Skip to main content
Log in

The Schrödinger problem on the non-commutative Fisher-Rao space

  • Published:
Calculus of Variations and Partial Differential Equations Aims and scope Submit manuscript

Abstract

We present a self-contained and comprehensive study of the Fisher-Rao space of matrix-valued non-commutative probability measures, and of the related Hellinger space. Our non-commutative Fisher-Rao space is a natural generalization of the classical commutative Fisher-Rao space of probability measures and of the Bures-Wasserstein space of Hermitian positive-definite matrices. We introduce and justify a canonical entropy on the non-commutative Fisher-Rao space, which differs from the von Neumann entropy. We consequently derive the analogues of the heat flow, of the Fisher information, and of the dynamical Schrödinger problem. We show the \(\Gamma \)-convergence of the \(\varepsilon \)-Schrödinger problem towards the geodesic problem for the Fisher-Rao space, and, as a byproduct, the strict geodesic convexity of the entropy.

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.

Similar content being viewed by others

References

  1. Ahmed, N.A., Gokhale, D.: Entropy expressions and their estimators for multivariate distributions. IEEE Trans. Inf. Theory 35(3), 688–692 (1989)

    Article  MathSciNet  MATH  Google Scholar 

  2. Albeverio, S., Yasue, K., Zambrini, J.: Euclidean quantum mechanics: analytical approach. Annales de l’IHP Physique théorique 50, 259–308 (1989)

    MathSciNet  MATH  Google Scholar 

  3. Alvarez-Esteban, P. C.,  del Barrio, E., Cuesta-Albertos, J. A., Matrán, C. et al.: Wide consensus aggregation in the wasserstein space. application to location-scatter families. Bernoulli, 24(4A):3147–3179, (2018)

  4. Ambrosio, L., Fusco, N., Pallara, D.: Functions of Bounded Variation and Free Discontinuity Problem. Clarendon Press, Oxford (2000)

    MATH  Google Scholar 

  5. Ambrosio, L., Gigli, N., Savaré, G.: Gradient Flows: in Metric Spaces and in the Space of Probability Measures. Birkhäuser Basel, Basel (2008)

    MATH  Google Scholar 

  6. Arendt, W., Kreuter, M.: Mapping theorems for Sobolev spaces of vector-valued functions. Studia Math. 240(3), 275–299 (2018)

    Article  MathSciNet  MATH  Google Scholar 

  7. Ay, N., Jost, J., Vân Lê, H., Schwachhöfer, L.: Information geometry. Springer, (2017)

  8. Baradat, A., Monsaingeon, L.: Small noise limit and convexity for generalized incompressible flows, schrödinger problems, and optimal transport. Arch. Ration. Mech. Anal. 235(2), 1357–1403 (2020)

    Article  MathSciNet  MATH  Google Scholar 

  9. Barbaresco, F.: Information geometry of covariance matrix: Cartan-Siegel homogeneous bounded domains, Mostow/Berger fibration and Fréchet median. In Matrix information geometry, pp. 199–255. Springer, (2013)

  10. Benamou, J.-D., Brenier, Y.: A computational fluid mechanics solution to the Monge -Kantorovich mass transfer problem. Numer. Math. 84(3), 375–393 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  11. Bengtsson, I., Życzkowski, K.: Geometry of Quantum States: An Introduction to Quantum Entanglement. Cambridge University Press, Cambridge (2017)

    Book  MATH  Google Scholar 

  12. Bhatia, R., Jain, T., Lim, Y.: On the Bures-Wasserstein distance between positive definite matrices. Exposit. Math. 37(2), 165–191 (2019)

    Article  MathSciNet  MATH  Google Scholar 

  13. Biane, P., Voiculescu, D.: A free probability analogue of the Wasserstein metric on the trace-state space. Geomet. Funct. Anal. GAFA 11(6), 1125–1138 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  14. Braides, A.: \(\Gamma \)-convergence for beginners. Oxford Lecture Series in Mathematics and its Applications, vol. 22. Oxford University Press, Oxford (2002)

  15. Brenier, Y., Vorotnikov, D.: On optimal transport of matrix-valued measures. SIAM J. Math. Anal 52(3), 2849–2873 (2020)

    Article  MathSciNet  MATH  Google Scholar 

  16. Bridson, M.R., Haefliger, A.: Metric spaces of non-positive curvature. Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences], vol. 319. Springer-Verlag, Berlin (1999)

  17. Burago, D., Burago, Y., Ivanov, S.: A course in metric geometry. AMS, (2001)

  18. Bures, D.: An extension of Kakutani’s theorem on infinite product measures to the tensor product of semifinite \(w^{\ast } \)-algebras. Trans. Amer. Math. Soc. 135, 199–212 (1969)

    MathSciNet  MATH  Google Scholar 

  19. Carlen, E.: Stochastic mechanics: a look back and a look ahead. Diffus. Quant.Theory Radical Element. Math. 47, 117–139 (2014)

    MathSciNet  Google Scholar 

  20. Carlen, E.A., Maas, J.: An analog of the 2-Wasserstein metric in non-commutative probability under which the fermionic Fokker-Planck equation is gradient flow for the entropy. Commun. Math. Phys. 331(3), 887–926 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  21. Carlen, E.A., Maas, J.: Gradient flow and entropy inequalities for quantum Markov semigroups with detailed balance. J. Funct. Anal. 273(5), 1810–1869 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  22. Carlier, G., Duval, V., Peyré, G., Schmitzer, B.: Convergence of entropic schemes for optimal transport and gradient flows. SIAM J. Math. Anal. 49(2), 1385–1418 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  23. Chen, Y., Gangbo, W., Georgiou, T. T., Tannenbaum, A.: On the matrix Monge-Kantorovich problem. Eur. J. Appl. Math (to appear)

  24. Chen, Y., Georgiou, T.T., Pavon, M.: On the relation between optimal transport and Schrödinger bridges: A stochastic control viewpoint. J. Optim. Theory Appl. 169(2), 671–691 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  25. Chen, Y., Georgiou, T. T., Pavon, M.: Stochastic control liasons: Richard Sinkhorn meets Gaspard Monge on a Schroedinger bridge. arXiv preprint arXiv:2005.10963, (2020)

  26. Chen, Y., Georgiou, T.T., Tannenbaum, A.: Matrix optimal mass transport: a quantum mechanical approach. IEEE Trans. Automat. Control 63(8), 2612–2619 (2018)

    Article  MathSciNet  MATH  Google Scholar 

  27. Chen, Y., Georgiou, T. T., Tannenbaum, A.: Wasserstein geometry of quantum states and optimal transport of matrix-valued measures. In Emerging applications of control and systems theory, Lect. Notes Control Inf. Sci. Proc., pp. 139–150. Springer, Cham (2018)

  28. Chen, Y., Georgiou, T.T., Tannenbaum, A.: Interpolation of matrices and matrix-valued densities: the unbalanced case. Eur. J. Appl. Math. 30(3), 458–480 (2019)

    Article  MathSciNet  MATH  Google Scholar 

  29. Chizat, L., Peyré, G., Schmitzer, B., Vialard, F.-X.: An interpolating distance between optimal transport and Fisher-Rao metrics. Found. Comput. Math. 18(1), 1–44 (2018)

    Article  MathSciNet  MATH  Google Scholar 

  30. Chizat, L., Peyré, G., Schmitzer, B., Vialard, F.-X.: Scaling algorithms for unbalanced optimal transport problems. Math. Comput. 87(314), 2563–2609 (2018)

    Article  MathSciNet  MATH  Google Scholar 

  31. Cuturi, M.: Sinkhorn distances: Lightspeed computation of optimal transport. In Advances in neural information processing systems, pp. 2292–2300, (2013)

  32. Demengel, F., Temam, R.: Convex functions of a measure and applications. India. Univ. Math. J. 33(5), 673–709 (1984)

    Article  MathSciNet  MATH  Google Scholar 

  33. Dittmann, J.: Explicit formulae for the Bures metric. J. Phys. A: Math. Gen. 32(14), 2663 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  34. Duran, A.J., Lopez-Rodriguez, P.: The \(L^p\) space of a positive definite matrix of measures and density of matrix polynomials in \(L^1\). J. Approx. Theory 90(2), 299–318 (1997)

    Article  MathSciNet  MATH  Google Scholar 

  35. Ferrante, A., Masiero, C., Pavon, M.: Time and spectral domain relative entropy: A new approach to multivariate spectral estimation. IEEE Trans. Autom. Control 57(10), 2561–2575 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  36. Ferrante, A., Pavon, M., Ramponi, F.: Hellinger versus Kullback-Leibler multivariable spectrum approximation. IEEE Trans. Autom. Control 53(4), 954–967 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  37. Fonseca, I., Leoni, G.: Modern Methods in the Calculus of Variations: \(L^p\) Spaces. Springer Science & Business Media, (2007)

  38. Goffman, C., Serrin, J.: Sublinear functions of measures and variational integrals. Duke Math. J. 31, 159–178 (1964)

    Article  MathSciNet  MATH  Google Scholar 

  39. Golse, F., Mouhot, C., Paul, T.: On the mean field and classical limits of quantum mechanics. Commun. Math. Phys. 343(1), 165–205 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  40. Itakura, F.: Analysis synthesis telephony based on the maximum likelihood method. In The 6th international congress on acoustics, 1968, pp. 280–292 (1968)

  41. Janati, H., Muzellec, B., Peyré, G., Cuturi, M.: Entropic optimal transport between (unbalanced) gaussian measures has a closed form arXiv:2006.02572. (2020)

  42. Jiang, X., Ning, L., Georgiou, T.T.: Distances and riemannian metrics for multivariate spectral densities. IEEE Trans. Autom. Control 57(7), 1723–1735 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  43. Jordan, R., Kinderlehrer, D., Otto, F.: The variational formulation of the Fokker-Planck equation. SIAM J. Math. Anal. 29(1), 1–17 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  44. Kakutani, S.: On equivalence of infinite product measures. Ann. Math. 2(49), 214–224 (1948)

    Article  MathSciNet  MATH  Google Scholar 

  45. Khesin, B., Lenells, J., Misiołek, G., Preston, S.C.: Geometry of diffeomorphism groups, complete integrability and geometric statistics. Geom. Funct. Anal. 23(1), 334–366 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  46. Khesin, B., Misiolek, G., Modin, K.: Geometric hydrodynamics of compressible fluids. arXiv preprint arXiv:2001.01143, (2020)

  47. Kondratyev, S., Monsaingeon, L., Vorotnikov, D.: A new optimal transport distance on the space of finite Radon measures. Adv. Differ. Equ. 21(11–12), 1117–1164 (2016)

    MathSciNet  MATH  Google Scholar 

  48. Laschos, V., Mielke, A.: Geometric properties of cones with applications on the Hellinger-Kantorovich space, and a new distance on the space of probability measures. J. Funct. Anal. 276(11), 3529–3576 (2019)

    Article  MathSciNet  MATH  Google Scholar 

  49. Léger, F.: A geometric perspective on regularized optimal transport. J. Dyn. Differ. Equ. 31(4), 1777–1791 (2019)

    Article  MathSciNet  MATH  Google Scholar 

  50. Léonard, C.: From the Schrödinger problem to the Monge-Kantorovich problem. J. Funct. Anal. 262(4), 1879–1920 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  51. Léonard, C.: A survey of the Schrödinger problem and some of its connections with optimal transport. Discrete Continuous Dyn. Syst. A. 34(4), 1533–1574 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  52. Liero, M., Mielke, A., Savaré, G.: Optimal transport in competition with reaction: the Hellinger-Kantorovich distance and geodesic curves. SIAM J. Math. Anal. 48(4), 2869–2911 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  53. Liero, M., Mielke, A., Savaré, G.: Optimal entropy transport problems and a new Hellinger-Kantorovich distance between positive measures. Invent. math. 211(3), 969–1117 (2018)

    Article  MathSciNet  MATH  Google Scholar 

  54. Lott, J., Villani, C.: Ricci curvature for metric-measure spaces via optimal transport. Annals of Mathematics, pp. 903–991 (2009)

  55. Malagò, L., Montrucchio, L., Pistone, G.: Wasserstein Riemannian geometry of Gaussian densities. Inf. Geom. 1(2), 137–179 (2018)

    Article  MathSciNet  MATH  Google Scholar 

  56. Mittnenzweig, M., Mielke, A.: An entropic gradient structure for Lindblad equations and couplings of quantum systems to macroscopic models. J. Stat. Phys. 167(2), 205–233 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  57. Modin, K.: Geometry of matrix decompositions seen through optimal transport and information geometry. J. Geomet. Mech. 9(3), 335–390 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  58. Monsaingeon, L., Tamanini, L., Vorotnikov, D.: The dynamical Schrödinger problem in abstract metric spaces. In preparation (2020)

  59. Musicus, B., Johnson, R.: Multichannel relative-entropy spectrum analysis. IEEE Trans. Acoust. Speech Signal Process. 34(3), 554–564 (1986)

    Article  MATH  Google Scholar 

  60. Ning, L., Georgiou, T.T., Tannenbaum, A.: On matrix-valued Monge-Kantorovich optimal mass transport. IEEE Trans. Autom. Control 60(2), 373–382 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  61. Otto, F.: The geometry of dissipative evolution equations: the porous medium equation. Commun. Partial Differ. Equ. 26(1–2), 101–174 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  62. Peyré, G., Chizat, L., Vialard, F.-X., Solomon, J.: Quantum entropic regularization of matrix-valued optimal transport. Eur. J. Appl. Math. 30(6), 1079–1102 (2019)

    Article  MathSciNet  MATH  Google Scholar 

  63. Peyré, G., Cuturi, M., et al.: Computational optimal transport: With applications to data science. Foundations and Trends® in Machine Learning, 11(5-6):355–607, (2019)

  64. Powers, R.T., Størmer, E.: Free states of the canonical anticommutation relations. Commun. Math. Phys. 16(1), 1–33 (1970)

    Article  MathSciNet  MATH  Google Scholar 

  65. Sakai, S.: \(C^*\)-algebras and \(W^*\)-algebras. Springer-Verlag, New York-Heidelberg, 1971. Ergebnisse der Mathematik und ihrer Grenzgebiete, Band 60

  66. Santambrogio, F.: Optimal transport for applied mathematicians. Birkhäuser/Springer, (2015)

  67. Séjourné, T., Feydy, J., Vialard, F.-X., Trouvé, A., Peyré, G.: Sinkhorn divergences for unbalanced optimal transport. arXiv preprint arXiv:1910.12958, (2019)

  68. Takatsu, A.: Wasserstein geometry of Gaussian measures. Osaka J. Math. 48(4), 1005–1026 (2011)

    MathSciNet  MATH  Google Scholar 

  69. Tao, T.: Topics in random matrix theory. American Mathematical Soc., (2012)

  70. Uhlmann, A.: The “transition probability” in the state space of a \(^*\)-algebra. Rep. Math. Phys. 9(2), 273–279 (1976)

    Article  MathSciNet  MATH  Google Scholar 

  71. Uhlmann, A.: The metric of Bures and the geometric phase. In Groups and related Topics, Gielerak et al. (eds.), pp. 267–274. Springer (1992)

  72. Villani, C.: Topics in optimal transportation. American Mathematical Soc., (2003)

  73. Villani, C.: Optimal transport: old and new. Springer Science & Business Media, (2008)

  74. Yasue, K.: Stochastic calculus of variations. J. Funct. Anal. 41(3), 327–340 (1981)

    Article  MathSciNet  MATH  Google Scholar 

  75. Zambrini, J.-C.: Variational processes and stochastic versions of mechanics. J. Math. Phys. 27(9), 2307–2330 (1986)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

Credit is due to Aymeric Baradat for our construction of recovery sequences, which was improved and adapted from [8]. We are very grateful to the anonymous referee for pointing out the link with the theory of \(C^*\)-algebras, cf. Remark 5.2. LM wishes to thank Jean-Claude Zambrini for numerous and fruitful discussions on the Schrödinger problem, and acknowledges support from the Portuguese Science Foundation through FCT project PTDC/MAT-STA/28812/2017 SchröMoka. DV was partially supported by the FCT projects UID/MAT/00324/2020 and PTDC/MAT-PUR/28686/2017.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Dmitry Vorotnikov.

Additional information

Communicated by J. Jost.

Publisher's Note

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

Appendix A: Some technical lemmas and postponed proofs

Appendix A: Some technical lemmas and postponed proofs

Lemma A.1

(Constant-speed reparametrization) Let \((G_t,U_t)\) be a curve connecting \(G_0,G_1\) with finite energy

$$\begin{aligned} E[G,U]=\int _0^1\int _\Omega \mathrm {d}G_t U_t:U_t\mathrm {d}t<+\infty . \end{aligned}$$

Then there exists a curve \(({\check{G}}_t,{\check{U}}_t)_{t\in [0,1]}\) connecting \({\check{G}}_0=G_0\) to \({\check{G}}_1=G_1\) with

$$\begin{aligned} \Vert {\check{U}}_t\Vert _{L^2(\mathrm {d}{\check{G}}_t)}\equiv cst ~\text {and}~ E[{\check{G}},{\check{U}}] = \left( \int _0^1\Vert U_t\Vert _{L^2(\mathrm {d}G_t)}\mathrm {d}t\right) ^2 \le E[G,U] \end{aligned}$$

with strict inequality unless \(\Vert U_t\Vert _{L^2(\mathrm {d}G_t)}\) is constant in time.

Proof

The argument is fairly standard (see e.g. [47, lemma 5.3] or [5, lemma 1.1.4]) hence we only sketch the idea. Consider the change of time variable

$$\begin{aligned} {\mathsf {s}}(t):=\int _0^t \Vert U_\tau \Vert _{L^2(\mathrm {d}G_\tau )}\mathrm {d}\tau , \text {and}~{\mathsf {t}}(s):= \min \{t\in [0,1]:\quad \mathsf s(t)=s\} \end{aligned}$$

(\({\mathsf {t}}\) is the left inverse of \({\mathsf {s}}\)). Setting

$$\begin{aligned} {\widetilde{G}}_s:=G_{{\mathsf {t}}(s)}, \qquad \widetilde{U}_s:=\frac{U_{{\mathsf {t}}(s)}}{\Vert U_{{\mathsf {t}}(s)}\Vert _{L^2(\widetilde{G}_s)}} \end{aligned}$$

gives an admissible path connecting \({\widetilde{G}}_0=G_0\) to \({\widetilde{G}}_L=G_1\) in time \(s\in [0,L]\) with

$$\begin{aligned} L={\mathsf {s}}(1)=\int _0^1 \Vert U_\tau \Vert _{L^2(\mathrm {d}G_\tau )}\mathrm {d}\tau \end{aligned}$$

and unit speed \(\Vert {\widetilde{U}}_s\Vert _{L^2(\mathrm {d}{\widetilde{G}}_s)}\equiv 1\). This is clear at least formally from the chain-rule

$$\begin{aligned}&\partial _s{\widetilde{G}}_s = \frac{\mathrm {d}{\mathsf {t}}}{\mathrm {d}s}(s) \partial _tG_t({\mathsf {t}}(s)) \\&\quad = \frac{1}{\frac{\mathrm {d}{\mathsf {s}}}{\mathrm {d}t}({\mathsf {t}}(s))}\left( G_{{\mathsf {t}}(s)} U_{\mathsf t(s)}\right) ^{Sym}\\&\quad = \frac{1}{ \Vert U_{\mathsf t(s)}\Vert _{L^2(\mathrm {d}G_{{\mathsf {t}}(s)})}}\left( G_{{\mathsf {t}}(s)} U_{{\mathsf {t}}(s)}\right) ^{Sym} \\&\quad =(\widetilde{G}_s {\widetilde{U}}_s)^{Sym} \end{aligned}$$

and can be made rigorous since the denominator \(\Vert U_{\mathsf t(s)}\Vert _{L^2(\mathrm {d}G_{{\mathsf {t}}(s)})}\) only vanishes at the discontinuity points of \({\mathsf {t}}(s)\), which are countable (being \({\mathsf {t}}\) monotone nondecreasing) and therefore \(\mathrm {d}s\)-negligible. Scaling \(t=Ls\)

$$\begin{aligned} ({\check{G}}_t,\check{U}_t):=({\widetilde{G}}_{tL},L{\widetilde{U}}_{tL}) \qquad \text{ for } t\in [0,1] \end{aligned}$$

back to the unit interval and noticing that \(\Vert \check{U}_t\Vert _{L^2(\mathrm {d}{\check{G}}_t)}\equiv L\) is constant, we get

$$\begin{aligned}&E[{\check{G}},{\check{U}}] = \int _0^1 \Vert \check{U}_t\Vert ^2_{L^2(\mathrm {d}{\check{G}}_t)} \,\mathrm {d}t = \int _0^1 L^2\mathrm {d}t=L^2 \\&\quad =\left( \int _0^1\Vert U_\tau \Vert _{L^2(\mathrm {d}G_\tau )}\mathrm {d}\tau \right) ^2 \le \int _0^1\Vert U_\tau \Vert _{L^2(\mathrm {d}G_\tau )}^2\mathrm {d}\tau = E[G,U] \end{aligned}$$

as desired. (The inequality is strict unless \((G,U)\) has constant speed as in our statement.) \(\square \)

Lemma A.2

(Refined Banach-Alaoglu [47]) Let \((X,\Vert \cdot \Vert )\) be a separable normed vector space. Assume that there exists a sequence of seminorms \(\{\Vert \cdot \Vert _k\}\) (\(k=0,1,2,\dots \)) on X such that for every \(x\in X\) one has

$$\begin{aligned} \Vert x\Vert _k\le C \Vert x\Vert \end{aligned}$$

with a constant C independent of kx, and

$$\begin{aligned} \Vert x\Vert _k\underset{k\rightarrow \infty }{\rightarrow } \Vert x\Vert _0. \end{aligned}$$

Let \(\varphi _k\) (\(k=1,2,\dots \)) be a uniformly bounded sequence of linear continuous functionals on \((X,\Vert \cdot \Vert _k)\), resp., in the sense that

$$\begin{aligned} c_k:=\Vert \varphi _k\Vert _{(X,\Vert \cdot \Vert _k)^*}\le C. \end{aligned}$$

Then the sequence \(\{\varphi _k\}\) admits a converging subsequence \(\varphi _{k_n}\rightarrow \varphi _0\) in the weak-\(*\) topology of \(X^*\), and

$$\begin{aligned} \Vert \varphi _0\Vert _{(X,\Vert \cdot \Vert _0)^*}\le c_0:=\liminf \limits _k c_{k}. \end{aligned}$$
(A.1)

Lemma A.3

(Refined Arzelà-Ascoli [5, 15]) Let \((X,\varrho )\) be a metric space. Assume that there exists a Hausdorff topology \(\sigma \) on X such that \(\varrho \) is sequentially lower semicontinuous with respect to \(\sigma \). Let \((x^k)_t\), \(t\in [0,1]\), be a sequence of curves lying in a common \(\sigma \)-sequentially compact set \(K\subset X\). Let it be equicontinuous in the sense that there exists a symmetric continuous function \(\omega :[0,1]\times [0,1]\rightarrow \mathbb {R}_+\), \(\omega (t,t)=0\), such that

$$\begin{aligned} \varrho ((x^k)_t,(x^k)_{\bar{t}})\le \omega (t,{\bar{t}}). \end{aligned}$$
(A.2)

for all \(t,{\bar{t}}\in [0,1]\). Then there exists a \(\varrho \)-continuous curve \(x_t\) such that

$$\begin{aligned} \varrho (x_t,x_{{\bar{t}}})\le \omega (t,{\bar{t}}), \end{aligned}$$
(A.3)

and (up to a not relabelled subsequence)

$$\begin{aligned} (x^k)_t\rightarrow x_t \end{aligned}$$
(A.4)

for all \(t\in [0,1]\) in the topology \(\sigma \).

Proof of Proposition 3.4

First of all, let us show that the right-hand side is always finite and that the minimum is always attained. For any fixed \(A_1\in {\mathcal {H}^+}\) it is easy to check that \((A_t,U_t):=(t^2A_1,\frac{2}{t} I)\) gives an admissible path connecting \(A_0=0\) to \(A_1\) with finite energy. In particular any two matrices \(A_0,A_1\in {\mathcal {H}^+}\) can be connected through zero as \(A_0\leadsto 0\leadsto A_1\) with finite cost, thus the problem is proper. For fixed \(A_0,A_1\) consider now a minimizing sequence \((A^n_t,U^n_t)_{t\in [0,1]}\). Note that our Lemma 4.2 applies in particular when \(\Omega =\{x\}\) is a one-point space, which gives here equicontinuity and pointwise relative compactness in the form

$$\begin{aligned} |A^n_t-A^n_s|_2\le C |t-s|^\frac{1}{2}, \qquad m^n_t=\hbox {tr}A^n_t\le M, \qquad s,t\in [0,1] \end{aligned}$$

uniformly in n. By the classical Arzelá-Ascoli theorem we get, up to extraction of a subsequence if needed,

$$\begin{aligned} A^n\rightarrow A \qquad \text{ uniformly } \text{ in } C([0,1];{\mathcal {H}^+}). \end{aligned}$$

This immediately shows that the matrix-valued measure \(\mu ^n:=A^n_t\mathrm {d}t\rightarrow A_t dt=:\mu \) at least weakly-\(*\) on \(\mathbb {H}^+(0,1)\). Because \( \Vert U^n\Vert _{L^2(\mathrm {d}\mu ^n)}^2=\int _0^1 A^n_tU^n_t:U^n_t\,\mathrm {d}t\le C \), an easy application of our Banach-Alaoglu variant (lemma A.2) in varying \(L^2(\mathrm {d}\mu ^n)\) spaces gives a limit \(U\in L^2(\mathrm {d}\mu )\) with

$$\begin{aligned} \int _0^1 A_t U_t:U_t\mathrm {d}t \le \liminf \limits _{n\rightarrow \infty } \int _0^1 A^n_tU^n_t:U^n_t\,\mathrm {d}t \end{aligned}$$

and such that \(\int _0^1 A^n U^n:V\,\mathrm {d}t\rightarrow \int _0^1 A U:V\,\mathrm {d}t\) for any reasonably smooth test function V. This shows that this limit \((A_t,U_t)_{t\in [0,1]}\) is an admissible curve joining \(A_0,A_1\) with energy \( E[A,U]\le \liminf E[A^n,U^n] \) and this pair is therefore a minimizer.

In order to identify now the left-hand side and the right-hand side of (3.5) we proceed in two steps.

Step 1: assume that \(A_0,A_1\in \mathcal {H}^{++}(d)\) are positive definite, and let \({\check{A}}_{0,1}:={\mathfrak {r}}(A_{0,1})\) be the corresponding real extensions as defined in (3.8). From Proposition 3.6 there holds \(d_B^2(A_0,A_1)=\frac{1}{2} W^2_2({\mathcal {N}}(\check{A}_0),{\mathcal {N}}({\check{A}}_1))\). In this real setting it is known [68, Prop. A] that

$$\begin{aligned} W^2_2({\mathcal {N}}({\check{A}}_0),{\mathcal {N}}({\check{A}}_1)) = \frac{1}{4} \min \limits _{{\check{A}}\in {\mathcal {A}}({\check{A}}_0,{\check{A}}_1)} \int _0^1 {\check{A}}_t {\check{U}}_t:{\check{U}}_t\,\mathrm {d}t, \end{aligned}$$

where the infimum runs of course over real pairs \(\frac{d\check{A}_t}{dt}=({\check{A}}_t {\check{U}}_t)^{Sym}\) with \({\check{U}}_t\in \mathcal S(2d)\). Complexifying back \(({\check{A}},{\check{U}})\leadsto (A,U)\) gives the result.

Step 2: if now either \(A_0\) or \(A_1\) are only semi-definite we approximate \(A^n_0:=A_0+\frac{1}{n} I\rightarrow A_0\) and \(A^n_1:=A_1+\frac{1}{n} I\rightarrow A_1\). Clearly \(A^n_{0,1}\in \mathcal {H}^{++}\) are positive-definite, hence step 1 applies.

Note from (3.4) that the left-hand side of (3.5) is of course continuous in \(A_0,A_1\), hence it suffices to show that the optimal value in the right-hand side is continuous for this particular choice of \(A^n_{0,1}\rightarrow A_{0,1}\). Observe that

$$\begin{aligned} {\mathcal {J}}^*(A_0,A_1):=\frac{1}{4} \min _{{\mathcal {A}}(A_0,A_1)}\int _0^1 A_t U_t :U_t \mathrm {d}t \end{aligned}$$

is a well-defined function of \(A_0,A_1\in {\mathcal {H}^+}\), and we proved earlier that there always exists a minimizer. Arguing precisely as for the existence of the minimizers, cf. also Lemma , it is easy to prove that \({\mathcal {J}}^*\) is lower semi-continuous in both arguments. Indeed, up to a subsequence, any sequence \((A^n_t,U^n_t)_{t\in [0,1]}\) of minimizers in \({\mathcal {J}}^*(A_0^n,A_1^n)\) converges to an admissible candidate \((A_t,U_t)_{t\in [0,1]}\) connecting \(A_0,A_1\), hence \({\mathcal {J}}^*(A_0,A_1)\le E[A,U]\le \liminf E[A^n,U^n]=\liminf {\mathcal {J}}^*(A^n_0,A^n_1)\) (regardless of the particular form of \(A_0^n,A_1^n\)).

In order to establish the upper continuity, let \(A_0=R D_0 R^*\) be a spectral decomposition of \(A_0\) and note that obviously \(A^n_0=A_0+\frac{1}{n} I=R(D_0+\frac{1}{n} I)R^*\). Since \(A^n_0\) and \(A_0\) commute it is easy to check that

$$\begin{aligned} A^n_{0t}:=R\left[ (1-t)\sqrt{D_0}+t\sqrt{D_0+\frac{1}{n} I}\right] ^2R^*, \qquad U^n_{0t}:=2 R\left( \sqrt{D_0+\frac{1}{n} I}-\sqrt{D_0}\right) R^*\sqrt{A_{0t}^{-1}} \end{aligned}$$

defines an admissible path \((A^n_{0t})_{t\in [0,1]}\) between \(A_0\) and \(A^n_0\). Moreover, a straightforward computation shows that the corresponding energy is

$$\begin{aligned} \frac{1}{4} \int _0^1A^n_{0t} U^n_{0t}:U^n_{0t}\,\mathrm {d}t = \left| \sqrt{D_0}-\sqrt{D_0+I/n}\right| ^2_2 \xrightarrow [n\rightarrow \infty ]{} 0. \end{aligned}$$
(A.5)

(Actually this path is exactly the Bures geodesic between \(A_0,A_0^n\).) A similar construction yields a path \((A^n_{1t})_{t\in [0,1]}\) connecting \(A_1^n\) to \(A_1\) with cost

$$\begin{aligned} \frac{1}{4} \int _0^1A^n_{1t} U^n_{1t}:U^n_{1t}\,\mathrm {d}t = \left| \sqrt{D_1}-\sqrt{D_1+I/n}\right| ^2_2 \xrightarrow [n\rightarrow \infty ]{} 0, \end{aligned}$$
(A.6)

where \(D_1\) is the spectral decomposition of \(A_1=QD_1Q^*\). Now pick a minimizer \(({\tilde{A}}_t,{\tilde{U}}_t)_{t\in [0,1]}\) in the definition of \({\mathcal {J}}^*(A_0,A_1)\) and fix a small \(\theta _n\in (0,1)\) to be determined shortly. Rescaling in time and concatenating the paths \(A_0^n\leadsto A_0\leadsto A_1\leadsto A_1^n\) in the intervals \(t\in [0,\theta _n]\), \(t\in [\theta _n,1-\theta _n]\), and \(t\in [1-\theta _n,1]\), respectively, we obtain an admissible path \(({\hat{A}}^n,{\hat{U}}^n)\) from \(A_0^n\) to \(A_1^n\) whose energy is bounded as

$$\begin{aligned}&4 {\mathcal {J}}^*(A_0^n,A_1^n)\le \int _0^1{\hat{A}}_t^n{\hat{U}}^n_t:{\hat{U}}^n_t\,\mathrm {d}t \\&\quad = \frac{1}{\theta _n} \int _0^1A^n_{0t} U^n_{0t}:U^n_{0t}\,\mathrm {d}t +\frac{1}{1-2\theta _n} \int _0^1 {\tilde{A}}_t{\tilde{U}}_t:\tilde{U}_t\,\mathrm {d}t +\frac{1}{\theta _n} \int _0^1A^n_{1t} U^n_{1t}:U^n_{1t}\,\mathrm {d}t. \end{aligned}$$

By (A.5), (A.6) we see that the first and third integrals in the right-hand side tend to zero, while the second integral is exactly \(4 \mathcal J^*(A_0,A_1)\) by definition of \(({\tilde{A}},{\tilde{U}})\). Choosing \(\theta _n\rightarrow 0\) sufficiently slowly and taking \(\limsup \) in the previous inequality gives

$$\begin{aligned} \limsup _{n \rightarrow \infty } 4{\mathcal {J}}^*(A_0^n,A_1^n)\le 0+\lim _{n\rightarrow \infty }\frac{1}{1-2\theta _n}\int _0^1 {\tilde{A}}_t{\tilde{U}}_t:{\tilde{U}}_t\,\mathrm {d}t + 0 = 4\mathcal J^*(A_0,A_1) \end{aligned}$$

and the proof is complete. \(\square \)

Proof of Proposition 3.8

Let \(\rho _0={\mathcal {N}}(A_0)\) and \(\rho _1={\mathcal {N}}(A_1)\). By [51, Thm. 3.3 and Thm. 3.4], if one could write the (fg) transform

$$\begin{aligned} {\left\{ \begin{array}{ll} \rho _0=f_0g_0\\ \rho _1=f_1g_1 \end{array}\right. } \end{aligned}$$
(A.7)

for \((f_t)_{t\in [0,1]}\) a forward solution of the heat equation with initial datum \(f_0\) and \((g_t)_{t\in [0,1]}\) a backward solution with terminal datum \(g_1\), then the solution of (3.16) would be given by

$$\begin{aligned} \rho _t=f_tg_t. \end{aligned}$$

Since the product of Gaussian distributions is Gaussian, it is legitimate to try and solve for \(f_0={\mathcal {N}}(B_0)\) and \(g_1={\mathcal {N}}(C_1)\) as Gaussians. Since the heat flow for Gaussians is explicitly given by (3.14) we see that the corresponding forward and backward solutions read

$$\begin{aligned} {\left\{ \begin{array}{ll} f_t={\mathcal {N}}(B_t) &{} \qquad \text{ with } B_t=B_0+2t I \\ g_t={\mathcal {N}}(C_t) &{} \qquad \text{ with } C_t=C_1+2(1-t) I \end{array}\right. } \end{aligned}$$

Exploiting the algebraic product rule \({\mathcal {N}}(B)\times {\mathcal {N}}(C)=\mathcal N\left( [B^{-1}+C^{-1}]^{-1}\right) \), we see that, given \(A_0,A_1\) the Schrödinger system (A.7) is equivalent to solving

$$\begin{aligned} {\left\{ \begin{array}{ll} {\mathcal {N}}(A_0)={\mathcal {N}}(B_0){\mathcal {N}}(C_0)\\ {\mathcal {N}}(A_1)={\mathcal {N}}(B_1){\mathcal {N}}(C_1) \end{array}\right. } \qquad \Leftrightarrow \qquad {\left\{ \begin{array}{ll} A_0^{-1}= B_0^{-1}+[C_1+2I]^{-1} \\ A_1^{-1} = [B_0+2I]^{-1}+ C_1^{-1} \end{array}\right. } \end{aligned}$$

in \(B_0,C_1\in \mathcal S^{++}\). It is then a simple exercise to check that this system has a unique solution, and in particular the covariance \(A_t\) of \(\rho _t=f_tg_t\) is fully determined by

$$\begin{aligned} A_t^{-1}=[B_0+2tI]^{-1} +[C_1+2(1-t)I]^{-1}. \end{aligned}$$

\(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Monsaingeon, L., Vorotnikov, D. The Schrödinger problem on the non-commutative Fisher-Rao space. Calc. Var. 60, 14 (2021). https://doi.org/10.1007/s00526-020-01871-w

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00526-020-01871-w

Mathematics Subject Classification

Navigation