Local semiconvexity of Kantorovich potentials on non-compact manifolds

Similar documents
Optimal transportation on non-compact manifolds

The optimal partial transport problem

C 1 regularity of solutions of the Monge-Ampère equation for optimal transport in dimension two

REGULARITY OF MONOTONE TRANSPORT MAPS BETWEEN UNBOUNDED DOMAINS

The Monge problem on non-compact manifolds

SEPARABILITY AND COMPLETENESS FOR THE WASSERSTEIN DISTANCE

Optimal transportation and action-minimizing measures

Spaces with Ricci curvature bounded from below

Introduction to Optimal Transport Theory

FULL CHARACTERIZATION OF OPTIMAL TRANSPORT PLANS FOR CONCAVE COSTS

PARTIAL REGULARITY OF BRENIER SOLUTIONS OF THE MONGE-AMPÈRE EQUATION

Optimal Transport for Data Analysis

LECTURE 15: COMPLETENESS AND CONVEXITY

NECESSARY AND SUFFICIENT CONDITIONS FOR CONTINUITY OF OPTIMAL TRANSPORT MAPS ON RIEMANNIAN MANIFOLDS

Weak KAM pairs and Monge-Kantorovich duality

Generic Aubry sets on surfaces

A few words about the MTW tensor

Logarithmic Sobolev Inequalities

Mañé s Conjecture from the control viewpoint

A NOTION OF NONPOSITIVE CURVATURE FOR GENERAL METRIC SPACES

Optimal Transportation. Nonlinear Partial Differential Equations

LARGE DEVIATIONS OF TYPICAL LINEAR FUNCTIONALS ON A CONVEX BODY WITH UNCONDITIONAL BASIS. S. G. Bobkov and F. L. Nazarov. September 25, 2011

The Minimum Speed for a Blocking Problem on the Half Plane

arxiv: v2 [math.ap] 23 Apr 2014

On a Class of Multidimensional Optimal Transportation Problems

Transport Continuity Property

arxiv: v2 [math.mg] 10 Apr 2015

A generic property of families of Lagrangian systems

EXAMPLE OF A FIRST ORDER DISPLACEMENT CONVEX FUNCTIONAL

Generalized Ricci Bounds and Convergence of Metric Measure Spaces

Alexandrov meets Lott-Villani-Sturm

On the regularity of solutions of optimal transportation problems

Topology of the set of singularities of a solution of the Hamilton-Jacobi Equation

M. Ledoux Université de Toulouse, France

A STRATEGY FOR NON-STRICTLY CONVEX TRANSPORT COSTS AND THE EXAMPLE OF x y p IN R 2

Stepanov s Theorem in Wiener spaces

A sufficient condition for observability of waves by measurable subsets

2 (Bonus). Let A X consist of points (x, y) such that either x or y is a rational number. Is A measurable? What is its Lebesgue measure?

Regularity for the optimal transportation problem with Euclidean distance squared cost on the embedded sphere

Metric Spaces and Topology

Uniqueness of the solution to the Vlasov-Poisson system with bounded density

BRUNN MINKOWSKI AND ISOPERIMETRIC INEQUALITY IN THE HEISENBERG GROUP

ON THE VOLUME MEASURE OF NON-SMOOTH SPACES WITH RICCI CURVATURE BOUNDED BELOW

Geometric and isoperimetric properties of sets of positive reach in E d

A CHARACTERIZATION OF STRICT LOCAL MINIMIZERS OF ORDER ONE FOR STATIC MINMAX PROBLEMS IN THE PARAMETRIC CONSTRAINT CASE

Moment Measures. Bo az Klartag. Tel Aviv University. Talk at the asymptotic geometric analysis seminar. Tel Aviv, May 2013

Recent developments in elliptic partial differential equations of Monge Ampère type

WEAK CURVATURE CONDITIONS AND FUNCTIONAL INEQUALITIES

OPTIMALITY CONDITIONS FOR GLOBAL MINIMA OF NONCONVEX FUNCTIONS ON RIEMANNIAN MANIFOLDS

MEAN CURVATURE FLOW OF ENTIRE GRAPHS EVOLVING AWAY FROM THE HEAT FLOW

REMOVABLE SINGULARITIES FOR BOUNDED p-harmonic AND QUASI(SUPER)HARMONIC FUNCTIONS ON METRIC SPACES

arxiv: v1 [math.ca] 20 Sep 2010

On the Intrinsic Differentiability Theorem of Gromov-Schoen

Sobolev regularity for the Monge-Ampère equation, with application to the semigeostrophic equations

SMOOTH OPTIMAL TRANSPORTATION ON HYPERBOLIC SPACE

A planning problem combining optimal control and optimal transport

Topological properties

Some Properties of the Augmented Lagrangian in Cone Constrained Optimization

Continuity of optimal transport maps and convexity of injectivity domains on small deformations of S 2

Spaces with Ricci curvature bounded from below

1 Cheeger differentiation

Introduction to optimal transport

Calculus of Variations

Inverse Brascamp-Lieb inequalities along the Heat equation

On the Hausdorff Dimension of the Mather Quotient

MATH 722, COMPLEX ANALYSIS, SPRING 2009 PART 5

Remarks on the Gauss-Green Theorem. Michael Taylor

On Shalom Tao s Non-Quantitative Proof of Gromov s Polynomial Growth Theorem

PACKING DIMENSIONS, TRANSVERSAL MAPPINGS AND GEODESIC FLOWS

3 (Due ). Let A X consist of points (x, y) such that either x or y is a rational number. Is A measurable? What is its Lebesgue measure?

MINIMAL GRAPHS PART I: EXISTENCE OF LIPSCHITZ WEAK SOLUTIONS TO THE DIRICHLET PROBLEM WITH C 2 BOUNDARY DATA

Integral Jensen inequality

On fine differentiability properties of horizons and applications to Riemannian geometry

and BV loc R N ; R d)

AN ELEMENTARY PROOF OF THE TRIANGLE INEQUALITY FOR THE WASSERSTEIN METRIC

arxiv: v1 [math.ap] 30 Jul 2015

ON A MEASURE THEORETIC AREA FORMULA

ON THE VOLUME MEASURE OF NON-SMOOTH SPACES WITH RICCI CURVATURE BOUNDED BELOW

APPLICATIONS OF DIFFERENTIABILITY IN R n.

ANSWER TO A QUESTION BY BURR AND ERDŐS ON RESTRICTED ADDITION, AND RELATED RESULTS Mathematics Subject Classification: 11B05, 11B13, 11P99

Robustness for a Liouville type theorem in exterior domains

PATH FUNCTIONALS OVER WASSERSTEIN SPACES. Giuseppe Buttazzo. Dipartimento di Matematica Università di Pisa.

Characterization of Self-Polar Convex Functions

i=1 β i,i.e. = β 1 x β x β 1 1 xβ d

On duality theory of conic linear problems

MAT 570 REAL ANALYSIS LECTURE NOTES. Contents. 1. Sets Functions Countability Axiom of choice Equivalence relations 9

The local equicontinuity of a maximal monotone operator

From now on, we will represent a metric space with (X, d). Here are some examples: i=1 (x i y i ) p ) 1 p, p 1.

SOME PROPERTIES ON THE CLOSED SUBSETS IN BANACH SPACES

Parcours OJD, Ecole Polytechnique et Université Pierre et Marie Curie 05 Mai 2015

A REPRESENTATION FOR THE KANTOROVICH RUBINSTEIN DISTANCE DEFINED BY THE CAMERON MARTIN NORM OF A GAUSSIAN MEASURE ON A BANACH SPACE

1 Topology Definition of a topology Basis (Base) of a topology The subspace topology & the product topology on X Y 3

ON A CLASS OF NONSMOOTH COMPOSITE FUNCTIONS

Problem Set 2: Solutions Math 201A: Fall 2016

Morse-Sard theorem for Clarke critical values

Math General Topology Fall 2012 Homework 11 Solutions

Optimal Transport for Data Analysis

Optimal transportation on the hemisphere

Behaviour of Lipschitz functions on negligible sets. Non-differentiability in R. Outline

Centre for Mathematics and Its Applications The Australian National University Canberra, ACT 0200 Australia. 1. Introduction

Transcription:

Local semiconvexity of Kantorovich potentials on non-compact manifolds Alessio Figalli, Nicola Gigli Abstract We prove that any Kantorovich potential for the cost function c = d / on a Riemannian manifold (M, g) is locally semiconvex in the region of interest, without any compactness assumption on M, nor any assumption on its curvature. Such a region of interest is of full µ-measure as soon as the starting measure µ does not charge n 1-dimensional rectifiable sets. 1 Introduction and main result Let (M, g) be an n-dimensional complete Riemannian manifold. We consider the Monge transport problem on M with cost function c(x, y) = d(x,y). This amounts to study the following problem: given two probability measures µ and ν on M, minimize c(x, S(x)) dµ(x) M among all maps S : M M such that T # µ = ν. The existence and uniqueness of an optimal transport map under the assumption that µ gives no mass to (n 1)-rectifiable sets and c(x, S(x)) dµ(x) < + inf S # µ=ν M has been proved in [4, 5]. The strategy was the following: first of all, one consider the Kantorovitch minimization problem inf c(x, y) dπ(x, y), γ Π(µ,ν) M M where Π(µ, ν) denotes the set of probability measures on M M which have µ and ν as first and second marginal, respectively. Then, it is a well-known fact in optimal transport theory that an optimal π exists, and it is contained in the c-subdifferential of a c-convex function ψ Centre de Mathématiques Laurent Schwartz, UMR 764, Ecole Polytechnique - 9118 Palaiseau, France. e-mail: figalli@math.polytechnique.fr University of Bordeaux. e-mail: nicolagigli@googlemail.com 1

(see [9, Theorems 4.1 and 5.1]): this means that there exist two measurable functions φ : M R {+ } and ψ : M R { } such that φ(x) = sup ψ(y) c(x, y), y M ψ(y) = inf φ(x) + c(x, y) (1.1) x M and π is concentrated on the set c φ := x M ( {x} c φ(x) ), where c φ(x) := { y M : φ(x) = ψ(y) c(x, y) }. Now, to complete the classical strategy in order to show existence and uniqueness of optimal maps, one needs to prove that φ is differentiable µ-a.e. Indeed, this allows to show that c φ(x) is a singleton for µ-a.e. x, which implies that the optimal plan π is concentrated on the graph of a function T. Moreover, such a function can be characterized µ-a.e. by the formula x c(x, y) y=t (x) = φ(x) T (x) = exp x ( φ(x)) (1.) (see for instance [4, Theorem 4.3] or [7, Proposition 1.15]). However, in [4, 5] the authors could not exactly complete the program described above, since they were not able to show that the function φ (which is also called Kantorovich potential) is differential µ-a.e. Indeed, since in the infimum appearing in (1.1) y varies in a non-compact set, φ does not a priori inherit the local semiconvexity property of the functions x d(x,y). Hence their strategy has been to prove a weaker statement, namely that φ is approximately differentiable µ-a.e., which turns out to be enough for obtaining existence and uniqueness of the optimal map. Moreover, at least when µ is absolutely continuous with respect to the volume measure, (1.) holds provided one replaces by the approximate gradient (see [4, Complement 3.4]). The aim of this paper is to fill the gap on the regularity of φ, showing that without any assumption on the manifolds nor on the measures µ and ν, the potential φ is always locally semiconvex on the region of interest. More precisely, we show the following: Theorem 1 Let φ be a Kantorovich potential as above. Then φ is locally semiconvex, and c φ(x) is non-empty and locally bounded for any x in the interior of {φ < + }. Moreover, the boundary of {φ < + } is (n 1)-rectifiable. Once the above theorem is proved, the existence and uniqueness of an optimal map T whenever µ does not charge (n 1)-rectifiable sets follows in a standard manner (see for instance [8, Lemma 7] or [4, Theorem 3.1]). Moreover (1.) holds µ-a.e., which allows to deduce (thanks to the fact that a semiconcave function is twice differentiable vol-a.e.) that T is differential µ-a.e. provided µ vol (see [3, Paragraph ] and [5, Proposition 3.4]). This last fact is important in order to compute the Jacobian of T and to obtain a change of variable formula (see [3, Corollary 4.7]). The next section is devoted to the proof of the above theorem. At the end of the section, we will also discuss some possible extensions of the above result to the case c = d p /p (p > 1), and to cost functions arising by minimizing a particular class of Tonelli Lagrangian. Moreover, we also make some general comments on how to deduce existence and uniqueness of optimal maps once the result is known in the compact case.

As a last comment, let us say that the approach we are going to describe is strongly inspired by a discussion made by Villani in [9, Chapter 1], where the author introduces the assumptions (H 1 ) and (H ), and he proves that if the cost function satisfies these assumptions, then a result closely related to ours holds (see [9, Theorem 1.4]). The point is that it is unclear - at least to us - whether the squared distance satisfies those assumptions or not (in particular, the problem is in checking (H )). Still, at least in the case c = d p /p or for costs coming from some special Tonelli Lagrangians, it is possible to conclude. We invite the reader to compare the proof of our result with the one of [9, Theorem 1.4]. Notice also that an approach similar to ours already appeared in [6, AppendixC]. Proof of the main theorem Set D := {φ < + }. We divide the proof in three steps. Step 1: φ is locally bounded in the interior of D. Since φ is defined by a supremum of continuous functions (see (1.1)), the fact that φ is locally bounded from below is immediate. Hence we only need to prove the bound from above. We argue by contradiction, and we assume the existence of a sequence x n x int(d) such that φ(x n ) +. For every n N, let us choose y n M a point such that φ(x n ) ψ(y n ) c(x n, y n ) + 1. (.1) In particular, as c = d, we have ψ(y n) + too. Hence, since R φ(x) ψ(y n ) c(x, y n ), we deduce that c(x, y n ) +, which further implies c(x n, y n ) = d(x n, y n ) +. Now, let γ n : [, d(x n, y n )] M be a minimizing geodesic parameterized by arc-length connecting x n to y n. Since d(x n, y n ) +, any geodesic γ n is defined at least on an interval [, l], for some l >. Let us define the following set: C n := { x M : there exists t [, l] s.t. d(x, γ n (t)) t/ }. (Observe that in an Euclidean space C n would just be a cone with height l and basis of radius l/.) We claim that inf φ + as n +. C n Indeed, if d(x, γ n (t)) t/ for some t [, l], thanks to the triangle inequality and (.1) we 3

have φ(x) ψ(y n ) d(x, y n) = ψ(y n ) [d(γ n(t), y n ) + t/] ψ(y n ) d(x n, y n ) φ(x n ) 1 + d(x n, y n ) t, ψ(y n ) [d(γ n(t), y n ) + d(x, γ n (t))] + d(x n, y n ) t = ψ(y n ) [d(x n, y n ) t/] (.) where at the second line we used the identity d(γ n (t), y n ) = d(x n, y n ) t. Thanks to the above inequality, we obtain that inf φ ψ(x n ) 1, (.3) C n which proves the claim. Now, letting n + and assuming l sufficiently small (say, 1/1 of the minimal injectivity radius in a neighborhood of x), it is easy to see by a simple compactness argument that the following holds: let v T x M be a limit point for γ n () T xn M (in some chart around x), and set C := { x M : there exists t [, l] s.t. d(x, exp x (tv)) t/ }. (.4) Then, up to subsequences, C n C in the Hausdorff distance, and φ in the interior of C. Since x C, this contradicts the fact that x int(d), and concludes the proof. Step : The boundary of D is (n 1)-rectifiable. This fact is a simple consequence of the proof of Step 1: assume that x D, and let {x n } D c be a sequence converging to x. Then, choosing a sequence {y n } M such that ψ(y n ) c(x n, y n ) n (this can always be done as φ(x n ) = + ), we deduce that inf φ n 1 C n (compare with (.3)), so that φ + in the interior of a cone C with vertex at x, height l and width l/ (see (.4)). Hence we have proved that at every boundary point x of D there exists an open cone of fixed height and width, with vertex at x, which is contained outside D. Then, it is a well-known result in geometric measure theory that the boundary of D is (n 1)- rectifiable, i.e. contained in a countable union of Lipschitz surfaces (see for instance the proof of [1, Theorem.61] or the one of [9, Theorem 1.48]). Step 3: c φ(x) is non-empty and bounded as x varies in a compact subset of int(d). In particular, φ is locally semiconcave inside D. 4

Let K int(d), take x K, and let y M be such that φ(x) ψ(y) c(x, y) + 1. We claim that d(x, y) is uniformly bounded, independently of y. Indeed, assuming without loss of generality d(x, y) 1, as in the proof of Step 1 we can consider the point γ k (l) on the (unit speed) geodesic from x to y k, where l dist(k, D)/. Then, by (.) we get ψ(γ(l)) ψ(x) 1 + d(x, y) l. Since by Step 1 ψ is uniformly bounded on the set K l := {x M : dist(x, K) l} D, the claim follows. The proof of Step 3 is now easy: thanks to (1.1) the function ψ(y) is upper semicontinuous. Hence, if x K and {y k } M is a maximizing sequence for φ, in the sense that φ(x) ψ(y k ) c(x, y k ) + 1 k, then by compactness (recall that that d(x, y k ) is uniformly bounded) there exists a point y such that φ(x) ψ(y) c(x, y). Since the opposite inequality is always true, we obtain that equality holds and y c φ(x). Moreover dist(k, y) C for some constant C depending only on K. Hence φ(x) = sup y M, dist(k,y) C ψ(y) c(x, y) x K, and the semiconvexity of φ in the interior of K follows. This concludes the proof of the main theorem. Remark (The case c = d p /p) The above result can be easily extended to the case c(x, y) = d(x,y) p p, p > 1. Indeed the only difference arises in the proof of (.), where by using the inequality (a b) p a p pa p 1 b for any b a we get φ(x) φ(x n ) 1 + d(x n, y n ) p 1 t x C n. Once one has the above inequality, the rest of the proof follows with no changes. Let us however point out that, since x d p (x, y)/p is not semiconcave in the classical sense (i.e. distributional locally second derivatives bounded from above) but only with a modulus of semiconcavity ω(t) = t p 1 (see [4, Appendices A and B]), the potential function will only be ω-semiconvex, which is however enough for the potential to be differentiable out of a countably (n 1)-rectifiable set. 5

Remark 3 (Costs induced by Lagrangian) A variant of the above argument still works when the cost function arises by minimizing a Lagrangian which behaves like θ( v x ), where θ : [, + ) [, + ) is convex, superlinear, and θ() = : c(x, y) = inf γ()=, γ(1)=y when L : T M R is a Lagrangian satisfying (a) L is C ; 1 L ( γ(t), γ(t) ) dt (b) for every (x, v) T M, L v (x, v) is positive definite on T x M; (c) there exist a convex superlinear function θ : R + R +, with θ() =, and A 1, A, A 3 positive constants, such that θ ( v x ) A1 L(x, v) A θ ( v x ) + A3, (x, v) T M. Indeed, the only main difference consists in the choice of the cone C n and the proof of (.): if x M, and (x n ), (y n ) are like in Step 1, we consider γ n : [, 1] M to be a minimizer for L going from x n to y n, that is and we set c(x n, y n ) = 1 L ( γ n (t), γ n (t) ) dt, γ n () = x n, γ n (1) = y n, C n := { x M : there exists t [, 1] s.t. d(x, γ n (t)) η d(x n, γ n (t)), with d(x n, γ n (t)) l }, where η, l > are small constants to be chosen. Let now x M be such that d(x, γ n (t)) η t for some t [, l], and take σ : [, t] M a constant-speed minimizing geodesic connecting γ n (t) to x. Then σ(τ) σ(τ) = d(x n, γ n (t))/t η d(x, γ n (t))/t, and thanks to the above assumptions on L we get c(x, y n ) c(x n, y n ) t t t L ( σ(τ), σ(τ) ) dτ t L ( γ n (τ), γ n (τ) ) dτ A θ ( ) t σ(τ) σ(τ) dτ θ ( γ n (τ) γn(τ)) dτ + (A1 + A 3 ) t ( η d(x, γ ) ( ) ] n(t)) d(x, γn (t)) θ + (A 1 + A 3 ), t t [ A θ where at the last step we used Jensen s inequality to estimate the second term. We now observe that, since θ() = and θ is convex, θ(η s) η θ(s) for all s. Hence, if we choose η = 1/(A ), arguing as in (.) we get φ(x) φ(x n ) 1 t θ ( d(x, γn (t)) t 6 ) + (A 1 + A 3 ) t.

This estimate allows to prove that if φ(x n ) +, then there exists a cone with (locally) fixed height and width, and with vertex at x, such that φ + inside this cone. This allows to prove Steps 1 and. To get Step 3, we have to show that, if φ(x) ψ(y) c(x, y) + 1, then d(x, y) is uniformly bounded. The proof of this fact is analogous of the one above, although a bit more involved: thanks to the above assumptions, the Hamiltonian H(x, p) behaves like θ ( p x ), where θ is the Legendre transform of θ. Hence, by exploiting the conservation of the Energy, as in proof of [4, Proposition B.17] one can show that γ(t ) at some t [, 1] controls max t [,1] γ(t) if γ is a minimizer. In this way, one can still prove d(x, y) has to be uniformly bounded, and the proof of Step 3 follows easily. (We leave the details to the interested reader.) Remark 4 (Compact vs. non-compact) Let us point out that, if one is simply interested in knowing that the optimal transport map exists and is unique, without having any information on its structure (namely, that is induced by a Kantorovich potential through the first formula in (1.)), then the argument for passing from the compact to the non-compact case is pretty simple (what we are going to say is not new - e.g. the argument was used already in [, Theorem 6..1] - still we believe it is worth to repeat it here). Assume indeed that existence and uniqueness of optimal maps is known whenever µ and ν are compactly supported (together with some suitable assumptions on µ). Then, in the general case, one simply takes an optimal plan π and consider its restriction to K n K n, where {K n } is an increasing sequence of compact sets such that n K n = M. If µ n and ν n denote the first and second marginal of π n := π (Kn Kn) π(k n K n) respectively, then it is well-known that π n is an optimal transport plan between them (see for instance [9, Theorem 4.6]). Hence any π n is concentrated on a graph, and letting n + also π has to be concentrated on a graph. This proves the existence. Moreover the fact that any optimal plan is concentrated on a graph immediately implies uniqueness, as a linear combination of optimal plans is still optimal. References [1] L.Ambrosio, N.Fusco & D.Pallara: Functions of bounded variation and free discontinuity problems. Oxford Mathematical Monographs. The Clarendon Press, Oxford University Press, New York,. [] L. Ambrosio, N. Gigli, and G. Savaré, Gradient flows in metric spaces and in spaces of probability measures, Lectures in Mathematics ETH Zürich. Birkhäuser Verlag, Basel, 5. [3] D. Cordero-Erasquin, R. J. McCann & M. Schmuckenschlager: A Riemannian interpolation inequality à la Borell, Brascamp and Lieb. Invent. Math., 146 (1), no., 19 57. [4] A. Fathi & A. Figalli: Optimal transportation on non-compact manifolds. Israel J. Math., to appear. 7

[5] A. Figalli: Existence, Uniqueness, and Regularity of Optimal Transport Maps. SIAM J. Math. Anal., 39 (7), no. 1, 16-137. [6] W. Gangbo & R. J. McCann: The geometry of optimal transportation. Acta Math., 177 (1996), 113-161. [7] N. Gigli Second order analysis on (P (M), W ). Submitted paper. Available at http://cvgmt.sns.it/cgi/get.cgi/papers/gig9/. [8] R. J. McCann: Polar factorization of maps on Riemannian manifolds. Geom. Funct. Anal., 11 (1), no. 3, 589-68. [9] C. Villani: Optimal transport, old and new. Grundlehren des mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences], Vol. 338, Springer-Verlag, Berlin-New York, 9. 8