RCD(0,N)-spaces with small linear diameter growth
Abstract
In this paper, we study some structure properties on the fundamental group of RCD() spaces. Our main result generalizes earlier work of Sormani [30] on Riemannian manifolds with nonnegative Ricci curvature and small linear diameter growth. We prove that the fundamental group is finitely generated if assuming small linear diameter growth on RCD() spaces.
Keywords: RCD(), finitely generated, fundamental group
1 Introduction
In recent years, the theory of RCD() spaces has a remarkable development. After Lott-Villani [24] and Sturm [33, 34] introduced the curvature dimension condition CD() independently, the notion of RCD() spaces was proposed and analyzed in [15, 4], as a finite dimensional counterpart of RCD() which was first introduced in [3]. Roughly speaking, a CD() space is a metric measure spaces with Ricci curvature bounded from below by and dimension bounded from above by , and RCD() spaces are those infinitesimally Hilbertian CD() spaces. One of the disadvantages of CD() for finite is the lack of a local-to-global property. To this aim, Bacher and Sturm [6] introduced the reduced curvature dimension condition CD∗() and as before, RCD∗() spaces are infinitesimally Hilbertian CD∗() spaces.
The motivation of studying RCD() (resp. RCD∗()) spaces is to single out the “Riemannian” class in CD() (resp. CD∗()) spaces, which excludes Finsler manifolds. Therefore, it is natural to expect that some analytical and topological properties of Riemannian manifolds also hold on RCD() or RCD∗() spaces.
It is well-known that for Riemannian manifolds, the Ricci curvature controls the fundamental group very well. In this area, one famous open problem in the past was that an open manifold with nonnegative Ricci curvature has a finitely generated fundamental group, conjectured by Milnor [26] in 1968. This conjecture was of great interests until Bruè-Naber-Semola [8] constructed a counterexample with such that . However, it is still worth studying under what conditions Milnor Conjecture holds. Some results in this direction have been accomplished by Anderson [5], Li [23] and Sormani [30] among others. Anderson and Li proved that if a manifold with has Euclidean volume growth, then the fundamental group is finite and this result has been extended to an RCD() space by Mondino-Wei [28]. A key technical point to get information on the fundamental group is to study the universal cover and the group of deck transformations, denoted by , which is a quotient of the fundamental group and called revised fundamental group in [31].
Moreover, in a more recent paper [37], Wang proved that any RCD() space is semi-locally simply connected, which implies that the universal cover is simply connected and is isomorphic to . Hence, the structure properties derived by Mondino-Wei [28] on hold on the fundamental group . We will review this point in the next section.
In this paper we extend the result of Sormani [30] to RCD() spaces and the main theorem is stated as following.
Theorem 1.1.
Let be an RCD(0,N) space for some . Then,
-
(1)
If , and X has small linear diameter growth, i.e.,
where
then the fundamental group is finitely generated.
-
(2)
If , then the fundamental group is finitely generated.
-
(3)
If , then the fundamental group is trivial.
To prove this result, we extend the Halfway Lemma and Uniform Cut Lemma established by Sormani in [30] to a non-smooth setting and the spirit of the proof is similar. Let us point out that Kitabeppu and Lakzian proved a similar result in [21], under additional non-branching and semi-locally simple connectedness assumptions. In our argument, we drop the non-branching assumption by carefully applying the excess estimate established for RCD() spaces in [17]. It can be seen in the proof that the excess estimate can prevent the structure of an RCD() space being too wild. Also, the semi-locally simple connectedness is necessary and sufficient for the existence of a simply connected universal cover, but by Wang’s result [37], we can directly drop this assumption.
Furthermore, we prove the finite generation of without assuming small linear diameter growth in the case , which differs from Kitabeppu and Lakzian’s result in [21]. Our results are more similar to the situation on manifolds. Notice that for 2-dimensional manifolds, is equivalent to , which implies that the the fundamental group is finitely generated (see [18]).
Also, since the Ricci-limit space is RCD space, it is straightforward to obtain the following corallary.
Corollary 1.2.
Let be a sequence of n-dimensional Riemannian manifold with and . Then,
-
(1)
If , and X has small linear diameter growth, i.e.,
where , then the fundamental group is finitely generated.
-
(2)
If , then the fundamental group is finitely generated.
Moreover, if an RCD() space has linear volume growth, then by Huang [20], its diameter growth is sublinear, which was proved on manifolds by Sormani [29]. Thus, we get the following corollary.
Corollary 1.3.
Let be an RCD(0,N) space with for some positive constant and . Then the fundamental group is finitely generated.
The paper is organized as follows. In section 2, we recall the definition of lower Ricci curvature bounds on metric measure spaces and review some basic properties and useful results. In particular, we will discuss Mondino and Wei’s work on the universal cover of RCD() spaces [28]. In section 3, we generalize Sormani’s technical lemmas to RCD() spaces, and the proof of Theorem 1.1, Corollary 1.2 and 1.3 are presented in section 4.
2 Preliminaries
Throughout this paper, is a metric measure space (mms. for short) where is a complete and separable geodesic metric space and is a locally finite nonnegative Borel measure with supp. We also assume is not a point.
2.1 RCD∗() spaces
In this subsection, we recall some basic definitions and properties of metric measure spaces with lower Ricci curvature bounds.
We denote by the set of Borel probability measures on and by the subset of probability measures with finite second moment, i.e.
For , the -distance is defined by
(2.1) |
where the infimum is taken over all such that .
Note that is a geodesic space provided that is a geodesic space. Then we define the evaluation map as
We will denote by Geo() () the space of (constant speed minimizing) geodesics on endowed with the sup distance.
Given , let OptGeo() be the space of all (Geo()) for which is a minimizer in (2.1) and recall that is in Geo() if and only if there exists OptGeo() such that for all .
Now, let us introduce the so-called reduced curvature dimension condition CD∗(), coming from [6]. For with , and , we set
(2.2) |
and let
(2.3) |
for all .
Definition 2.1 (Reduced curvature dimension condition).
Let and . We say that a mms. is a CD∗() space if for any two measures with bounded support, there exists a measure OptGeo() such that for any and , we have
(2.4) |
where we have written with , for all .
Remark 2.2.
On a CD∗() space , a natural version of Bishop-Gromov volume comparison holds (see [6] for precise statement) and this leads to the properness of (i.e. closed bounded sets in are compact).
Remark 2.3.
The original curvature dimension condition, denoted by CD(), has the same definition as CD∗() except that the coefficients and in (2.4) are replaced by and respectively. In general, CD∗() is weaker than CD(), while in the case which we will mainly consider in this paper, these two notions are identical.
It is possible to see that Finsler manifolds are allowed as CD∗() spaces. In order to single out the “Riemannian class”, the CD∗() condition might be strengthened by requiring additionally that the Sobolev space is Hilbert, inspired by the fact that a smooth Finsler manifold is Riemannian if and only if the space is Hilbert. We briefly review the definition of Sobolev space for mms. below.
Recall that the Cheeger energy is defined through
(2.5) |
where, . Then, we define
(2.6) |
By looking at the optimal approximating sequence in (2.5), one can identify a canonical object , called minimal relaxed gradient, which provides the integral presentation
See [2] for more details on this topic.
The Sobolev space endowed with the norm is a Banach space, but it is not Hilbert in general. If is Hilbert, then we say that is infinitesimally Hilbertian.
Definition 2.4.
An RCD∗() (resp. RCD()) space is an infinitesimally Hilbertian CD∗() (resp. CD()) space.
A basic property of RCD∗() spaces is the stability under pointed measured Gromov-Hausdorff (pmGH for short) convergence. See [16, 13] for a proof.
Proposition 2.5 (Stability).
Let . If is a sequence of RCD∗() spaces with , then is also an RCD∗() space.
Finally, we state a few properties of RCD() spaces (the first one is proved in [17], the second in [14], the third in [20] and the fourth in [19]). Note that RCD() condition is exactly the same as RCD∗() condition.
Theorem 2.6 (Abresch-Gromoll excess estimate).
Let be an RCD(0,N) space with . Fix and a minimizing geodesic joining them. Define
Then for any with , we have
(2.7) |
where and is the excess function w.r.t p and q.
Note that in [17], Gigli and Mosconi proved excess estimates for all RCD() spaces with , but for our purposes, we only need the case here.
Theorem 2.7 (Splitting).
Let be an RCD(0,N) space with . If contains a line, then is isomorphic to , where is the Euclidean metric, is the Lebesgue measure and is an RCD(0,N-1) space if and a singleton if .
Theorem 2.8.
Let be an RCD(0,N) space with for some positive constant and . If does not split, then
Theorem 2.9.
Let be a noncompact RCD(0,N) space with , then for every , there exists a constant such that
2.2 The topology on metric measure spaces
We first recall the definition of the universal cover of a metric space [32].
Definition 2.10 (Universal cover of a metric space).
Let be a connected metric space. We say that a connected metric space is a universal cover of if is a cover for with the following property: for any cover of , there is a commutative diagram formed by a continuous map and the two covering projections onto :
Remark 2.11.
The universal cover may not exist in general. However, if it exists then it is unique. Moreover, if a space is locally path connected and semi-locally simply connected (i.e., for all , there is a neighborhood such that any loop in is contractible in ), then it has a simply connected universal cover. On the other hand, the universal cover of a locally path connected space may not be simply connected (See [32]).
If a space has a universal cover, then one can consider the revised fundamental group introduced in [31]. We first recall that for a covering , a deck transformation is a homeomorphism such that and all deck transformations form a group .
Definition 2.12 (Revised fundamental group).
is a metric space which admits a universal cover . Then the revised fundamental group of , denoted by , is the group of deck transformations .
Notice that can be seen as a quotient of and the trivial element in is represented by those loops in based at , which are still loops when lifted to . Thus if the universal cover is simply connected, then is isomorphic to . Before Wang [37] proved the semi-locally simply connectedness for spaces, the existence of the universal cover of an RCD∗() space was confirmed by Mondino and Wei [28].
Theorem 2.13.
Let be an RCD space with . Then admits a universal cover which is also an RCD space.
Remark 2.14.
By Mondino and Wei’s contruction, naturally inherits the length structure of the base and is locally isometric to . Then by completeness and locally compactness, is a geodesic space. In our argument below, we will always assume the universal cover to be geodesic.
Mondino and Wei [28] also derived several structure properties on the revised fundamental group of an RCD∗() space in the similar manner as Riemannian geometry. We present one of them, which is an extension of the celebrated result by Cheeger-Gromoll [12] for compact manifolds with nonnegative Ricci curvature.
Theorem 2.15.
Let be a compact RCD(0,N) space with . Then the revised fundamental group contains a finite normal subgroup such that contains a subgroup of finite index.
Remark 2.16.
Finally, we point out that it has been proved by Wang in [37] that any space is semi-locally simply connected, which generalizes the same author’s result in [36].
Theorem 2.17.
Let be an space with . Then is semi-locally simply connected. In particular, the universal cover is simply connected and is isomorphic to .
By Theorem 2.17, in the RCD setting, we get rid of the notion of revised fundamental group and the revised fundamental group in Theorem 2.15 can be replaced by . In the remaining part of this paper, we still use the notation to denote the deck transformation group, which acts on discretely. Note that in this paper, we typically consider RCD spaces and is isomorphic to in most of our context.
2.3 Structure of RCD() spaces
The main purpose of this subsection is to provide some metric measure structure theory of RCD spaces, which we will need in the proof of Theorem 1.1 for the case . We use the notion RCD() instead of RCD∗() in this subsection, since most previous works reviewed in this subsection selected this stronger notion, though it seems that all these results hold on RCD∗() spaces.
Given an RCD() space with , we first recall the notion of tangent cones.
Definition 2.18 (tangent cones).
We say that a pointed metric measure space is a tangent cone of at if there exists a sequence such that
The collection of all tangent cones of at is denoted by .
A compactness result on RCD() spaces yields that is non-empty for any (see Chapter 27 in [35] for instance). We are now in the position to introduce the notions of -regular set and essential dimension as follows.
Definition 2.19 (-regular set).
For any integer , we denote by the set of all points such that , where is the volume of the unit ball in . We call the -regular set of .
The following result is proved by Bruè-Semola in [9].
Theorem 2.20.
Let be an RCD space with . Then there exists a unique integer , called the essential dimension of , denoted by , such that .
When the essential dimension reaches its maximum value , Brena-Gigli-Honda-Zhu obtain the following result (see Theorem 1.3 and Theorem 2.20 in [7]).
Theorem 2.21.
If an RCD space satisfies , then for some constant . In particular, is an RCD space.
Finally, let us recall that in dimension 2, the synthetic notions of lower bounds on sectional and Ricci curvature coincide (see [25]).
Theorem 2.22.
If is an RCD space, then is an Alexandrov space with curvature .
3 Halfway Lemma and Uniform Cut Lemma on RCD(0,N) Spaces
In this section, we extend two technical lemmas given by Sormani in [30] to a non-smooth context. First of all, we recall a notion introduced by Sormani [30].
Definition 3.1.
Let be a geodesic metric space which admits a universal cover . Given , we say that is a minimal representative geodesic loop of if , where is a minimal geodesic from to .
Lemma 3.2 (Halfway Lemma).
Let (X,d) be a proper geodesic metric space. Assume that (X,d) admits a universal cover (,). Then there exists an ordered set of independent generators of G(X) with minimal representative geodesic loops of length such that
If G(X) is infinitely generated, then we obtain a sequence of such generators.
Proof.
Fix and let be a lift of . Since is proper and acts discretely on , there exists a non-trivial element such that
Let . Define each and inductively by
Notice that is nonempty for all if is infinitely generated. Let be a unit speed minimal geodesic from to . Define (i.e., is the minimal representative geodesic loop of based at ).
It only remains to prove:
Suppose that there is a such that . Then there exists such that . Thus we can find a minimal geodesic from to with length .
Denote and . Then
and
Therefore, . But , which is a contradiction since . ∎
Lemma 3.3 (Uniform Cut Lemma).
Let (X,d,m) be a RCD(0,N) space with . Let be a geodesic loop based at with and is nontrivial. Suppose satisfies the following two conditions:
-
(1)
If based at is a loop such that in , then
-
(2)
is minimal on and .
Then there is a universal constant defined in Theorem 1.1, such that for any with ,
Proof.
We first prove this lemma for and argue by contradiction. Suppose that there exists a point such that
Let be a unit speed minimal geodesic from to . Let be the universal cover of and be the lift of . By Theorem 2.13, is an RCD() space.
Denote . Then is a minimal geodesic from to by the first condition of . Thus . We can also lift the curve to where runs from to . Note that and
The excess of w.r.t and satisfies
We can now apply the excess estimate (Theorem 2.6).
Notice that
(3.1) |
where and
(3.2) | ||||
(3.3) |
Combining (3.1)-(3.3), we get
and
This contradicts with the definition of ().
For and , let be a point on a minimal geodesic from to . Then
Thus we complete the proof. ∎
4 Proof of Main Theorems
Proof of Theorem 1.1.
We first point out that there is a complete classification for RCD∗() spaces when . Indeed, is isometric to or (see Corollary 1.2 in [22]). Thus we only need to consider the case .
(1) .
Suppose is infinitely generated. Construct a sequence of independent generators , as in Lemma 3.2, with minimal representative geodesic loops based at some point . Notice that satisfies the hypothesis in Lemma 3.3. Let be the universal cover of .
We observe that diverges to infinity, since otherwise the orbit would be contained in a closed ball . Since is compact and acts discretely on , must be finite which is a contradiction.
Choose a sequence , then by Lemma 3.3,
There exists a point on the minimal geodesic from to and satisfies
Then,
which is a contradiction.
(2) .
In this case, 1 or 2. If , then and is isometric to or (see Theorem 1.1 in [22]). Thus, we may assume . By Theorem 2.21 and Theorem 2.22, we know that is a 2-dimensional Alexandrov space with curvature . Then we may go through Gromov’s arguments in [18].
Let (which is also an Alexandrov space with curvature ) be the universal cover of . Notice that acts on via isometries. Without loss of generality, we assume that is a regular point (i.e., at , the space of directions is isometric to standard , or equivalently, the tangent cone is isometric to ).
Choose a generating set inductively such that
-
(1)
for all ,
-
(2)
, ,
-
(3)
for all .
Clearly, if . Let be the minimal geodesic from to . We claim that for . Otherwise, consider the comparison triangle in . We get that . But then,
This contradicts our choice of .
Recall that an equivalent class of a minimal geodesic starting at is a direction at and the distance between two classes is the angle between them. Since is isometric to and for , the generating set we construct above contains at most 6 elements, i.e.
Thus we complete the proof. ∎
Proof of corollary 1.2.
Proof of Corollary 1.3.
As in Theorem 1.1, we only need to consider the case . If does not split, then by Theorem 2.8 and Theorem 1.1, we get the conclusion.
If splits, then by Theorem 2.7, is isomorphic to , where is an RCD() space. We claim that is compact. Otherwise by Theorem 2.9 and hence, where . This is a contradiction with our linear volume growth assumption.
Now suppose is infinitely generated. We apply Lemma 3.2 to obtain a sequence of independent generators of with minimal representative geodesic loops of length , and
Note that is of the form , where is a loop in and is a loop in . Thus is homotopic to and in . Since is minimal in the equivalent class, must be of the form .
Therefore, we get a sequence of loops on with
But is compact, which is a contradiction to . ∎
The author would like to thank Bobo Hua for helpful suggestions on this research project. The author would also like to thank Andrea Mondino for useful comments on an earlier version of this paper.
References
- [1] U. Abresch and D. Gromoll, On complete manifolds with nonnegative Ricci curvature, Journal of the American Mathematical Society, 3 (1990), pp. 355–374.
- [2] L. Ambrosio, N. Gigli, and G. Savaré, Calculus and heat flow in metric measure spaces and applications to spaces with ricci bounds from below, Inventiones mathematicae, 195 (2014), pp. 289–391.
- [3] , Metric measure spaces with riemannian Ricci curvature bounded from below, Duke Mathematical Journal, 163 (2014), pp. 1405–1490.
- [4] L. Ambrosio, A. Mondino, and G. Savaré, Nonlinear diffusion equations and curvature conditions in metric measure spaces, vol. 262, American Mathematical Society, 2019.
- [5] M. T. Anderson, On the topology of complete manifolds of non-negative Ricci curvature, Topology, 29 (1990), pp. 41–55.
- [6] K. Bacher and K.-T. Sturm, Localization and tensorization properties of the curvature-dimension condition for metric measure spaces, Journal of Functional Analysis, 259 (2010), pp. 28–56.
- [7] C. Brena, N. Gigli, S. Honda, and X. Zhu, Weakly non-collapsed RCD spaces are strongly non-collapsed, Journal für die reine und angewandte Mathematik (Crelles Journal), 2023 (2023), pp. 215–252.
- [8] E. Bruè, A. Naber, and D. Semola, Fundamental Groups and the Milnor Conjecture, arXiv preprint arXiv:2303.15347, (2023).
- [9] E. Brué and D. Semola, Constancy of the dimension for rcd (k, n) spaces via regularity of lagrangian flows, Communications on Pure and Applied Mathematics, 73 (2020), pp. 1141–1204.
- [10] J. Cheeger and T. H. Colding, On the structure of spaces with Ricci curvature bounded below. I, Journal of Differential Geometry, 46 (1997), pp. 406–480.
- [11] , On the structure of spaces with Ricci curvature bounded below. II, Journal of Differential Geometry, 54 (2000), pp. 13–35.
- [12] J. Cheeger and D. Gromoll, The splitting theorem for manifolds of nonnegative Ricci curvature, Journal of Differential Geometry, 6 (1971), pp. 119–128.
- [13] M. Erbar, K. Kuwada, and K.-T. Sturm, On the equivalence of the entropic curvature-dimension condition and Bochner’s inequality on metric measure spaces, Inventiones mathematicae, 201 (2015), pp. 993–1071.
- [14] N. Gigli, An overview of the proof of the splitting theorem in spaces with non-negative Ricci curvature, Analysis and Geometry in Metric Spaces, 2 (2014).
- [15] , On the differential structure of metric measure spaces and applications, American Mathematical Soc., 2015.
- [16] N. Gigli, A. Mondino, and G. Savaré, Convergence of pointed non-compact metric measure spaces and stability of ricci curvature bounds and heat flows, Proceedings of the London Mathematical Society, 111 (2015), pp. 1071–1129.
- [17] N. Gigli and S. Mosconi, The Abresch-Gromoll inequality in a non-smooth setting, Discrete & Continuous Dynamical Systems, 34 (2014), p. 1481.
- [18] M. Gromov, Manifolds of negative curvature, Journal of differential geometry, 13 (1978), pp. 223–230.
- [19] X.-t. Huang, Noncompact RCD (0, N) spaces with linear volume growth, arXiv preprint arXiv:1603.05221, (2016).
- [20] X.-T. Huang, An almost rigidity theorem and its applications to noncompact RCD (0, N) spaces with linear volume growth, Communications in Contemporary Mathematics, 22 (2020), p. 1850076.
- [21] Y. Kitabeppu and S. Lakzian, Non-branching RCD (0, N) geodesic spaces with small linear diameter growth have finitely generated fundamental groups, Canadian Mathematical Bulletin, 58 (2015), pp. 787–798.
- [22] , Characterization of low dimensional RCD*(K, N) spaces, Analysis and Geometry in Metric Spaces, 4 (2016).
- [23] P. Li, Large time behavior of the heat equation on complete manifolds with non-negative Ricci curvature, Annals of Mathematics, 124 (1986), pp. 1–21.
- [24] J. Lott and C. Villani, Ricci curvature for metric-measure spaces via optimal transport, Annals of Mathematics, (2009), pp. 903–991.
- [25] A. Lytchak and S. Stadler, Ricci curvature in dimension 2, Journal of the European Mathematical Society, 25 (2022), pp. 845–867.
- [26] J. Milnor, A note on curvature and fundamental group, Journal of Differential geometry, 2 (1968), pp. 1–7.
- [27] I. Mondello, A. Mondino, and R. Perales, An upper bound on the revised first Betti number and a torus stability result for RCD spaces, Comment. Math. Helv, 97 (2022), pp. 555–609.
- [28] A. Mondino and G. Wei, On the universal cover and the fundamental group of an RCD*(K, N)-space, Journal für die reine und angewandte Mathematik (Crelles Journal), 2019 (2019), pp. 211–237.
- [29] C. Sormani, The almost rigidity of manifolds with lower bounds on Ricci curvature and minimal volume growth, Communications in Analysis and Geometry, 8 (2000), pp. 159–212.
- [30] , Nonnegative Ricci curvature, small linear diameter growth and finite generation of fundamental groups, Journal of Differential Geometry, 54 (2000), pp. 547–559.
- [31] C. Sormani and G. Wei, Hausdorff convergence and universal covers, Transactions of the American Mathematical Society, 353 (2001), pp. 3585–3602.
- [32] E. H. Spanier, Algebraic topology, Springer Science & Business Media, 1989.
- [33] K.-T. Sturm, On the geometry of metric measure spaces. I, Acta Math, 196 (2006), pp. 65–131.
- [34] , On the geometry of metric measure spaces. II, Acta Math, 196 (2006), pp. 133–177.
- [35] C. Villani, Optimal transport: old and new, vol. 338, Springer, 2009.
- [36] J. Wang, Ricci limit spaces are semi-locally simply connected, arXiv preprint arXiv:2104.02460, (2021).
- [37] , RCD*(K, N) spaces are semi-locally simply connected, Journal für die reine und angewandte Mathematik (Crelles Journal), (2023).