof of Theorem 2.1, the sectional curvature of is also bounded uniformly from above. By [16, Theorem 4.7] the injectivity radius of is bounded below by . By Cheeger-Gromov’s -precompactness, is regular. Then by the -bi-Lipschitz equivalence between and the original limit and Lemma 4.2 again, is also regular.
Therefore, by the first part of (0.3.1) is a -Riemannian manifold that admits a positive harmonic radius . In particular, for any , is -close to . Then is also -close to an Euclidean space for fixed and any large . Then harmonic coordinate charts of a definite radius can be constructed at by the proof of Theorem 0.1. It contradicts to that .
For limit spaces under bounded local covering geometry, we have the following -regularity that depends on the space itself.
Let be the set consisting of all compact Ricci-limit spaces of Riemannian -manifolds with and -local covering geometry, such that each element is -almost regular.
Corollary 6.1.
Let , where are Riemannian -manifolds with and -local covering geometry. Then the followings hold for .
-
(6.1.1)
For any sufficient large and any , the preimages of in the universal cover of admit a uniform -harmonic radius .
-
(6.1.2)
Any element is regular in the sense that for any , any tangent cone at is isometric to .
Proof.
Suppose that converges to , and converges to a tangent cone that is -close to , where . Let us consider the equivariant pointed Gromov-Hausdorff convergence
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 , where is the radius appearing in the Reifenberg condition and bounded local covering geometry.
Indeed, let be a closed Riemannian -manifold with bounded Ricci curvature and -Reifenberg local covering geometry, and a closed -Riemannian -manifold in for and , such that
If , up to a rescaling on the metric and , is of -Reifenberg local covering geometry and with
By viewing and as the rescaled metrics and respectively, we have and
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 be a closed Riemannian -manifold with and -Reifenberg local covering geometry, and . If , then there is a -smooth fibration that satisfies
-
(7.1.1)
is a -almost Riemannian submersion, where after fixing , as ;
-
(7.1.2)
the fiber has intrinsic diameter for any ;
-
(7.1.3)
the second fundamental form of satisfies ;
-
(7.1.4)
is diffeomorphic to an infra-nilmanifold.
Proof of Proposition 7.1.
Let be an -Gromov-Hausdorff approximation (for simplicity, -GHA) with . Without loss of generality, we assume .
Step 1. First, we prove that for , , and any , there is a local fibration , which is a -almost Riemannian submersion with a uniform second derivative control
(7.1) |
and satisfies
(7.2) |
Indeed, by the assumptions is of -Reifenberg local covering geometry. Let us consider the equivariant closeness of the normal cover in the diagram (7.3):
(7.3) |
where is a -GHA, whose restriction on coincides with .
As the arguments before Lemma 5.3 in Section 5.2, in the following we identify with an open domain with the the pullback metric via a diffeomorphism , such that and is -close to .
Let and . Since by Theorem 0.1 the -harmonic radius of is no less than , the rescaled has -harmonic radius at least . By Lemma 5.2, is -close to an Euclidean ball .
Let us consider the harmonic -splitting map constructed in (5.14),
(7.4) |
Since the Ricci curvature of satisfies . By the triangle inequality, is -close to . Then there is such that for , (5.15) holds for and the Buseman functions as in (5.14).
In the following we fix . By the proof of Proposition 5.1 applied on (7.3), the lifted harmonic functions on together with other harmonic functions form a -harmonic coordinate chart , whose coordinate functions admit a uniform -norm bound .
By the -compactness for , is --close to an adapted harmonic coordinate chart for , which by Lemma 3.1 descends to an almost harmonic coordinate chart on , whose coordinate functions are .
By (7.6), is -close to , which is a Riemannian submersion. Hence is an -Riemannian submersion.
By and the uniform bound on up to the 2nd covariant derivative in Lemma 3.1, 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 -GHA .
Let us consider the nearby metric provided by Theorem 0.1 for fixed . For , let and . By (0.1.1) for , , and the -harmonic radius of is at least . It follows that the injectivity radius of admits a lower bound . Hence the convexity radius of is no less than . Without loss of generality, we assume that .
In the following we view the rescaled metric as a rescaling of , where is provided by Step 1. Let be a -net in with depends only on the dimension of . For each , let be the local fibrations constructed in Step 1. Since is -bi-Lipschitz to , (7.2) implies that
(7.7) |
with respect to .
By taking such that , we will glue such local fibrations together with respect to the smoothed metric to a -Riemannian submersion with respect to the rescaled original metric .
Let be a smooth cut-off function such that , , and . Let . Let us consider the energy function
By the construction of and (7.7), for , is a strictly convex function in , and it takes a unique minimum point, , that is -close to measured in . We define
Let us prove (7.1.1)-(7.1.3) first. By its definition is determined by the equations
where is the gradient of with respect to . Note that, in the normal coordinates at , can be written in the form
where is the position vector of in the normal coordinates of . Then the differential of is determined by , where
By the sectional curvature bound of provided by (0.1.1.b), is -close to the identity matrix , which is the Hessian of squared Euclidean distance. At the same time, by the Bishop-Gromov’s relative volume comparison, the count of with non-vanishing can be chosen at most . Hence is also -close to . It follows that for all sufficient small ,
(7.8) |
Since by (7.7) and the choice of , is -close to whenever is well-defined,
(7.9) |
Combing (7.8) and (7.7) together, we derive
(7.10) |
In order to show that is -almost Riemannian submersion, it suffices to show that the local fibrations nearby are -close to each other.
Indeed, by the definition of in Step 1, is a harmonic map, which by (7.7) is -close to each other up to a transformation in the intersection of their domains. In particular, for
(7.11) |
At the same time, by the construction of (see also Proposition 5.1 and Lemma 3.1), on is - close to a constant isometric transformation on .
Now by Cheng-Yau’s gradient estimate [20] for the component harmonic functions of
in the context of uniform lower Ricci curvature bound, it follows that, for and ,
(7.12) |
(Note that the support of lies in .)
Since is a -Riemannian submersion, by -closeness (7.12) for , the average of in (7.10) is a -Riemannian submersion. Then (7.10) implies (7.1.1).
The uniform bound on the 2nd fundamental form of 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 and of .
Indeed, the 2nd fundamental form of with respect to the Euclidean metric in the normal coordinates from can be expressed in matrix by
(7.13) |
where consists of the Hessian of each components of with respect to such that
(7.14) |
For the last term of (7.13), we note that is a combination of , which is -close to the identity matrix . So is its inverse . For and , we have
Observe that, in the normal coordinates at . It follows that admits a uniform upper norm bound with respect to and . So are and .
At the same time, by (7.9) and (7.1) for , each term in (7.14) is bounded by . It follows that with respect to the rescaled metric and .
After rescaling back, we derive the bound of 2nd fundamental form in (7.1.3) with respect to . By the -compactness in Theorem 0.1, it can be seen that after replacing with the original metric , (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 and injectivity radius of is , where . Suppose for some , the intrinsic diameter
where . By the second fundamental form bound in (7.1.3), the extrinsic diameter of other fibers over is no less than . By (7.1.1), at least many of -balls are required to cover . However, by the existence of -GHA , which is -close to , and the harmonic coordinates on , at most such balls are required. Hence .
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, can be smoothed to by the Ricci flow [23], whose sectional curvature is uniformly bounded by . By the -compactness, lifted to the universal cover of -balls on , the -norm between and is uniformly bounded by . So is the Levi-Civita connections of and . By the expression of the second fundamental form in terms of Christoffel symbols, it is easy to see that with respect to still satisfies (7.1.3). Hence, with the induced metric by admits a uniform sectional curvature bound with a small diameter . The Gromov’s almost flat theorem [33] (cf. also [59, 58]) implies . ∎
Proof of Theorem 0.8.
The only difference from Proposition 7.1 is the -smoothness of the fibration . Note that the -regularity of in Proposition 7.1 is due to the limit coordinate chart on in the definition of ; see the paragraph below (7.6).
Remark 7.2.
We point out that the local fibration with fixed are modeled on the Riemannian submersion 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 , one picks up for each a local “-splitting map” modeled on the Euclidean space as that in (5.14), then by Cheeger-Colding’s well-known -estimates on their gradients and Hessians [11, 10], it also yields a -almost Riemannian submersion for (cf. [50, Proposition 6.6]), such that
(7.15) |
and
(7.16) |
But one immediately encounters the following issues:
-
(7.2.1)
to guarantee the local fibrations and the global fibration glued together are -almost Riemannian submersions, has to approach ;
-
(7.2.2)
after the local fibrations are re-defined locally for each , the 2nd derivative (7.16) after rescaling back blows up as .
Without a solution of the above, only a rescaling invariant 2nd derivative control on a -almost Riemannian submersion can be derive, such as
(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 with and -local covering geometry can be smoothed to a nearby metric that admits a uniform bounded sectional curvature depending on . Hence by Theorem 0.6 a fibration exists such that (0.6.1-4) hold with respect to .
In order to remove the perturbing error that arises from , such that the fibration remains to be a -Gromov-Hausdorff approximation as , must go to zero. Moreover, since the sectional curvature of generally blows up as , some explicit curvature control on (e.g., in [40, Theorem 1.6] by Perelman’s pseudo-locality [52]) and an arbitrary small distance distortion (e.g., 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 by the Ricci flow method and by suitably choosing the flow time with respect to , 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 . 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. -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 -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 -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.