Stability and Finiteness of Wasserstein Spaces
Abstract.
Under Gromov–Hausdorff convergence, and equivariant Gromov–Hausdorff convergence, we prove stability results of Wasserstein spaces over certain classes of singular and non-singular spaces. For example, we obtain an analogue of Perelman’s stability theorem on Wasserstein spaces.
Key words and phrases:
equivariant Gromov–Hausdorff convergence, Gromov–Hausdorff convergence, Wasserstein spaces, stability2020 Mathematics Subject Classification:
53C23, 53C20, 51K101. Introduction
Finiteness theorems for finite dimensional spaces with curvature bounds can be traced back to 1967, with Weinstein [32] proving that given any even natural number , and any , there are only finitely many homotopy types of -dimensional, -pinched manifolds. Shortly after, in 1970, Cheeger obtained several finiteness theorems in his celebrated work [10]. For instance, he proved that given any natural number , and real numbers , there are only finitely many diffeomorphism types of compact -dimensional manifolds admitting a Riemannian metric such that diameter , vol and , where denotes the sectional curvature. These results served as departure points for further seminal and important developments in Riemannian geometry (see for instance [3, 21, 22, 20, 29]).
In 1981, Gromov [19] introduced a powerful notion of convergence, now known as Gromov–Hausdorff convergence, that has been useful in establishing (among other things) finiteness theorems through the notion of “stability”. For instance, using Gromov–Hausdorff convergence, Grove, Petersen and Wu [22], in 1990, relaxed the curvature condition in Cheeger’s finiteness theorem and showed that for any , real number , and any positive numbers and , the class consisting of compact -dimensional Riemannian manifolds with diameter, , and vol contains finitely many homeomorphism types.
In 1991, Perelman [29, 25], using Gromov–Hausdorff convergence, showed that given a non-collapsing convergent sequence of compact Alexandrov spaces, with a uniform lower curvature bound, uniform finite dimension and uniform upper diameter bound, the tail of the sequence stabilizes topologically. That is, all spaces in the tail of the sequence are mutually homeomorphic. As a corollary, one obtains homeomorphism finiteness (in all dimensions) for a more general class of spaces that need not be manifolds.
In all cases of finiteness and stability, attention has been restricted to finite dimensional spaces. Thus, the question of stability in infinite dimensions remains open. In this paper, we focus our attention on Wasserstein spaces, which are infinite-dimensional, and prove stability and finiteness results for these spaces in different settings. Our proofs are not technical, and these results are, to the best of our knowledge, the first stability and finiteness results for infinite-dimensional spaces.
Let denote the class of compact Alexandrov spaces with dimension at most , curvature bounded below by , and diameter bounded above by . We denote by all Alexandrov spaces in with dimension . Given , we will denote by the subset of where the Alexandrov spaces under consideration have volume bounded below by . Furthermore, given a sequence of compact metric spaces and a compact metric space , we will denote by the Gromov–Hausdorff convergence of to . We will denote by the Wasserstein space over equipped with the -Wasserstein metric (see Definition 2.5).
Theorem A.
Assume and are in . Then the following assertions are equivalent:
-
(1)
.
-
(2)
.
-
(3)
.
We point out a few remarks concerning the above theorem. First, the interesting implication is . The implication is well known (see Corollary 4.3 in [28]) and holds for general compact metric spaces. Second, the notion of convergence in is known as Gromov–Hausdorff convergence of metric pairs. It was introduced in [9], and later studied extensively in [1]. This notion of convergence generalizes Gromov–Hausdorff convergence and takes into account pairs of the form , where is a metric space, and is a closed non-empty subset of . In our theorem, we identify the base space with the set of Dirac deltas in the Wasserstein space . Thus, becomes trivial. However, the new idea for this implication is that we give an equivalent characterization of convergence of metric pairs using the notion of -isometries (see Proposition 2.1). This equivalent characterization will be useful when we introduce the notion of relative equivariant Gromov–Hausdorff convergence (see Definition 2.2) which takes into account a triple , where is a compact metric space, is a closed, non-empty -invariant subset of and is a closed subgroup of , the group of isometries of .
A consequence of the above theorem, is the following analogue of Perelman’s stability theorem for Wasserstein spaces.
Corollary B.
Let and be spaces in . If then for all sufficiently large , and are homeomorphic, and the homeomorphisms can be taken to be Gromov–Hausdorff approximations.
Note that in the preceding corollary we do not impose any other structure on the Wasserstein spaces other than that coming from the base spaces.
We further have the following finiteness result. For any , and , let denote the class consisting of Wasserstein spaces with base spaces in and equipped with the -Wasserstein metric.
Corollary C.
There are only finitely many topological types in .
We note that the technique in this paper can be applied whenever we have a stability result on the base spaces and essentially reduces to the following result.
Theorem D.
Let denote a precompact class (with respect to the Gromov–Hausdorff topology) of compact non-branching metric measure spaces having good transport behavior and their reference measures are qualitivatively non-degenerate. Assume that if is a sequence in , and , then is a non-branching compact metric space admitting a measure that is qualitatively non-degenerate and has good transport behavior (we will take such a measure as a reference measure on ). Then, the following assertions are equivalent:
-
(1)
.
-
(2)
.
The notion of good transport behavior was first introduced in the paper by Galaz-García, Kell, Mondino, and Sosa [18], and further studied by Kell [26]. The class of metric measure spaces satisfying such condition is large. For instance, according to [18], strong and essentially non-branching spaces admit good transport behavior. The notion of a measure being qualitatively non-degenerate was introduced by Cavalletti and Huesmann [7] and later studied by Kell [26]. It is technical to state, and we refer the reader to [26, 7, 30] for more details, but we note a measure being qualitatively non-degenerate yields desirable information. For example, a qualitatively non-degenerate measure is doubling (Lemma 2.8 in [30]).
We note that the classes , satisfy the hypotheses of Theorem D. However, the properties in Theorem D are not closed under Gromov–Hausdorff convergence in general. This is because a compact metric measure space endowed with a qualitatively non-degenerate measure is doubling and thus must have finite dimension.
Using almost verbatim the proofs of Theorems D and A, and noting that spaces are non-branching [14], we obtain the following corollary.
Corollary E.
Assume each is an compact space with uniform upper diameter bound. Assume is an -dimensional compact Riemannian manifold. If then for all large , there exists a homeomorphism that is Lipschitz and its inverse is Hölder.
A consequence of our next main theorem is an equivariant stability result for Wasserstein spaces. We first discuss notation. Given a closed subgroup of , where is a compact metric space, we denote by the induced subgroup of given by the push-forward of maps. Given a sequence of pairs , where each is a compact metric space and is a closed subgroup of , we will denote, by the equivariant Gromov–Hausdorff convergence of to . This convergence, originally introduced by Fukaya (see [16, 17, 15]), has been useful in establishing interesting results in both the singular and non-singular setting (see for instance [34, 12, 5, 23, 33, 13, 8]). Furthermore, it has been useful in obtaining finiteness and stability results in the singular setting. For instance, using equivariant Gromov–Hausdorff convergence, Zamora extended Anderson finiteness to the setting [33].
Theorem F.
Let and be closed Riemannian manifolds with uniform lower sectional curvature bound , uniform upper diameter bound and uniform upper dimension bound. Assume and are closed subgroups of and respectively. Then the following assertions are equivalent:
-
(1)
.
-
(2)
There exists unique closed subgroups of and of such that and and such that .
-
(3)
.
We emphasize that we identify the base space with the (closed) subset of Dirac deltas in the Wasserstein space. In the presence of a uniform lower curvature bound, we have the following corollary, an equivariant stability result on Wasserstein spaces.
Corollary G.
Let and be closed Riemannian manifolds with uniform lower sectional curvature bound , uniform upper diameter bound and uniform dimension. Assume and are closed subgroups of and respectively of the same dimension. If , then for all large , there exists a homeomorphism that is compatible with some group isomorphism .
Our article is organized as follows. In section 2, we discuss the preliminaries that will be useful in this paper. Moreover, we will give a new characterization of convergence of metric pairs. In section 3 we give proofs of the theorems and corollaries.
Acknowledgements.
It is a pleasure to thank Fernando Galaz–García, Martin Kerin, Kohei Suzuki, Jaime Santos-Rodríguez, and Mauricio Che for helpful conversations and comments during the Durham Metric Geometry Reading Seminar. I would also like to thank Sergio Zamora and Yanpeng Zhi for several interesting conversations and Aseel alnajjar for several comments on the first draft of this manuscript. Part of this work was completed at the Erwin Schrödinger International Institute for Mathematics and Physics (ESI), where the author received financial support to participate in the workshop Synthetic Curvature Bounds for Non-Smooth Spaces: Beyond Finite Dimension. The author would like to thank ESI and the conference organizers for the excellent atmosphere and working conditions.
2. Preliminaries
In this section, we collect preliminary definitions and results we will use in the proof of our main results.
2.1. Convergence of Metric Pairs and Equivariant Convergence of Metric Pairs
In this section, we give a new equivalent characterization of the notion of convergence of metric pairs which will allow us to introduce, canonically, the notion of relative equivariant Gromov–Hausdorff convergence. The equivalent characterization is demonstrated in the following proposition. We refer the reader to the paper by Ahumada Gómez and Che [1] for results concerning the (non-equivariant) convergence of metric pairs.
Proposition 2.1.
Let be a sequence of compact metric spaces. Let be a collection of closed non-empty subspaces . Let be a compact metric space, and a closed non-empty subset. Then the following assertions are equivalent:
-
(1)
.
-
(2)
There exists a sequence of positive numbers such that and -Gromov–Hausdorff approximations , such that and are -close, i.e., for all .
Proof.
Assume . Then there exists a sequence of numbers that tends to and -Gromov–Hausdorff approximations such that . Here, denotes the Hausdorff distance. We define as follows. Since , for each choose such that . We will show that the map is a -Gromov–Hausdorff approximation. Indeed, for each , there exists an such that . Therefore, by the triangle inequality, we get . Now we show that is almost distance preserving up to an error . We only show one side of the inequality, for the other is similar. Let . Then
Now note that and . Thus, as is an -Gromov–Hausdorff approximation, we get
Now we show the reverse direction. Define by
Since is an -Gromov–Hausdorff approximation, it follows that . It remains to verify that is a Gromov–Hausdorff approximation. It is clear that is -surjective. Therefore, it remains to prove that is almost distance preserving up to an error that tends to . Once again, we only verify one direction of the inequality, since the other is similar. Let . It is clear that it suffices to assume that and that . Then
Since is an -Gromov–Hausdorff approximation, . Since and are -close, . ∎
The proposition above suggests the following definition of relative equivariant Gromov–Hausdorff convergence.
Definition 2.2.
Let and be compact metric spaces. Suppose and are closed subgroups of and respectively. Assume is a closed non-empty -invariant subset of , and is a closed non-empty -invariant subset of . For , we define an -equivariant Gromov–Hausdorff approximation between and to be a quadruple of maps
subject to the following conditions:
-
(1)
The triple is an -equivariant Gromov–Hausdorff approximation (in the usual sense).
-
(2)
is an -Gromov–Hausdorff approximation.
-
(3)
The maps and are -close.
Remark 2.3.
We use the terms “equivariant Gromov–Hausdorff approximations” and “equivariant approximations” interchangeably.
Definition 2.4.
Given a sequence of triples , where is a compact metric space, is a closed subgoup of and is a closed non-empty -invariant subset of , we say that Gromov–Hausdorff equivariantly converges to , denoted , where is a compact metric space, is a closed subgroup of Isom and is a closed subset of that is -invariant, if there exists a sequence of positive real numbers such that and -equivariant Gromov–Hausdorff approximations .
2.2. Further definitions and auxiliary results
For the convenience of the reader, we include the definitions of Wasserstein spaces, good transport behavior and isometric rigidity.
Definition 2.5.
Let be a Polish metric space and let . Given any two probability measure and on , we define the -Wasserstein metric between and , denoted by , to be
where the infimum is taken over all admissible measures having marginals and . The -Wasserstein metric is a metric on the space of probability measures with finite -moments, which is denoted by .
Definition 2.6.
A metric measure space is said to have good transport behavior if, for any two probability measures in , where , any optimal transport map between them is induced by a map.
We motivate the concept of isometric rigidity by the following example.
Example 2.7.
If is a compact metric space and is an isometry, then induces an isometry .
Thus, it is natural to formulate the following definition.
Definition 2.8.
Let be a metric space. Then, is said to be isometrically rigid if given any isometry , there exists an isometry such that . In particular, every isometry of is induced by an isometry of ; the base space.
The concept of isometric rigidity was first introduced by Kloeckner in the context of euclidean space in [27], where he showed, among other things, that, is not isometrically rigid. Later, Bertrand and Kloeckner [4] studied the concept of isometric rigidity and showed that if is a negatively curved geodesically complete Hadamard space, then is isometrically rigid. In 2021, Santos Rodriguez showed [30] that if is a positively curved closed Riemannian manifold, then is isometrically rigid and that, if is a CROSS, then, in fact, the isometry groups of and are the same for all . In the same paper, Santos Rodriguez showed the following result which we will use frequently in this paper (Corollary 3.8, [30]).
Proposition 2.9.
Let and be two compact non-branching metric measure spaces equipped with qualitatively non-degenerate measures and such that they have good transport behavior. Suppose that there exists an isometry . Then and are isometric.
In fact, Santos Rodriguez in [30] showed the above proposition for Wasserstein spaces with finite -moments, where .
3. Proofs
In what follows, we assume that the maps are sufficiently regular. Indeed, this is possible because if is a compact metric space then is compact with the compact-open topology. Moreover, the compact open topology on can be metrized with the uniform metric. One can always replace -isometries between compact metric spaces by almost-isometries that are measurable. In particular, we have the following simple lemma.
Lemma 3.1.
Assume and are compact metric spaces and and are closed subgroups of and respectively. Equip and with the uniform metrics and respectively. If is an -equivariant approximation, then, , and can be chosen to be measurable.
Proof.
By Corollary 3.4 in [2], it follows that is an -approximation. Furthermore, observe that is an almost inverse to . In particular, for , and , one has and . Hence, as shown in Lemma 4.1 in [11], one can obtain measurable approximations , and that remain close to , and respectively. Hence the result follows. ∎
Proof of Theorem D
The implication is well known (see Corollary 4.3 in [28]). Let us show . Assume for the sake of obtaining a contradiction that holds but does not hold. Hence, there exists a such that (up to a subsequence), for all . Since the class is pre-compact, and by our assumptions, it follows that up to a further subsequence, , where is a compact non-branching metric space. Hence, . By assumption, and are isometric. Now, by Proposition 2.9, and are therefore isometric, which is a contradiction. ∎
Proof of Theorem A
The proof is similar to the proof of Theorem D. However, we give details for the convenience of the reader. The implications and are clear. The implication follows at once from the following observation. Let be -approximations by measurable maps. Then each induces an -approximation , with , and such that, on Dirac deltas, is a Hausdorff approximation. Hence, one may use Proposition 2.1 to establish this implication. We shall now show .
We must prove that . Assume otherwise. Hence, there exists such that, up to a subsequence, for all . Here, denotes the Gromov–Hausdorff distance. Since is, with respect to , compact [6], there exists a compact Alexandrov space such that, up to a further subsequence, . Hence, . By assumption, . Therefore, and are isometric. Alexandrov spaces have good transport behavior, and on them, the Hausdorff measure is qualitatively non-degenerate, therefore and are isometric. Hence we have . ∎
Proof of Corollary B
The result follows from Perelman’s stability Theorem. Note that if is a map, then is continuous if and only if is continuous (see Remark 4.14 in [31]). ∎
Proof of Corollary C
We proceed by contradiction. Assume we have a sequence of topologically pairwise inequivalent Wasserstein spaces in . Then, since is precompact and we have volume bounded below by , there exists a compact -dimensional Alexandrov space such that (up to a subsequence), . Hence, . Therefore, by Corollary B, for all large , and are homeomorphic, which is a contradiction.
∎
Proof of Corollary E
This follows in an analogous manner as in the proof of Corollary B. However, one uses the result in [24] instead of Perelman’s stability theorem.
∎
Now we will prove Theorem F. First, we need two lemmas.
Lemma 3.2.
Let and be compact metric spaces and and are closed subgroups of and respectively. An -equivariant approximation, between and induces an -equivariant approximation between and , where tends to as tends to .
Proof.
Define and by the rules and . By Corollary 4.3 in [28], the map is an -approximation, where
Let . Then,
The last inequality is similar. ∎
As a corollary, we have the following result.
Corollary 3.3.
Let and be compact metric spaces. If , then .
We will also need the following lemma, an equivariant analogue of Proposition 2.9.
Lemma 3.4.
Let and be positively curved closed Riemannian manifolds and let and be closed subgroups of and , respectively. If and are equivariantly isometric then so are and .
Proof.
Let be an isometry and let be an isomorphism such that for any , one has . Then, from Corollary 3.8 in [30], it follows that sends Dirac deltas onto Dirac deltas. Thus, there exists an isometry such that and agree on Dirac deltas. Moreover, there exists an isomorphism such that for , . Thus it follows that . ∎
We are now ready to prove theorem F.
Proof of Theorem F
The implication follows from Lemma 3.2. The implication is trivial. Now we shall verify the implication Since and are isometrically rigid, there exists unique closed subgroups of and of such that and . It now remains to verify that . As usual, assume otherwise. So, there exists a such that, up to a subsequence, for all . Here, denotes the equivariant Gromov–Hausdorff distance. Since forms a family of compact Alexandrov spaces with curvature uniformly bounded below, dimension uniformly bounded above, and uniform upper diameter bound, it follows that there exists a compact Alexandrov space , such that, up to a subsequence, . Up to a further subsequence, there exists a closed subgroup of such that . Hence, by Lemma 3.2, it follows that . Therefore, and are equivariantly isometric. The previous lemma shows that and are equivariantly isometric, which is a contradiction. ∎
Proof of Corollary G
References
- [1] Andrés Ahumada Gómez and Mauricio Che “Gromov–Hausdorff convergence of metric pairs and metric tuples” In Differential Geom. Appl. 94, 2024, pp. Paper No. 102135 DOI: 10.1016/j.difgeo.2024.102135
- [2] Mohammad Alattar “Stability and equivariant Gromov–Hausdorff convergence” In Bulletin of the London Mathematical Society 47, 2024 DOI: https://doi.org/10.1112/blms.13073
- [3] Michael T. Anderson and Jeff Cheeger “Diffeomorphism finiteness for manifolds with Ricci curvature and -norm of curvature bounded” In Geom. Funct. Anal. 1.3, 1991, pp. 231–252 DOI: 10.1007/BF01896203
- [4] Jérôme Bertrand and Benoît R. Kloeckner “A geometric study of Wasserstein spaces: isometric rigidity in negative curvature” In Int. Math. Res. Not. IMRN, 2016, pp. 1368–1386 DOI: 10.1093/imrn/rnv177
- [5] Elia Brue, Aaron Naber and Daniele Semola “Stability of Tori under Lower Sectional Curvature”, 2023 arXiv:2307.03824 [math.DG]
- [6] Yuri. Burago, Misha Gromov and G. Perelman “A. D. Aleksandrov spaces with curvatures bounded below” In Uspekhi Mat. Nauk 47.2(284), 1992, pp. 3–51\bibrangessep222 DOI: 10.1070/RM1992v047n02ABEH000877
- [7] Fabio Cavalletti and Martin Huesmann “Existence and uniqueness of optimal transport maps” In Ann. Inst. H. Poincaré C Anal. Non Linéaire 32.6, 2015, pp. 1367–1377 DOI: 10.1016/j.anihpc.2014.09.006
- [8] Nicola Cavallucci “A GH-compactification of CAT-groups via totally disconnected, unimodular actions”, 2023 arXiv:2307.05640 [math.MG]
- [9] Mauricio Che, Fernando Galaz-García, Luis Guijarro and Ingrid Membrillo Solis “Metric Geometry of Spaces of Persistence Diagrams”, 2022 arXiv:2109.14697 [math.MG]
- [10] Jeff Cheeger “Finiteness theorems for Riemannian manifolds” In Amer. J. Math. 92, 1970, pp. 61–74 DOI: 10.2307/2373498
- [11] Nhan-Phu Chung “Gromov-Hausdorff distances for dynamical systems” In Discrete Contin. Dyn. Syst. 40.11, 2020, pp. 6179–6200 DOI: 10.3934/dcds.2020275
- [12] Diego Corro “Collapsing regular Riemannian foliations with flat leaves”, 2024 arXiv:2403.11602 [math.DG]
- [13] Diego Corro, Jaime Santos-Rodríguez and Jesús Núñez-Zimbrón “Cohomogeneity one RCD-spaces”, 2024 arXiv:2405.09448 [math.MG]
- [14] Qin Deng “Hölder continuity of tangent cones in RCD(K,N) spaces and applications to non-branching”, 2020 arXiv:2009.07956 [math.DG]
- [15] Kenji Fukaya “Hausdorff convergence of Riemannian manifolds and its applications” In Recent topics in differential and analytic geometry 18, Adv. Stud. Pure Math. Academic Press, Boston, MA, 1990, pp. 143–238 DOI: 10.2969/aspm/01810143
- [16] Kenji Fukaya “Theory of convergence for Riemannian orbifolds” In Japan. J. Math. (N.S.) 12.1, 1986, pp. 121–160 DOI: 10.4099/math1924.12.121
- [17] Kenji Fukaya and Takao Yamaguchi “The fundamental groups of almost non-negatively curved manifolds” In Ann. of Math. (2) 136.2, 1992, pp. 253–333 DOI: 10.2307/2946606
- [18] Fernando Galaz-García, Martin Kell, Andrea Mondino and Gerardo Sosa “On quotients of spaces with Ricci curvature bounded below” In J. Funct. Anal. 275.6, 2018, pp. 1368–1446 DOI: 10.1016/j.jfa.2018.06.002
- [19] Mikhael Gromov “Groups of polynomial growth and expanding maps” In Inst. Hautes Études Sci. Publ. Math., 1981, pp. 53–73 URL: http://www.numdam.org/item?id=PMIHES_1981__53__53_0
- [20] Karsten Grove “Finiteness theorems in Riemannian geometry” In Explorations in complex and Riemannian geometry 332, Contemp. Math. Amer. Math. Soc., Providence, RI, 2003, pp. 101–120 DOI: 10.1090/conm/332/05931
- [21] Karsten Grove and Peter Petersen “Bounding homotopy types by geometry” In Ann. of Math. (2) 128.1, 1988, pp. 195–206 DOI: 10.2307/1971439
- [22] Karsten Grove, Peter Petersen and Jyh Yang Wu “Geometric finiteness theorems via controlled topology” In Invent. Math. 99.1, 1990, pp. 205–213 DOI: 10.1007/BF01234417
- [23] John Harvey “Convergence of isometries, with semicontinuity of symmetry of Alexandrov spaces” In Proceedings of the American Mathematical Society 144.8, 2016, pp. 3507–3515
- [24] Shouhei Honda and Yuanlin Peng “A note on the topological stability theorem from RCD spaces to Riemannian manifolds” In Manuscripta Math. 172.3-4, 2023, pp. 971–1007 DOI: 10.1007/s00229-022-01418-7
- [25] Vitali Kapovitch “Perelman’s stability theorem” In Surveys in differential geometry. Vol. XI 11, Surv. Differ. Geom. Int. Press, Somerville, MA, 2007, pp. 103–136 DOI: 10.4310/SDG.2006.v11.n1.a5
- [26] Martin Kell “Transport maps, non-branching sets of geodesics and measure rigidity” In Adv. Math. 320, 2017, pp. 520–573 DOI: 10.1016/j.aim.2017.09.003
- [27] Benoît Kloeckner “A geometric study of Wasserstein spaces: Euclidean spaces” In Ann. Sc. Norm. Super. Pisa Cl. Sci. (5) 9.2, 2010, pp. 297–323
- [28] John Lott and Cédric Villani “Ricci curvature for metric-measure spaces via optimal transport” In Ann. of Math. (2) 169.3, 2009, pp. 903–991 DOI: 10.4007/annals.2009.169.903
- [29] Grigori Perelman “Alexandrov spaces with curvatures bounded from below II” In preprint, 1991
- [30] Jaime Santos-Rodríguez “On isometries of compact -Wasserstein spaces” In Adv. Math. 409, 2022, pp. Paper No. 108632\bibrangessep21 DOI: 10.1016/j.aim.2022.108632
- [31] Cédric Villani “Optimal transport: old and new” Springer, 2009
- [32] Alan Weinstein “On the homotopy type of positively-pinched manifolds” In Arch. Math. (Basel) 18, 1967, pp. 523–524 DOI: 10.1007/BF01899493
- [33] Sergio Zamora “Anderson finiteness for RCD spaces” In MPIM Preprint Series (4), 2023 URL: https://archive.mpim-bonn.mpg.de/id/eprint/5009/1/mpim-preprint_2023-4.pdf
- [34] Sergio Zamora “First Betti number and collapse”, 2022 arXiv:2209.12628 [math.DG]