Skip to main content
Log in

On the Existence of Monge Maps for the Gromov–Wasserstein Problem

  • Published:
Foundations of Computational Mathematics Aims and scope Submit manuscript

Abstract

The Gromov–Wasserstein problem is a non-convex optimization problem over the polytope of transportation plans between two probability measures supported on two spaces, each equipped with a cost function evaluating similarities between points. Akin to the standard optimal transportation problem, it is natural to ask for conditions guaranteeing some structure on the optimizers, for instance, if these are induced by a (Monge) map. We study this question in Euclidean spaces when the cost functions are either given by (i) inner products or (ii) squared distances, two standard choices in the literature. We establish the existence of an optimal map in case (i) and of an optimal 2-map (the union of the graphs of two maps) in case (ii), both under an absolute continuity condition on the source measure. Additionally, in case (ii) and in dimension one, we numerically design situations where optimizers of the Gromov–Wasserstein problem are 2-maps but are not maps. This suggests that our result cannot be improved in general for this cost. Still in dimension one, we additionally establish the optimality of monotone maps under some conditions on the measures, thereby giving insight into why such maps often appear to be optimal in numerical experiments.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Algorithm 1
Fig. 3
Algorithm 2
Fig. 4
Fig. 5
Fig. 6

Similar content being viewed by others

Notes

  1. Note that in the inner product case, \({\text {GW}}(\mu ,\nu )=0\) implies that the metric measure spaces are isomorphic, but the converse is not true—for instance, in cases where \((\mathcal {X}, d_\mathcal {X}, \mu )\) and \((\mathcal {Y}, d_\mathcal {Y}, \nu )\) are point clouds in an Euclidean space that simply differ by a translation, meaning \(\mathcal {Y}=T(\mathcal {X})\), \(d_{\mathcal {X}}=d_{\mathcal {Y}}\) the Euclidean distance, and \(\nu =T_\#\mu \), where \(T:x\mapsto x+x_0\).

  2. Note that we cannot apply Proposition 2.9 to the measures \((\mu _u,\nu _{t_B(u)})_u\) and the maps \((T_u)_u\) directly, as \(T_u\) may not be the unique Monge map between the measures, a required assumption of the proposition.

  3. We consider \(\nu \) as a measure on \(\mathbb {R}^n\) with d-dimensional support.

  4. The link of the code is https://github.com/theodumont/monge-gromov-wasserstein.

  5. Note: this is where the choice is made, as in the proof of Lemma 3.16, between the non-decreasing and non-increasing rearrangements. Using this convention, the non-decreasing rearrangement is shown to be optimal.

References

  1. Luigi Ambrosio, Nicola Gigli, and Giuseppe Savaré. Gradient flows: in metric spaces and in the space of probability measures. Springer Science & Business Media, 2005.

  2. David Alvarez-Melis and Tommi S Jaakkola. Gromov–Wasserstein alignment of word embedding spaces. 2018. arXiv: 1809.00013

  3. Najma Ahmad, Hwa Kil Kim, and Robert J McCann. “Optimal transportation, topology and uniqueness”. In: Bulletin of Mathematical Sciences 1.1 (2011), pp. 13–32.

  4. Alexandr Andoni, Assaf Naor, and Ofer Neiman. “Snowflake universality of Wasserstein spaces”. In: Annales Scientifiques de íÉcole Normale Supérieure. Vol. 51. Societe Mathematique de France. 2018, pp. 657–700.

  5. Brandon Amos, Lei Xu, and J Zico Kolter. “Input convex neural networks”. In: International Conference on Machine Learning. PMLR. 2017, pp. 146–155.

  6. Alexander C Berg, Tamara L Berg, and Jitendra Malik. “Shape matching and object recognition using low distortion correspondences”. In: 2005 IEEE computer society conference on computer vision and pattern recognition (CVPR’05). Vol. 1. IEEE. 2005, pp. 26–33.

  7. Robert Beinert, Cosmas Heiss, and Gabriele Steidl. On assignment problems related to Gromov–Wasserstein distances on the real line. 2022. arXiv: 2205.09006.

  8. Mathieu Blondel, Olivier Teboul, Quentin Berthet, and Josip Djolonga. “Fast differentiable sorting and ranking”. In: International Conference on Machine Learning. PMLR. 2020, pp. 950–959.

  9. Yann Brenier. “Décomposition polaire et réarrangement monotone des champs de vecteurs”. In: CR Acad. Sci. Paris Sér. I Math. 305 (1987), pp. 805–808.

  10. Charlotte Bunne, David Alvarez-Melis, Andreas Krause, and Stefanie Jegelka. Learning generative models across incomparable spaces. 2019. arXiv: 1905.05461.

  11. Guillaume Carlier. “Optimal transportation and economic applications”. In: Lecture Notes (2012).

  12. David Chiron. “On the definitions of Sobolev and BV spaces into singular spaces and the trace problem”. In: Communications in Contemporary Mathematics 9.04 (2007), pp. 473–513.

  13. Samir Chowdhury and Facundo Mémoli. “The Gromov–Wasserstein distance between networks and stable network invariants”. In: Information and Inference: A Journal of the IMA 8.4 (2019), pp. 757–787.

  14. Pierre-André Chiappori, Robert J McCann, and Lars P Nesheim. “Hedonic price equilibria, stable matching, and optimal transport: equivalence, topology, and uniqueness”. In: Economic Theory 42.2 (2010), pp. 317–354.

  15. Pınar Demetçi, Rebecca Santorella, Björn Sandstede, William Stafford Noble, and Ritambhara Singh. Gromov–Wasserstein optimal transport to align single-cell multi-omics data. 2020. https://doi.org/10.1101/2020.04.28.066787.

    Article  Google Scholar 

  16. Pınar Demetçi, Rebecca Santorella, Manav Chakravarthy, Bjorn Sandstede, and Ritambhara Singh. “SCOTv2: single-cell multiomic alignment with disproportionate cell-type representation”. In: Journal of Computational Biology 29.11 (2022), pp. 1213–1228.

  17. Nicoló De Ponti and Andrea Mondino. “Entropy–Transport distances between unbalanced metric measure spaces”. In: Probability Theory and Related Fields 184.1 (2022), pp. 159–208.

  18. Joaquin Fontbona, Hélène Guérin, and Sylvie Méléard. “Measurability of optimal transportation and strong coupling of martingale measures”. In: Electronic communications in probability 15 (2010), pp. 124–133.

    Article  MathSciNet  Google Scholar 

  19. Jacques Faraut and Khélifa Harzallah. “Distances hilbertiennes invariantes sur un espace homogène”. In: Annales de l’Institut Fourier. Vol. 24. 1974, pp. 171–217.

  20. Edouard Grave, Armand Joulin, and Quentin Berthet. “Unsupervised alignment of embeddings with wasserstein procrustes”. In: The 22nd International Conference on Artificial Intelligence and Statistics. PMLR. 2019, pp. 1880–1890.

  21. Mikhael Gromov, Misha Katz, Pierre Pansu, and Stephen Semmes. Metric structures for Riemannian and non-Riemannian spaces. Vol. 152. Springer, 1999.

  22. YoonHaeng Hur, Wenxuan Guo, and Tengyuan Liang. Reversible Gromov–Monge sampler for simulation-based inference. 2021. arXiv: 2109.14090.

  23. Tjalling C Koopmans and Martin Beckmann. “Assignment problems and the location of economic activities”. In: Econometrica: journal of the Econometric Society (1957), pp. 53–76.

  24. Hiroshi Konno. “Maximization of a convex quadratic function under linear constraints”. In: Mathematical programming 11.1 (1976), pp. 117–127.

  25. Alexander Korotin, Lingxiao Li, Aude Genevay, Justin M Solomon, Alexander Filippov, and Evgeny Burnaev. “Do neural optimal transport solvers work? a continuous wasserstein-2 benchmark”. In: Advances in Neural Information Processing Systems 34 (2021), pp. 14593–14605.

  26. Alexander Korotin, Daniil Selikhanovych, and Evgeny Burnaev. “Neural optimal transport”. In: arXiv preprint arXiv:2201.12220 (2022).

  27. Grégoire Loeper. “On the regularity of solutions of optimal transportation problems”. In: Acta mathematica 202.2 (2009), pp. 241–283.

  28. Russell Lyons. “Distance covariance in metric spaces”. In: The Annals of Probability 41.5 (2013), pp. 3284–3305.

  29. Ashok Makkuva, Amirhossein Taghvaei, Sewoong Oh, and Jason Lee. “Optimal transport mapping via input convex neural networks”. In: International Conference on Machine Learning. PMLR. 2020, pp. 6672–6681.

  30. Robert J McCann. “Polar factorization of maps on Riemannian manifolds”. In: Geometric & Functional Analysis GAFA 11.3 (2001), pp. 589–608.

  31. Robert J McCann. A glimpse into the differential topology and geometry of optimal transport. 2012. arXiv: 1207.1867.

  32. Facundo Mémoli, Axel Munk, Zhengchao Wan, and Christoph Weitkamp. “The ultrametric Gromov–Wasserstein distance”. In: Discrete & Computational Geometry (2023), pp. 1–73.

  33. Facundo Mémoli. “On the use of Gromov–Hausdorff distances for shape comparison”. In: Eurographics Symposium on Point-Based Graphics. The Eurographics Association, 2007.

  34. Facundo Mémoli. “Gromov–Hausdorff distances in Euclidean spaces”. In: 2008 IEEE Computer Society Conference on Computer Vision and Pattern Recognition Workshops. IEEE. 2008, pp. 1–8.

  35. Facundo Mémoli. “Gromov–Wasserstein distances and the metric approach to object matching”. In: Foundations of Computational Mathematics 11.4 (2011), pp. 417–487.

  36. Robert J McCann and Nestor Guillen. “Five lectures on optimal transportation: geometry, regularity and applications”. In: Analysis and geometry of metric measure spaces: lecture notes of the s é minaire de Math é matiques Sup é rieure (SMS) Montr é al (2011), pp. 145–180.

  37. Haggai Maron and Yaron Lipman. “(Probably) concave graph matching”. In: Advances in Neural Information Processing Systems 31 (2018).

  38. Facundo Mémoli and Tom Needham. Comparison results for Gromov–Wasserstein and Gromov–Monge distances. 2022. arXiv: 2212.14123.

  39. Facundo Mémoli and Tom Needham. “Distance distributions and inverse problems for metric measure spaces”. In: Studies in Applied Mathematics 149.4 (2022), pp. 943–1001.

  40. Abbas Moameni. “A characterization for solutions of the Monge–Kantorovich mass transport problem”. In: Mathematische Annalen 365.3 (2016), pp. 1279–1304.

  41. Robert J McCann, Brendan Pass, and Micah Warren. “Rectifiability of optimal transportation plans”. In: Canadian Journal of Mathematics 64.4 (2012), pp. 924–934.

  42. Abbas Moameni and Ludovic Rifford. “Uniquely minimizing costs for the Kantorovitch problem”. In: Annales de la Facult é des sciences de Toulouse: Math é matiques. Vol. 29. 2020, pp. 507–563.

  43. Facundo Mémoli and Guillermo Sapiro. “Comparing point clouds”. In: Proceedings of the 2004 Eurographics/ACM SIGGRAPH symposium on Geometry processing. 2004, pp. 32–40.

  44. Facundo Mémoli and Guillermo Sapiro. “A theoretical and computational framework for isometry invariant recognition of point cloud data”. In: Foundations of Computational Mathematics 5.3 (2005), pp. 313–347.

  45. Facundo Mémoli, Zane Smith, and Zhengchao Wan. The Gromov–Hausdorff distance between ultrametric spaces: its structure and computation. 2021. arXiv: 2110.03136.

  46. Adam Paszke, Sam Gross, Francisco Massa, Adam Lerer, James Bradbury, Gregory Chanan, Trevor Killeen, Zeming Lin, Natalia Gimelshein, Luca Antiga, Alban Desmaison, Andreas Kopf, Edward Yang, Zachary DeVito, Martin Raison, Alykhan Tejani, Sasank Chilamkurthy, Benoit Steiner, Lu Fang, Junjie Bai, and Soumith Chintala. “PyTorch: An Imperative Style, High-Performance Deep Learning Library”. In: Advances in Neural Information Processing Systems. Vol. 32. 2019, pp. 8024–8035.

  47. Gabriel Peyré and Marco Cuturi. “Computational optimal transport: with applications to data science”. In: Foundations and Trends in Machine Learning 11.5-6 (2019), pp. 355–607.

  48. Ivegen Redko, Titouan Vayer, Rémi Flamary, and Nicolas Courty. “CO-optimal transport”. In: Advances in Neural Information Processing Systems 33 (2020), pp. 17559–17570.

    Google Scholar 

  49. R Tyrrell Rockafellar and Roger J-B Wets. Variational analysis. Vol. 317. Springer Science & Business Media, 2009.

  50. Filippo Santambrogio. “Optimal transport for applied mathematicians”. In: Birk ä user, NY 55.58-63 (2015), p. 94.

  51. Isaac J Schoenberg. “Metric spaces and positive definite functions”. In: Transactions of the American Mathematical Society 44 (1938), pp. 522–536.

  52. Antoine Salmona, Julie Delon, and Agnès Desolneux. Gromov–Wasserstein distances between Gaussian distributions. 2021. arXiv: 2104.07970.

  53. Karl-Theodor Sturm. “On the geometry of metric measure spaces”. In: Acta mathematica 196.1 (2006), pp. 65–131.

  54. Karl-Theodor Sturm. The space of spaces: curvature bounds and gradient flows on the space of metric measure spaces. 2012. arXiv: 1208.0434.

  55. Thibault Séjourné, François-Xavier Vialard, and Gabriel Peyré. “The unbalanced Gromov Wasserstein distance: Conic formulation and relaxation”. In: Advances in Neural Information Processing Systems 34 (2021), pp. 8766–8779.

    Google Scholar 

  56. Shinji Umeyama. “An eigendecomposition approach to weighted graph matching problems”. In: IEEE transactions on pattern analysis and machine intelligence 10.5 (1988), pp. 695–703.

  57. Aad W Van der Vaart. Asymptotic statistics. Vol. 3. Cambridge university press, 2000.

  58. Titouan Vayer, Rémi Flamary, Nicolas Courty, Romain Tavenard, and Laetitia Chapel. “Sliced Gromov–Wasserstein”. In: Advances in Neural Information Processing Systems 32 (2019).

  59. Titouan Vayer. “A contribution to optimal transport on incomparable spaces”. PhD thesis. 2020. arXiv:2011.04447.

  60. Cédric Villani. Optimal transport: old and new. Vol. 338. Springer, 2008.

  61. Hongteng Xu, Dixin Luo, Ricardo Henao, Svati Shah, and Lawrence Carin. Learning autoencoders with relational regularization. 2020. arXiv: 2002.02913.

  62. Feng Zhou and Fernando De la Torre. “Factorized graph matching”. In: IEEE transactions on pattern analysis and machine intelligence 38.9 (2015), pp. 1774–1789.

  63. Zhengxin Zhang, Youssef Mroueh, Ziv Goldfeld, and Bharath Sriperumbudur. “Cycle consistent probability divergences across different spaces”. In: International Conference on Artificial Intelligence and Statistics. PMLR. 2022, pp. 7257–7285.

Download references

Acknowledgements

We thank Facundo Mémoli and Gabriel Peyré for their helpful remarks. We also thank the anonymous reviewers who substantially helped us improve the clarity of the paper. This work was supported by the Bézout Labex (New Monge Problems), funded by ANR, reference ANR-10-LABX-58.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Théo Dumont.

Additional information

Communicated by Eitan Tadmor.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendix A. Proofs of Lemmas 3.3 and 3.4

In this section, we prove Lemmas 3.3 and 3.4. Let us first recall the statement of Lemma 3.3:

Lemma 3.3

Let \(\mu ,\nu \in \mathcal {P}(E)\) and let \(\psi _1,\psi _2:E\rightarrow F\) be homeomorphisms. Let \({\tilde{c}}:F\times F\rightarrow \mathbb {R}\) and consider the cost \(c(x,y)= c( \psi _1(x),\psi _2(y))\). Then a map is optimal for the cost c between \(\mu \) and \(\nu \) if and only if it is of the form \(\psi _2^{-1}\circ T\circ \psi _1\) with \(T:F\times F\) optimal for the cost \({\tilde{c}}\) between \(\psi _{1\#}\mu \) and \(\psi _{2\#}\nu \).

Proof

Remark that the continuity of \(\psi _1\), \(\psi _2\) and their inverse ensures their measurability. We have the following equalities:

$$\begin{aligned} \mathop {\mathrm {arg\,min}}\limits _{\pi \in \Pi (\mu ,\nu )} \int {\tilde{c}}(\psi _1 (x), \psi _2(y))\mathop {}\!\mathrm d\pi (x,y)&=\mathop {\mathrm {arg\,min}}\limits _{\pi \in \Pi (\mu ,\nu )} \int \tilde{c}( u,v)\mathop {}\!\mathrm d(\psi _1,\psi _2)_\#\pi (u,v) \\&=(\psi _1 ^{-1} ,\psi _2^{-1} )_\#\mathop {\mathrm {arg\,min}}\limits _{{\tilde{\pi }} \in \Pi (\psi _{1\#} \mu , \psi _{2\#}\nu )} \int \tilde{c}(u,v)\mathop {}\!\mathrm d{\tilde{\pi }}(u,v) \end{aligned}$$

since the mapping \((\psi _1 ^{-1},\psi _2^{-1} )_\#\) is a one-to-one correspondence from \(\Pi (\psi _{1\#} \mu , \psi _{2\#}\nu )\) to \(\Pi (\mu ,\nu )\) by bijectivity of \(\psi _1\) and \(\psi _2\). This bijectivity ensures that any optimal deterministic transport plan \({\tilde{\pi }}^\star \) between \(\psi _{1\#}\mu \) and \(\psi _{2\#}\nu \) induces an optimal deterministic transport plan \(\pi ^\star \) between \(\mu \) and \(\nu \), and vice versa. Writing \({\tilde{\pi }}^\star =(\mathop {\textrm{id}}\limits ,T)_\#(\psi _{1\#}\mu )\), this plan \(\pi ^\star \) is given by

$$\begin{aligned} \pi ^\star&= (\psi _1 ^{-1} ,\psi _2^{-1} )_\#{\tilde{\pi }} ^\star \\&=(\psi _1 ^{-1} ,\psi _2^{-1} )_\#(\mathop {\textrm{id}}\limits ,T)_\#\psi _{1\#} \mu \\&=(\mathop {\textrm{id}}\limits ,\psi _2^{-1}\circ T\circ \psi _1)_\#\mu . \square \end{aligned}$$

We now recall the statement of Lemma 3.4:

Lemma 3.4

Let \(h\ge 1\) and \(\mu ,\nu \in \mathcal {P}(\mathbb {R}^h)\) with compact supports and such that \(\mu \ll {\text {Leb}}_h\). Consider the cost \(c(x, y)= -\langle \psi _1(x),\psi _2(y)\rangle \) where \(\psi _1,\psi _2:\mathbb {R}^h\rightarrow \mathbb {R}^h\) are diffeomorphisms. Then there exists a unique optimal transport plan between \(\mu \) and \(\nu \) for the cost c, and it is induced by a map \(t:\mathbb {R}^h\rightarrow \mathbb {R}^h\) of the form \(t=\psi _2^{-1}\circ \nabla f\circ \psi _1\), with \(f:\mathbb {R}^h\rightarrow \mathbb {R}\) convex.

Proof

As \(\psi _{1\#}\mu \) has a density with respect to the Lebesgue measure since \(\psi _1\) is a diffeomorphism and \(\psi _{1\#}\mu \) and \(\psi _{2\#}\nu \) have compact support, Brenier’s theorem states that there exists a unique optimal transport plan between \(\psi _{1\#}\mu \) and \(\psi _{2\#}\nu \) and that it is induced by a map \(\nabla f\), where \(\phi \) is a convex function. Using Lemma 3.3 then yields the result. \(\square \)

Remark A.1

(Hypothesis of Lemma 3.4) In the proof of Lemma 3.4, we only needed (i) \(\psi _1\), \(\psi _2\) and their inverse to be measurable, (ii) \(\psi _{1\#}\mu \) to have a density with respect to the Lebesgue measure, and (iii) \(\psi _{1\#}\mu \) and \(\psi _{2\#}\nu \) to have compact support. Imposing \(\psi _1\) to be a diffeomorphism and \(\psi _2\) to be a homeomorphism ensures both (i) and (ii) and is natural to expect.

Appendix B. Measurable selection of maps in the manifold setting

1.1 B.1 Measurability of set-valued maps

Let X and U be two topological spaces, and let \(\mathcal {B}\) denote the Borel \(\sigma \)-algebra on X. A set-valued map S is a map from X to P(U) (the set of subsets of U) and will be denoted by \(S: X \rightrightarrows U\). The idea is to introduce a notation that is consistent with the case where \(S(x) = \{u\}\) for all x in X, where we want to retrieve the standard case of maps \(X \rightarrow U\). Definitions are taken from Rockafellar and Wets [49], where measurability is studied when \(U = \mathbb {R}^n\). Most results and proofs adapt to a more general setting, in particular when U is a complete Riemannian manifold M, which we shall assume in the following. For the sake of completeness, we provide all the proofs and highlight those that require specific care when replacing \(\mathbb {R}^n\) by a manifold. Of importance for our proofs, we define the pre-image of a set \(B \subset U\) by S,

$$\begin{aligned} S^{-1}(B) = \{ x \in X \mid S(x) \cap B \ne \varnothing \}, \end{aligned}$$

and the domain of S, defined as \(S^{-1}(U)=\{ x \in X \mid S(x) \ne \varnothing \}\). We will often use the following relation: If a set A can be written as \(A = \bigcup _k A_k\), then \(S^{-1}(A) = \bigcup _k S^{-1}(A_k)\). Indeed,

$$\begin{aligned}{} & {} x \in S^{-1}(A) \iff S(x) \cap A \ne \varnothing \iff \exists k,\, S(x) \cap A_k \ne \varnothing \iff \exists k,\,\\{} & {} x \in S^{-1}(A_k) \iff x \in \textstyle \bigcup _k S^{-1}(A_k). \end{aligned}$$

A set-valued map \(S: X \rightrightarrows U\) is said to be measurable if, for any open set \(O \subset U\),

$$\begin{aligned} S^{-1}(O) \in \mathcal {B}. \end{aligned}$$

Note that if S is measurable (as a set-valued map), then its domain must be measurable as well (as an element of \(\mathcal {A}\)). We say that \(S: X \rightrightarrows U\) is closed-valued if S(x) is a closed subset of U for all \(x \in X\).

Proposition B.1

[49, Theorem 14.3.c] A closed-valued map \(S: X \rightrightarrows U\) is measurable if and only if the property “\(S^{-1}(B) \in \mathcal {B}\)” holds either:

  1. (a)

    for all \(B \subset U\) open (the definition);

  2. (b)

    for all \(B \subset U\) compact;

  3. (c)

    for all \(B \subset U\) closed.

Proof

  • (a) \(\Longrightarrow \) (b): For a compact \(B \subset U\), let \(B_k = \{ x \in U \mid d(x,B) < k^{-1} \}\) for \(k \ge 0\) (which is open). Note that

    $$\begin{aligned} x \in S^{-1}(B) \iff S(x) \cap B \ne \varnothing \iff S(x) \cap B_k \ne \varnothing \text { for all { k},} \end{aligned}$$

    because S(x) is a closed set. Hence, \(S^{-1}(B) = \bigcap _k S^{-1}(B_k)\). All the \(S^{-1}(B_k)\) are measurable, and so is \(S^{-1}(B)\) as a countable intersection of measurable sets.

  • (b) \(\Longrightarrow \) (a): Fix O an open set of U. As we assumed U to be a complete separable Riemannian manifold, O can be written as a countable union of compact balls: \(O = \bigcup _n \overline{B(x_n, r_n)}\).

  • (b) \(\Longrightarrow \) (c): Immediate.

  • (c) \(\Longrightarrow \) (b): A closed set B can be obtained as a countable union of compact sets by letting \(B = \bigcup _n B \cap \overline{B(x_0, n)}\) for some \(x_0\). Hence, \(S^{-1}(B) = \bigcup _n S^{-1}(B \cap \overline{B(x_0,n)})\) is in \(\mathcal {B}\).

\(\square \)

Now, we state a result on operations that preserve the measurability of closed–set-valued maps. The proof requires adaptation from the one of [49] because the latter uses explicitly the fact that one can compute Minkowski sums of sets (which may not make sense on a manifold).

Proposition B.2

([49, Proposition 14.11], adapted to the manifold case) Let \(S_1\) and \(S_2: X \rightrightarrows U\) be two measurable closed–set-valued maps. Then

  • \(P: x \mapsto S_1(x) \times S_2(x)\) is measurable as a closed-valued map in \(U \times U\) (equipped with the product topology).

  • \(Q: x \mapsto S_1(x) \cap S_2(x)\) is measurable.

Proof

The first point can be proved in the same spirit as the proof proposed by Rockafellar and Wets. Namely, let \(O'\) be an open set in \(U \times U\). By definition of the product topology, \(O'\) can be obtained as \(\bigcup _n O_1^{(n)} \times O_2^{(n)}\) where \(O_1^{(n)}\) and \(O_2^{(n)}\) are open sets in U. Then \(P^{-1}(O') = \bigcup _n P^{-1}(O_1^{(n)} \times O_2^{(n)})\). Now, observe that

$$\begin{aligned}{} & {} P^{-1}(A \times B) = \{ x\mid S_1(x) \times S_2(x) \in A \times B\} = \{x\mid S_1(x) \in A \text { and } S_2(x) \in B \}\\{} & {} = S_1^{-1}(A) \cap S_2^{-1}(B), \end{aligned}$$

so that finally, \(P^{-1}(O') = \bigcup _n S_1^{-1}(O_1^{(n)}) \cap S_2^{-1}(O_2^{(n)})\) is measurable as a countable union of (finite) intersection of measurable sets (given that \(S_1\) and \(S_2\) are measurable). Note that this does not require \(S_1\) and \(S_2\) to be closed-valued.

Now, let us focus on the second point, which requires more attention. Thanks to Proposition B.1, it is sufficient to show that \(Q^{-1}(C) \in \mathcal {B}\) for any compact set \(C \subset U\). In [49], this is done by writing

$$\begin{aligned} Q^{-1}(C) = \{ x\mid S_1(x) \cap S_2(x) \cap C \ne \varnothing \} = \{x\mid R_1(x) \cap R_2(x) \ne \varnothing \} \\= \{x\mid 0 \in (R_1(x) - R_2(x)) \} = (R_1 - R_2)^{-1}(0), \end{aligned}$$

where \(R_j(x) = S_j(x) \cap C\) (which is also closed-valued), and using the fact that the (Minkowski) difference of measurable closed-valued maps is measurable as well [49, Proposition 14.11.c]. To adapt this idea (we cannot consider Minkowski differences in our setting), we introduce the diagonal \(\Delta = \{(u,u)\mid u \in U\} \subset U \times U\). Now, observe that

$$\begin{aligned} R_1(x) \cap R_2(x) \ne \varnothing \iff (R_1(x) \times R_2(x)) \cap \Delta \ne \varnothing , \end{aligned}$$

that is \(x \in R^{-1}(\Delta )\), where \(R(x) = R_1(x) \times R_2(x)\). Since the maps \(R_1\) and \(R_2\) are measurable closed-valued maps (inherited from \(S_1,S_2\)), so is R according to the previous point. And since \(\Delta \) is closed, \(R^{-1}(\Delta ) = Q^{-1}(C)\) is measurable. \(\square \)

1.2 B.2 Proof of Proposition 2.10

Let us first recall the result.

Proposition 2.10

(Measurable selection of maps, manifold case) Let M be a complete Riemannian manifold and \((B, \Sigma , m)\) a measure space. Consider a measurable function \(B\ni u \mapsto (\mu _u, \nu _u) \in P(M)^2\). Assume that for m-almost every \(u\in B\), \(\mu _u\ll {\text {vol}}_M\) and \(\mu _u\) and \(\nu _u\) have a finite transport cost. Let \(T_u\) denote the (unique by Theorem 1.8) optimal transport map induced by the squared distance cost \(d_M^2\) on M between \(\mu _u\) and \(\nu _u\). Then there exists a measurable function \((u,x)\mapsto T(u,x)\) such that for m-almost every u, \(T(u,x)=T_{u}(x)\), \(\mu _{u}\)-a.e.

The proof is essentially an adaptation of the one of Theorem 1.1 from Fontbona et al. [18], with additional care required since we do not have access to a linear structure on the manifold M. It relies on the measurability of set-valued maps (see [49, Chapters 5 and 14] and “Appendix B.1” for a summary). In the following, we consider a partition \((A_{nk})_n\) of M made of cells with diameter \(\le 2^{-k}\) and such that \((A_{nj})_n\) is a refinement of \((A_{nk})_n\) in sense that each \(A_{nk}\) is itself partitioned by some \(A_{n'j}\). The crucial point regarding measurability is the following proposition.

Proposition B.3

The set \(B_{nk} = \{ (u,x)\mid T_u(x) \in A_{nk} \}\) is measurable.

Its proof relies on a core lemma:

Lemma B.4

Let \(F \subset M\) be a closed set. Then the set \(B_F = \{ (u,x)\mid T_u(x) \in F \}\) is measurable.

The key will be to identify this set as the domain of a measurable set-valued map, see Section “Appendix B.1.”

Proof of Lemma B.4

Observe that

$$\begin{aligned} B_F = \{(u,x)\mid (\{x\} \times F) \cap \mathop {\textrm{gph}}\limits (T_u) \ne \varnothing \}, \end{aligned}$$

where \(\mathop {\textrm{gph}}\limits (T_u) :=\{(x,T_u(x))\mid x \in M\}\) denotes the topological closure of the graph of the optimal transport map \(T_u\) that pushes \(\mu _u\) onto \(\nu _u\). Let \(S_1: (u,x) \mapsto \{x\} \times F\) and \(S_2: (u,x) \mapsto \mathop {\textrm{gph}}\limits (T_u)\), so that \(B_F = \textrm{dom}(S)\), where \(S(x) = S_1(x) \cap S_2(x)\). According to Proposition B.2, given that \(S_1\) and \(S_2\) are closed-valued, if they are measurable, so is S, and so is \(B_F\) as the domain of a measurable map. Establishing the measurability of these two maps can be easily done by adapting the work of [18]. We give details here for the sake of completeness.

  • Measurability of \(S_1\): Let \(O \subset M \times M\) be open. Observe that

    $$\begin{aligned} S_1^{-1}(O) = \{(u,x) \mid \{x\} \times F \cap O \ne \varnothing \} = B \times \{x \mid \{x\} \times F \cap O \ne \varnothing \}. \end{aligned}$$

    Fix \(z \in F\). There exists \(\epsilon > 0\) such that \(B(x,\epsilon ) \times \{z\} \subset O\) (since O is open), and thus \(B(x,\epsilon ) \times F \cap O \ne \varnothing \), proving that there is a neighborhood of x included in \(\{x \mid \{x\} \times F \cap O \ne \varnothing \}\) which is thus open, making in turn \(S_1^{-1}(O)\) open (hence measurable), hence the measurability of \(S_1\).

  • Measurability of \(S_2\): Given that \(u \mapsto (\mu _u,\nu _u)\) is measurable by assumption, and that measurability is preserved by composition, we want to show that

    1. (i)

      the map \(S: (\mu ,\nu ) \mapsto \Pi ^\star (\mu ,\nu )\) (the set of optimal transport plans between \(\mu \) and \(\nu \) for the squared distance cost on M) is measurable, and

    2. (ii)

      the map \(\mathfrak {U}: \pi \in P(M^2) \mapsto \mathop {\textrm{supp}}\limits \pi \) satisfies that \(\mathfrak {U}^{-1}(O)\) is open for any open set \(O \subset P(M^2)\).

    From these two points, we obtain the measurability of \((U \circ S)^{-1}(O)\), thus the measurability of \(S_2\). To get (i), observe first that S is closed-valued, so that it is sufficient to prove that \(S^{-1}(C)\) is measurable for any closed set \(C \subset P(M^2)\) according to Proposition B.1. Let \(C \subset P(M^2)\) be closed. Then, \(S^{-1}(C) = \{(\mu ,\nu )\mid \Pi ^\star (\mu ,\nu ) \cap C \ne \varnothing \}\). Consider a sequence \((\mu _n, \nu _n)_n\) in \(S^{-1}(C)\) that converges to \((\mu ,\nu )\) for the weak topology. Let \(\pi _n \in \Pi ^\star (\mu _n,\nu _n) \cap C\). According to [60, Theorem 5.20], \((\pi _n)_n\) admits a weak limit \(\pi \) in \(\Pi ^\star (\mu ,\nu )\); but since C is closed, \(\pi \in C\), so \((\mu ,\nu ) \in S^{-1}(C)\), which is closed (hence measurable), proving the measurability of S. (ii) simply follows from the fact that

    $$\begin{aligned} \mathfrak {U}^{-1}(O) = \{ \pi \mid \mathop {\textrm{supp}}\limits \pi \cap O \ne \varnothing \} = \{ \pi \mid \pi (O) > 0 \}, \end{aligned}$$

    which is open. Indeed, the Portmanteau theorem shows that if \(\pi _n \rightarrow \pi \) (weakly) and \(\pi _n(O) = 0\), then \(0 = \liminf \pi _n(O) \ge \pi (O) \ge 0\), so \(\pi (O) = 0\). The complementary set of \(U^{-1}(O)\) is closed, that is \(\mathfrak {U}^{-1}(O)\) is open.

\(\square \)

Proof of Proposition B.3

Using that \(A_{nk}\) can be inner-approximated by a sequence of closed set \(F_j \subset A_{nk}\), we obtain a sequence of sets \((B_{F_j})_j\) such that \(\bigcup _j B_{F_j} = A_{nk}\). By Lemma B.4, the \((B_{F_j})_j\) are measurable, so is \(A_{nk}\) as the (countable) union of measurable sets. \(\square \)

We can now prove our main proposition on measurable selections of maps.

Proof of Proposition 2.10

Recall that we assume that \(M = \bigsqcup _n A_{nk}\). For each (nk), select (in a measurable way) an element \(a_{nk}\in A_{nk}\). Then, define the map

$$\begin{aligned} T^{(k)}: (u,x) \mapsto a_{nk},\ \text { where } T_u(x) \in A_{nk}\,. \end{aligned}$$

This map is measurable: for any open \(O \subset M\),

$$\begin{aligned} T^{(k),-1}(O) = \{(u,x) \mid T_u(x) \in A_{nk} \} = B_{nk}, \end{aligned}$$

which is measurable by Proposition B.3. Now, for two maps \(f,g: B \times M \rightarrow M\), let \(D_1\) denotes the natural \(L^1\) distance on M, that is

$$\begin{aligned} D_1(f,g) = \int _B \int _M d\big (f(u,x), g(u,x)\big ) \mathop {}\!\mathrm d\mu _u(x) \mathop {}\!\mathrm dm(u). \end{aligned}$$

This yields a complete metric space [12], and \((T^{(k)})_k\) is a Cauchy sequence for this distance. Indeed, for \(k \le j\) two integers, recall that we assumed that \((A_{nj})_n\) is a refinement of \((A_{nk})_n\), yielding

$$\begin{aligned} D_1(T^{(k)}, T^{(j)})&= \int _{B} \int _M d\big (T^{(k)}(u,x), T^{(j)}(u,x)\big ) \mathop {}\!\mathrm d\mu _u (x) \mathop {}\!\mathrm dm(u) \\&= \int _{B} \int _M \sum _n \sum _{n' :\,A_{n'j} \subset A_{nk}} \mathbb {1}_{B_{n'j}}(u,x) d(a_{nk}, a_{n'j}) \mathop {}\!\mathrm d\mu _u(x) \mathop {}\!\mathrm dm(u) \\&= \int _B \int _M \sum _n \sum _{n' :\,A_{n'j} \subset A_{nk}} d(a_{nk},a_{n'j}) \mathop {}\!\mathrm d\nu _u(A_{n'j}) \mathop {}\!\mathrm dm(u) \\&\le 2^{-k} \end{aligned}$$

where we successively used (i) that for all u, \(\int _{M} \mathbb {1}_{B_{n'j}}(u,x) \mathop {}\!\mathrm d\mu _u(x) = \nu _u(A_{n'j})\) by construction, (ii) that the diameter of the partition \(A_{nk}\) is less than or equal to \(2^{-k}\) and (iii) that \(\nu _u\) and m are probability measures. Note that claim (i) is given by

$$\begin{aligned} (u,x) \in B_{n'j} \iff T_u(x) \in A_{n'j} \iff x \in \mu _u(T_u^{-1}(A_{n'j})) = T_{u \#} \mu _u(A_{n'j})=\nu _u(A_{n'j}). \end{aligned}$$

Now, let T denote the limit of \((T^{(k)})_k\) (which is measurable). It remains to show that \(T(u,x) = T_u(x)\), m-a.e. This can be obtained by proving that

$$\begin{aligned} \int _M g(x) f(T(x,u)) \mathop {}\!\mathrm d\mu _u(x) = \int _M g(x) f(T_u(x)) \mathop {}\!\mathrm d\mu _u(x), \end{aligned}$$

for any pair \(f,g: M \rightarrow \mathbb {R}\) of bounded Lipschitz-continuous functions [57, Lemma 2.24]. As in [18], let \(\Vert f\Vert :=\sup _{x \ne y} \frac{|f(x) - f(y)|}{d(x,y)} + \sup _x |f(x)|\). The difference between these two terms can be bounded using the partition \((A_{nk})_n\). We have for m-a.e. u:

$$\begin{aligned}{} & {} \left| \int _M g(x) f(T_u(x)) \mathop {}\!\mathrm d\mu _u(x) - \int _M g(x) f(T(u,x)) \mathop {}\!\mathrm d\mu _u(x) \right| \\{} & {} \quad \le \left| \int _M g(x) f(T_u(x)) \mathop {}\!\mathrm d\mu _u(x) - \int _M g(x) f(T^{(k)}(u,x)) \mathop {}\!\mathrm d\mu _u(x) \right| \\{} & {} \qquad + \Vert g\Vert \Vert f\Vert \int _M d\big ( T^{(k)}(u,x), T(u,x) \big ) \mathop {}\!\mathrm d\mu _u(x). \end{aligned}$$

Since \(T^{(k)} \rightarrow T\) in \(D_1\), it implies that up to a subsequence, for m-a.e. u, the second term tends to zero:

$$\begin{aligned} \int _M d\big (T^{(k)}(u,x), T(u,x)\big ) \mathop {}\!\mathrm d\mu _u(x)\xrightarrow []{k\rightarrow \infty }0. \end{aligned}$$

To treat the first term and show that it tends to 0 as \(k \rightarrow \infty \) for a subset of B with full m-measure, we write for m-a.e. u:

$$\begin{aligned}&\left| \int _M g(x) f(T_u(x)) \mathop {}\!\mathrm d\mu _u(x) - \int _M g(x) f(T^{(k)}(u,x)) \mathop {}\!\mathrm d\mu _u(x) \right| \\&\quad \le \int _M |g(x)| \big | f(T_u(x)) - f(T^{(k)}(u,x))\big | \mathop {}\!\mathrm d\mu _u(x) \\&\quad \le \Vert g\Vert \Vert f\Vert \int _M d\big (T_u(x),T^{(k)}(u,x)\big ) \mathop {}\!\mathrm d\mu _u(x) \\&\quad \le \Vert g\Vert \Vert f\Vert \ 2^{-k} \sum _n \nu _u(A_{nk})\xrightarrow []{k\rightarrow \infty }0, \end{aligned}$$

which concludes the proof. \(\square \)

Appendix C. Measure disintegration

The following definition and theorem are taken from Ambrosio et al. [1].

Definition C.1

(Measure disintegration) Let \(\mathcal {X}\) and \(\mathcal {Z}\) be two Radon spaces, \(\mu \in \mathcal {P}(\mathcal {X})\) and \(\varphi : \mathcal {X}\rightarrow \mathcal {Z}\) a Borel-measurable function. A family of probability measures \(\{\mu _{u}\}_{u \in \mathcal {Z}} \subset \mathcal {P}(\mathcal {X})\) is a disintegration of \(\mu \) by \(\varphi \) if:

  1. (i)

    the function \(u \mapsto \mu _{u}\) is Borel-measurable;

  2. (ii)

    \(\mu _{u}\) lives on the fiber \(\varphi ^{-1}(u)\), that is for \(\varphi _\#\mu \)-a.e. \(u \in \mathcal {Z}\),

    $$\begin{aligned} \mu _{u}\big (\mathcal {X}\setminus \varphi ^{-1}(u)\big )=0, \end{aligned}$$

    and therefore, \(\mu _{u}(B)=\mu _{u}(B \cap \varphi ^{-1}(u))\) for any Borel \(B\subset \mathcal {X}\);

  3. (iii)

    for every measurable function \(f: \mathcal {X}\rightarrow [0, \infty ]\),

    $$\begin{aligned} \int _{\mathcal {X}} f(x) \mathop {}\!\mathrm d\mu (x)=\int _{\mathcal {Z}} \Big (\int _{\varphi ^{-1}(u)} f(x) \mathop {}\!\mathrm d\mu _{u}(x)\Big ) \mathop {}\!\mathrm d(\varphi _\#\mu )(u). \end{aligned}$$

    In particular, for any Borel \(B \subset \mathcal {X}\), taking f to be the indicator function of B,

    $$\begin{aligned} \mu (B)=\int _{\mathcal {Z}} \mu _{u}(B) \mathop {}\!\mathrm d(\varphi _\#\mu )(u). \end{aligned}$$

Theorem C.2

(Disintegration theorem) Let \(\mathcal {X}\) and \(\mathcal {Z}\) be two Radon spaces, \(\mu \in \mathcal {P}(\mathcal {X})\) and \(\varphi : \mathcal {X}\rightarrow \mathcal {Z}\) a Borel-measurable function. There exists a \(\varphi _\#\mu \)-a.e. uniquely determined family of probability measures \(\{\mu _{u}\}_{u \in \mathcal {Z}} \subset \mathcal {P}(\mathcal {X})\) that provides a disintegration of \(\mu \) by \(\varphi \).

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Dumont, T., Lacombe, T. & Vialard, FX. On the Existence of Monge Maps for the Gromov–Wasserstein Problem. Found Comput Math (2024). https://doi.org/10.1007/s10208-024-09643-0

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s10208-024-09643-0

Keywords

Mathematics Subject Classification

Navigation