Representation of the polar cone of convex functions and applications

Similar documents
On a Class of Multidimensional Optimal Transportation Problems

The optimal partial transport problem

On a weighted total variation minimization problem

Optimization and Optimal Control in Banach Spaces

PARTIAL REGULARITY OF BRENIER SOLUTIONS OF THE MONGE-AMPÈRE EQUATION

Integral Jensen inequality

L p Spaces and Convexity

MATH MEASURE THEORY AND FOURIER ANALYSIS. Contents

3. (a) What is a simple function? What is an integrable function? How is f dµ defined? Define it first

Existence of minimizers for the pure displacement problem in nonlinear elasticity

Division of the Humanities and Social Sciences. Supergradients. KC Border Fall 2001 v ::15.45

Subdifferential representation of convex functions: refinements and applications

Existence and monotonicity of solutions to moral hazard problems

Local strong convexity and local Lipschitz continuity of the gradient of convex functions

Functional Analysis I

Chapter 2 Convex Analysis

Bodies of constant width in arbitrary dimension

5 Measure theory II. (or. lim. Prove the proposition. 5. For fixed F A and φ M define the restriction of φ on F by writing.

If Y and Y 0 satisfy (1-2), then Y = Y 0 a.s.

BASICS OF CONVEX ANALYSIS

SYMMETRY RESULTS FOR PERTURBED PROBLEMS AND RELATED QUESTIONS. Massimo Grosi Filomena Pacella S. L. Yadava. 1. Introduction

Legendre-Fenchel transforms in a nutshell

Optimality Conditions for Nonsmooth Convex Optimization

2 Lebesgue integration

On a theorem of Alexandrov

Lecture 4 Lebesgue spaces and inequalities

Minimizing within convex bodies using a convex hull method

Convex Functions and Optimization

6. Duals of L p spaces

Convex Analysis and Economic Theory AY Elementary properties of convex functions

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

Probability and Measure

On John type ellipsoids

Convexity in R n. The following lemma will be needed in a while. Lemma 1 Let x E, u R n. If τ I(x, u), τ 0, define. f(x + τu) f(x). τ.

STOKES PROBLEM WITH SEVERAL TYPES OF BOUNDARY CONDITIONS IN AN EXTERIOR DOMAIN

APPROXIMATE ISOMETRIES ON FINITE-DIMENSIONAL NORMED SPACES

FULL CHARACTERIZATION OF OPTIMAL TRANSPORT PLANS FOR CONCAVE COSTS

+ 2x sin x. f(b i ) f(a i ) < ɛ. i=1. i=1

REGULAR LAGRANGE MULTIPLIERS FOR CONTROL PROBLEMS WITH MIXED POINTWISE CONTROL-STATE CONSTRAINTS

GEOMETRIC APPROACH TO CONVEX SUBDIFFERENTIAL CALCULUS October 10, Dedicated to Franco Giannessi and Diethard Pallaschke with great respect

Helly's Theorem and its Equivalences via Convex Analysis

Problem Set. Problem Set #1. Math 5322, Fall March 4, 2002 ANSWERS

Vector Integration and Disintegration of Measures. Michael Taylor

THEOREMS, ETC., FOR MATH 515

g 2 (x) (1/3)M 1 = (1/3)(2/3)M.

Continuous Sets and Non-Attaining Functionals in Reflexive Banach Spaces

Functional Analysis. Franck Sueur Metric spaces Definitions Completeness Compactness Separability...

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

Signed Measures. Chapter Basic Properties of Signed Measures. 4.2 Jordan and Hahn Decompositions

Chapter 6. Integration. 1. Integrals of Nonnegative Functions. a j µ(e j ) (ca j )µ(e j ) = c X. and ψ =

CHAPTER 2: CONVEX SETS AND CONCAVE FUNCTIONS. W. Erwin Diewert January 31, 2008.

Chapter 8. General Countably Additive Set Functions. 8.1 Hahn Decomposition Theorem

Convex Optimization Notes

Calculus of Variations. Final Examination

(convex combination!). Use convexity of f and multiply by the common denominator to get. Interchanging the role of x and y, we obtain that f is ( 2M ε

PCA sets and convexity

Optimality Conditions for Constrained Optimization

1/12/05: sec 3.1 and my article: How good is the Lebesgue measure?, Math. Intelligencer 11(2) (1989),

SOME PROPERTIES ON THE CLOSED SUBSETS IN BANACH SPACES

On the distributional divergence of vector fields vanishing at infinity

CONTENTS. 4 Hausdorff Measure Introduction The Cantor Set Rectifiable Curves Cantor Set-Like Objects...

Solutions to Tutorial 11 (Week 12)

Recall that if X is a compact metric space, C(X), the space of continuous (real-valued) functions on X, is a Banach space with the norm

CHAPTER 6. Differentiation

ZERO DUALITY GAP FOR CONVEX PROGRAMS: A GENERAL RESULT

Inequalities of Babuška-Aziz and Friedrichs-Velte for differential forms

Legendre-Fenchel transforms in a nutshell

MA651 Topology. Lecture 9. Compactness 2.

Lecture 1: Entropy, convexity, and matrix scaling CSE 599S: Entropy optimality, Winter 2016 Instructor: James R. Lee Last updated: January 24, 2016

On Semicontinuity of Convex-valued Multifunctions and Cesari s Property (Q)

Real Analysis, 2nd Edition, G.B.Folland Signed Measures and Differentiation

Convex Optimization Theory. Chapter 5 Exercises and Solutions: Extended Version

Brøndsted-Rockafellar property of subdifferentials of prox-bounded functions. Marc Lassonde Université des Antilles et de la Guyane

On the uniform Poincaré inequality

Extreme Abridgment of Boyd and Vandenberghe s Convex Optimization

Dual methods for the minimization of the total variation

Convex analysis and profit/cost/support functions

b i (µ, x, s) ei ϕ(x) µ s (dx) ds (2) i=1

Lebesgue-Radon-Nikodym Theorem

SOME REMARKS ON THE SPACE OF DIFFERENCES OF SUBLINEAR FUNCTIONS

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

REGULARITY OF MONOTONE TRANSPORT MAPS BETWEEN UNBOUNDED DOMAINS

SEPARABILITY AND COMPLETENESS FOR THE WASSERSTEIN DISTANCE

A Dual Condition for the Convex Subdifferential Sum Formula with Applications

Overview of normed linear spaces

for all subintervals I J. If the same is true for the dyadic subintervals I D J only, we will write ϕ BMO d (J). In fact, the following is true

Existence of minimizers for the Newton s problem of the body of minimal resistance under a single impact assumption

Math The Laplacian. 1 Green s Identities, Fundamental Solution

HILBERT SPACES AND THE RADON-NIKODYM THEOREM. where the bar in the first equation denotes complex conjugation. In either case, for any x V define

Notions such as convergent sequence and Cauchy sequence make sense for any metric space. Convergent Sequences are Cauchy

GENERALIZED CONVEXITY AND OPTIMALITY CONDITIONS IN SCALAR AND VECTOR OPTIMIZATION

6 Classical dualities and reflexivity

02. Measure and integral. 1. Borel-measurable functions and pointwise limits

CHAPTER I THE RIESZ REPRESENTATION THEOREM

l(y j ) = 0 for all y j (1)

OPTIMAL POTENTIALS FOR SCHRÖDINGER OPERATORS. 1. Introduction In this paper we consider optimization problems of the form. min F (V ) : V V, (1.

Free energy estimates for the two-dimensional Keller-Segel model

2 Measure Theory. 2.1 Measures

Lecture Notes in Advanced Calculus 1 (80315) Raz Kupferman Institute of Mathematics The Hebrew University

ASYMPTOTIC BEHAVIOUR OF NONLINEAR ELLIPTIC SYSTEMS ON VARYING DOMAINS

Transcription:

Representation of the polar cone of convex functions and applications G. Carlier, T. Lachand-Robert October 23, 2006 version 2.1 Abstract Using a result of Y. Brenier [1], we give a representation of the polar cone of monotone gradient fields in terms of measure-preserving maps, or bistochastic measures. Some applications to variational problems subject to a convexity constraint are given. Keywords: Convexity constraint, Euler-Lagrange equation, measure-preserving maps, bistochastic measures. This article was completed after Thomas sudden death. It is dedicated to his memory. CEREMADE, UMR CNRS 7534, Université Paris IX Dauphine, Pl. de Lattre de Tassigny, 75775 Paris Cedex 16, carlier@ceremade.dauphine.fr. Laboratoire de Mathématiques, UMR CNRS 5127, Université de Savoie, 73376 Le Bourget-du-Lac Cedex, France. 1

1 Introduction Minimization problems in the set of convex functions, of the following form: g(x, u(x), u(x)) dx (1) inf u H 1 () u convex arise in several fields of applied mathematics (see [2], [5]). Even for a very simple and well-behaved g, the convexity constraint raises a number of mathematical difficulties. Even the simplest case, that is the projection problem: inf u H 1 () u convex u q 0 2 (2) (where q 0 L 2 () N is given) has yet to be investigated in detail, in particular regarding regularity issues. When dealing with this type of minimization problems on a closed cone (here the cone of convex functions), it is natural to try to identify the polar cone in order to write an Euler-Lagrange equation. A first proposal in this direction has been given in [7]; unfortunately the corresponding equation is difficult to handle in general. In the present paper we propose a different representation of the polar cone of the set of convex functions (more precisely: of the polar cone of the set of gradients of convex functions). It makes use of the set S of measurepreserving maps (see the precise definition and statements in the next section). We prove that every vector field in L 2 () N can be decomposed as the sum of the gradient of a convex function and an element of the closed cone generated by (S id) (Section 3). This is to be compared to the well-kown polar decomposition of Y. Brenier [1], which is a composition instead of a sum. This result is actually a direct consequence of a result of Y. Brenier (see [1] and Proposition 2.1). Another way to express the same decomposition is to make use of the so-called bistochastic measures, a special case of Young measures (Section 4). Some applications to variational problems of the form (1) are given in sections 5 (Euler-Lagrange equations) and 6 (regularity). 2 Representation of the polar cone of gradients of convex functions In the whole paper, is some bounded open convex subset of R N, N 1. We denote by, the usual scalar product in the Hilbert space L 2 () N. We consider K := {q L 2 () N : q = u, for some convex function u}. 2

It is a closed convex cone in L 2 () N. Its polar cone is defined as usual: K := {p L 2 () N : p, q 0, q K}. We make use of the set of measure-preserving maps: S := {s measurable : : s dx = dx}. Here dx denotes Lebesgue s measure and s dx is the measure defined by ϕ(x)s dx = ϕ(s(x))dx for every bounded continuous function ϕ. Let us start with a characterization of K: Proposition 2.1 (Brenier) Let q L 2 () N ; then q K if and only if q, s id 0, for all s S. The proof is already given in [1], we recall it here briefly for the sake of completeness. Proof. Assume first that q K, q = u with u convex. Let s S; by convexity, one has u(x) (s(x) x) u(s(x)) u(x), hence: u, s id u(s(x))dx u(x)dx = 0. Conversely, assume that q satisfies the inequalities of the Proposition and let us prove that q K. Let (x 0, x 1,..., x n = x 0 ) n+1 be a cycle of distinct Lebesgue points of q, let ε > 0 be such that the balls B(x i, ε) are disjoint and included in. Define then the measure preserving map s ε by x if x \ n 1 B(x i, ε) s ε (x) = i=0 x + (x i+1 x i ) if x B(x i, ε), i = 0,..., n 1. We have by assumption. This yields n 1 i=0 q, s ε id 0 1 q(x) (x i+1 x i ) dx 0. (3) B(x i, ε) B(x i,ε) 3

Letting ε go to zero, we get then n 1 q(x i ) (x i+1 x i ) 0. i=0 In other words the set A := {(x, q(x)), x Lebesgue point of q} is cyclically monotone, so that from Rockafellar [9] there exists a convex function u such that A u; in particular q = u almost everywhere. As a consequence of Proposition 2.1 one may characterize the polar cone of K: Corollary 2.1 The polar cone K is the closure (in L 2 ) of the convex cone generated by (S id) K = cone(s id). Proof. Proposition 2.1 can be stated as: K = (S id) = (cone(s id)). Taking the polar cones of those two closed convex cones gives the desired result. One may ask what is the difference between the two cones cone(s id) and R + co(s id). Indeed, the latter is much easier to charaterize, as we shall see in the next sections. It has also some interesting properties: Proposition 2.2 Let p R + co(s id), p 0, so that p = λ(τ id) with λ > 0 and τ co S \ {id}. Then, for any convex function u H 1 (), one has u(τ(x)) dx u(x) dx. (4) Additionnally, for any convex function u H 1 () such that p, u = 0, one has: u(τ(x)) dx = u(x) dx (5) u(τ(x)) u(x) = u(x) (τ(x) x) a.e. x. (6) Notice that (6) means that the convex function u is actually affine on the segment [x, τ(x)]. In particular, p R + co(s id) and p, u = 0 for some strictly convex u implies p = 0. 4

Let us postpone the proof for a while and continue the discussion on the links between cone(s id) and R + co(s id). One has obviously R + co(s id) cone(s id) = R + co(s id), (7) but it turns out that R + co(s id) is not closed and therefore there is no equality in the inclusion above. Indeed, consider the set ker div := {p (L 2 ()) N : p, v = 0, v H 1 ()}. (That is, divergence-free vector fields and whose normal component at the boundary vanishes, whenever regular.) Corollary 2.2 One has ker div cone(s id) ker div R + co(s id) = {0}. In particular, R + co(s id) is not closed. Proof of Proposition 2.2 and Corollary 2.2. Since τ co S, there exists some sequence of the form n i=1 αn i s n i with s n i S, αi n 0, i αn i = 1, converging in L 2 and almost everywhere to τ. For any u H 1 () convex, we have u ( i αn i sn i (x)) i αn i u(sn i (x)), for almost every x. Hence [ u ( n αi n s n i (x) ) ] n u(x) dx (u(s n i (x)) u(x)) dx = 0, i=1 taking into account that each s n i is measure preserving. In the limit we get (4) by Fatou s Lemma. If we assume p, u = 0, then we get: [ ] u(τ(x)) u(x) u(x) (τ(x) x) dx 0. Since u is convex, the integrand is nonnegative. Hence we obtain (5 6). It is clear that ker div cone(s id), since from the very definitions p ker div and q K implies p, q = 0. Now assume p ker div R + co(s id), p 0. Then p, u = 0 for all u, so (6) holds also for any u convex. Since this equation means that u is affine on the segment [x, τ(x)], there is clearly a contradiction whenever u is strictly convex. 5 i=1 α n i

The non-closedeness of R + co(s id) makes difficult the manipulation of cone(s id) in general. One has to notice the following, however: Proposition 2.3 Let p cone(s id). Then either p R + co(s id), or, for any sequences (λ n ) R +, (τ n ) co S such that λ n (τ n id) p strongly in L 2 () N, one has λ n + and τ n id. Proof. Since p cone(s id), the closure of R + co(s id), it can be written as the limit of a sequence (p n ) such that p n = λ n (τ n id), with λ n R +, τ n co S. If λ n does not converge to +, then we may extract a subsequence and assume that it converges to some limit λ R +. If λ = 0, then p n 0 since co S is bounded: so p = 0 R + co(s id). If λ > 0, then τ n = id + λ 1 n p n converges to τ = id + λ 1 p, and τ co(s) since this is a closed set. Now p = λ(τ id) R + co(s id). If λ n + on the other hand, one get τ n id = λ 1 n p + o(1) 0 as n. Remark 1. We have seen that R + co(s id) is dense in K and that R + co(s id) is not closed in L 2, a natural question at this point is then : how big is K \R + co(s id)? To partially answer this question, let us remark that R + co(s id) L and that K contains many non bounded elements: for instance all the (not necessarily bounded) vector fields in ker div, but also v for every v H 1 0() with v 0, since v u = u, v. One can also construct non bounded elements of K as follows: let f 0 be a scalar function in L 2 \ L. Denoting by S f the set of maps σ : preserving the measure f(x) dx, one actually has (σ id)f L 2 and K for every σ S f. For instance if is the ball and f is further assumed to be radial then (R id)f K for every rotation R. Remark 2. In view of corollary 2.2, it is also natural to wonder whether one has K = ker div + R + co(s id)? The previous identity is equivalent to ker div +R + co(s id) being closed in L 2. Again, the answer is unfortunately negative as the following counter-example shows. Assume that is the unit ball of R 2 and define v(x) := (1 x ) 2/3, we already know that v K. Now, if v admitted a decomposition of the form v = p + λ(τ id) with (p, λ, τ) ker div R + co(s) then for every φ W 1,1 () the quantity v. φ would be finite. Taking for instance φ(x) := (1 x )1/4 yields the desired contradiction. Remark 3. One should finally notice that, even for a p R + co(s id), the writing p = λ(τ id) is not unique. This is obvious if p = 0, where one can 6

choose λ = 0 and any τ, or τ = id and any λ. But it is also true if p 0. Indeed, for any given τ co S \ {id}, let us consider I τ := {r 0; τ r := rτ + (1 r)id co S}. Since co S is closed, convex, and bounded, I τ is an interval of the form [0, R], with R 1. Now we have p = λ(τ id) = λ r (τ r id) for any r in this interval. 3 Decomposition of vector fields Given q 0 L 2 () N we consider the problem (2), which may also be stated as follows: given q 0 L 2 () N, solve inf q K q q 0 2 L 2. (8) There exists a unique solution q = u of this projection problem. Moreover the characterization of this solution is given by: Theorem 3.1 Let q 0 L 2 () N. Then there exist a unique (up to a constant) u H 1, convex, a unique p K = cone(s id), such that: q 0 = u + p (9) u, p = 0. (10) Here q = u solves (8), while p solves the dual problem (projection on K ). Moreover, if p 0 ( i.e. q 0 / K), then either p = lim n λ n (τ n id) where λ n R + +, τ n co(s) id, (11) or p = λ(τ id) with λ > 0, τ co S and u(τ(x)) dx = u(x) dx (12) u(τ(x)) u(x) = u(x) (τ(x) x) a.e. x. (13) Proof. The minimization problem (8) can also be written as: inf q L 2 () N where L(p, q) := 1 2 q q 0 2 + p, q. sup L(p, q) (14) p K 7

It is well-known that L admits a unique saddle point (q, p) = ( u, p) in K K (see for instance [6]) where q = u is the solution of (14) and p is the solution of the dual problem: Since sup p K inf L(p, q). (15) q L 2 () N inf L(p, q) = 1 q L 2 () N 2 p q 0 2 + 1 2 q 0 2, the solution of (15), that is p, is the projection of q 0 on K = cone(s id), i.e. the solution to inf p q 0 2. (16) p cone(s id) The saddle-point (q, p) is characterized by (9 10) since K, K are polar cones. Conversely if (9-10) hold with u convex and p K then u (respectively p) is the projection of q 0 onto K (respectively onto K ). If p 0, we deduce the remaining properties from Proposition 2.2 and Proposition 2.3. If, in the previous satetement, p = λ(τ id) with λ > 0 and τ co S\{id}, then u is affine between x and τ(x) for almost every x; in particular if x is such that (x, u(x)) does not belong to a face of dimension 1 of the epigraph of u, then p(x) = 0. When p cone(s id) \ R + co(s id) (a situation that, as already noted, may unfortunately occur), we don t have a tractable representation of p so that we don t have such a nice property as (13). 4 Bistochastic measures In this section, we investigate a little bit further what elements of co S look like (the closure is still taken in L 2 ). This can conveniently be done in terms of bistochastic measures. 4.1 Bistochastic measures and co S Let a sequence of elements of co S of the form τ n := k n i=1 αn i sn i (αi n 0, i αn i = 1) converge in L 2 to some limit τ. For each n and each x one may define the probability measure µ n x := k n i=1 α n i δ s n i (x) 8

so that for every x one has τ n (x) = y dµ n x(y). (17) One may also define the nonnegative Radon measure γ n on by ( ) f(x, y) dγ n (x, y) := f(x, y) dµ n x(y) dx, f C0( 0 ). Since each s n i is volume-preserving, one has for every continuous bounded function f on : f(x) dγ n (x, y) = f(y) dγ n (x, y) = f. This expresses that the two projections (marginals) of γ n equal Lebesgue measure on ; hence γ n is a so-called bistochastic measure. Note that bistochastic measures are special cases of Young measures [11]. It is standard then to prove that the limit τ can also be represented by the measure given by the weak limit γ of (γ n ) (or some subsequence if necessary). More precisely, γ is also a bistochastic measure which therefore admits a disintegration given by a measurable family of probability measures {µ x } x : ( ) f(x, y)dγ(x, y) := f(x, y) dµ x (y) dx, f C0( 0 ), and τ(x) = y dµ x (y), a.e. x. (18) Note finally for further use that since γ is bistochastic, one has for every continuous bounded function f on : ( ) f(y) dµ x (y) dx = f(x)dx. (19) It is also interesting to notice that v convex, v(τ(x)) dx v(x) dx. (20) This comes from (18) and Jensen s inequality. We see in (5) that u saturates the inequality (20) with respect to τ. 9

4.2 The decomposition with bistochastic measures Let us see now how to use bistochastic measures for the projection problem (8). As already noticed u is the projection of q 0 onto K if and only if q 0 can be decomposed as in (9 10). We are interested in the nontrivial case q 0 / K, denoting by p the projection of q 0 on K, we then have p 0. Let us assume that p R + co(s id) (which, unfortunately, is not the case for any q 0 L 2 () N ) then p = λ(τ id) with λ > 0 and τ co S. In this case, (12 13) hold. Using bistochastic measures we see that there exists a measurable family {µ x } x of probability measures on, satisfying (19) and such that (18) hold (where τ now is given by Theorem 3.1). Using (19), (5) and (18), we get ( ) ( ) u(y) dµ x (y) dx = u = u(τ(x))dx = u y dµ x (y) dx. Since u is convex this implies that for a.e. x : ( ) u y dµ x (y) = u(τ(x)) = u(y) dµ x (y). This means that u is essentially affine on Supp(µ x ). In particular if u is differentiable at x then Supp(µ x ) u ( u(x)), or equivalently: u(y) = u(x) + u(x) (y x), µ x -a.e. y. (21) Remark 4. At least formally, one may characterize u, the solution of (2), in terms of balayages, a well-known notion in potential theory and probabilities (see Meyer [8]). Indeed if K is a convex compact subset of R n, the polar cone of convex functions for the duality between C 0 (K, R) functions and Radon measures in K (a duality that is not adapted to our variational problem) can be represented using Cartier-Fell-Meyer Theorem [4]. This theorem states that if a signed measure with Jordan decomposition λ µ is in the polar cone of C 0 (K, R) convex functions (so that λ and µ have the same mass and the same moment), then there exists a measurable family {ν x } x K of probability measures on K such that the barycenter of ν x is x for all x and for every f C 0 (K, R): ( ) f(x) dµ(x) = f(y) dν x (y) dλ(x). (22) K K If u C 0 (K, R) is convex and u dµ = u dλ then it is easy to see that u K K is affine on Supp(ν x ). Let us also note that if we specify K :=, then there K 10

is no reason for the measures λ and µ above to vanish on. Formally one would like to use this representation to characterize u, solution of (2), by: u q 0, v = v d(µ λ), v H 1, u d(µ λ) = 0 (23) where λ and µ satisfy (22). We refer to Rochet and Choné [5] for the use of balayages in variational problems subject to a convexity constraint. Even though what one formally obtains with balayages and our representation using bistochastic measures look similar, one has to take the divergence of our representation to obtain Rochet and Choné s one. In other words, there is no obvious link with the kernel {µ x } we obtain and the kernel {ν x } formally obtained from Cartier-Fell-Meyer Theorem. 5 Euler-Lagrange Equations We may use now Proposition 2.1 to derive the Euler-Lagrange equation of problems of the form (1). In the following, we assume that g is differentiable with respect to u and its third argument q and satisfies some suitable growth conditions. Since we consider a minimization problem in H 1, we can add constant functions to the solutions; we deduce that any solution u of (1) satisfies: g (x, u, u) dx = 0. (24) u This implies in particular that there exists a solution ψ to the Laplace equation with homogeneous Neumann boundary condition: ψ = g (x, u, u) in, u (25) ψ n = 0 on. The variational inequalities associated to (1) can be written as: ( ) g (x, u, u) ψ (v u) dx 0, for all convex v H 1 (), q or equivalently ψ g (x, u, u) cone(s id), q ( ψ g ) (x, u, u) u dx = 0. q 11

This yields: g (x, u, u) = ψ p (26) q for some p cone(s id) with p, u = 0. This implies in particular: g ( g ) u (x, u, u) div q (x, u, u) + p = 0. (27) In the special case where p is of the form p = λ(τ id) with λ > 0 and τ co S \ {id} (again, this is not always true and almost impossible to check a priori), then for almost every x such that (x, u(x)) does not belong to a face of dimension 1 of the epigraph of u, then p(x) = 0. In this case, we get the following alternative (or complementary slackness condition) for the solutions of (1): for a.e. x, either there is a line segment in the graph of u containing (x, u(x)), or g (x, u, u) = ψ. q 6 Regularity In this final section, given q 0 L 2 () N, we consider the projection problem (2). We recall that its solution u is characterized by the conditions q 0 = u + p with p cone(s id), u, p = 0. Proposition 6.1 Let us assume that p R + co(s id). 1. if q 0 is continuous at x 0 then u is differentiable at x 0, 2. if q 0 C 0 (, R N ) then u C 1 (, R) and if q 0 C 0,α (, R N ) for some α (0, 1) then u C 1,α (, R). Proof. If p = 0, there is nothing to prove, by assumption we therefore have p = λ(τ id) with λ > 0 and τ co S \ {id}. Using the representation (18) of elements of co S as moments of bistochastic measures, we deduce that there exists a measurable family of probability measures {µ x } x such that: τ(x) = y dµ x (y), a.e. x. (28) 12

1. Assume by contradiction that q 0 is continuous at x 0 and u(x 0 ) is not reduced to a point. With no loss of generality we may assume that u(x 0 ) = 0 and that 0 belongs to the relative interior of u(x 0 ). By a standard convex analysis argument, there exist k N + 1, (α 1,, α k ) (0, 1) k such that k ı=1 α i = 1, sequences (x i n ) n N, for i = 1,, k such that u is differentiable at every x i n, q 0 (x i n) = u(x i n) + λ(τ(x i n) x i n), the measure µ x i n is well-defined at x i n and: lim n x i n = x 0, lim n u(x i n ) = γ i 0, i = 1,, k, and We then have: hence: k α i γ i q 0 (x i n) = i=1 q 0 (x i n ) = u(xi n ) + λ ( k α i γ i u(x i n) + λ i=1 ydµ x i n (y) x i n k ( α i i=1 ) k α i γ i = 0 ı=1 y γ i dµ x i n (y) γ i x i n (29) (30) Using the fact that q 0 is continuous at x 0 and passing to the limit yields: 0 > k α i γ i 2 = λ lim n i=1 k ( α i i=1 ) y γ i dµ x i n (y) Moreover, Supp(µ x i n ) u ( u(x i n )) so that for µ x i -a.e. y one has: n which can also be written as ) (31) u(y) = u(x i n ) + u(xi n ) (y xi n ) (32) γ i y = u(y) + ε i n (y) (33) with ε i n (y) = u(xi n ) xi n u(xi n ) + (γ i u(x i n )) y which tends to 0 as n tends to. Using (33), and the fact that u u(x 0 ) = 0 we get: lim y γ i dµ x i n n (y) 0 for all i = 1,, k and the latter contradicts (31). 2. The previous argument actually proves that u is continuous on provided q 0 is. This implies that τ is continuous on too and q 0 (x) = u(x) + λ(τ(x) x), x. (34) 13

In addition, since u is affine between x and τ(x), we also have: Let (x, y), using (34), one gets Using (35) yields u(x) = u(τ(x)), x. (35) u(x) u(y) 2 + λ(τ(x) τ(y)) ( u(x) u(y)) = (q 0 (x) q 0 (y) + λ(x y)) ( u(x) u(y)). (36) (τ(x) τ(y)) ( u(x) u(y)) = (τ(x) τ(y)) ( u(τ(x)) u(τ(y)) 0. With (36), we thus obtain: u(x) u(y) q 0 (x) q 0 (y) + λ x y If q 0 C 0,α (, R N ) for some α (0, 1), the previous inequality implies that u C 1,α (, R) and the proof is complete. Theorem 6.1 states two kinds of regularity results that are of quite different nature: the first assertion states a pointwise regularity result while the second states a global regularity result. Both statements are obtained under the assumption that p R + co(s id). This assumption is not satisfactory, since it is not always satisfied and at least very difficult to check. Concerning the global regularity result, we have not been able to remove it although we believe that the result remains true for a general p cone(s id). Let us mention however that the proof above carries over when p is of the form (σ id)f with f C 0 (, R + ) and σ in the closed convex hull of S f defined in Remark 1 (but again this is a special case). We refer the reader to [3] for other C 1 regularity results obtained with totally different arguments. The pointwise regularity part is more surprising (from the point of elliptic regularity theory for instance), and it does not hold in general (i.e. when p / R + co(s id)) as the next counter-example shows. Let := ( 3, 3) ( 1, 1) and define for all x = (x 1, x 2 ) : max( x 1 + 2, x 2 ) 1 if max( x 1 + 2, x 2 ) 1 u 0 (x) := max( x 1 2, x 2 ) 1 if max( x 1 2, x 2 ) 1 0 otherwise. 14

Define q 0 := u 0 and u as the largest convex minorant of u 0. We then have q 0 = u+ (u 0 u). Since u 0 u 0 and u 0 = u on, (u 0 u) K (see Remark 1 ). On the other hand, it is easy to check that u (u 0 u) = 0. This proves that u is the projection of q 0 := u 0 on K. Let us remark now that u is singular on the segment [( 2, 0), (2, 0)] whereas q 0 is continuous on this segment except at the points ( 2, 0), ( 1, 0), (1, 0) and (2, 0). This proves that the pointwise regularity statement is not true without additional assumption on the multiplier p. Acknowledgements: The authors thank Yann Brenier for bringing to their attention Proposition 2.1 and for stimulating discussions. References [1] Y. Brenier, Polar factorization and monotone rearrangement of vectorvalued functions, Comm. Pure Appl. Math. 44 4, pp. 375 417 (1991). [2] G. Buttazzo, V. Ferone and B. Kawohl, Minimum problems over sets of concave functions and related questions. Math. Nachr. 173, pp. 71 89 (1995). [3] G. Carlier, T. Lachand-Robert Regularity of solutions for some variational problems subject to a convexity constraint, Comm. Pure Appl. Math., 54, pp. 583 594 (2001). [4] P. Cartier, J.M.G. Fell, P.-A. Meyer Comparaison des mesures portées par un ensemble convexe compact, Bull. Soc. Math. France, 92, pp. 435 445 (1964). [5] P. Choné, J.-C. Rochet, Ironing, Sweeping and Multidimensional screening, Econometrica, vol. 66 (1998), pp. 783 826. [6] I. Ekeland, R. Temam, Convex Analysis and Variational Problems, North-Holland (1972). [7] P.-L. Lions, Identification du cône dual des fonctions convexes et applications, C. R. Acad. Sci. Paris (1998). [8] P.-A. Meyer, Probabilités et Potentiel, Hermann, Paris (1966). [9] R. T. Rockafellar, Characterization of the subdifferentials of convex functions, Pacific J. Math. 17, pp. 497 510 (1966). [10] R. T. Rockafellar, Convex Analysis, Princeton University Press (1970). 15

[11] L. C. Young, Lectures on the calculus of variations, Chelsea, New York (1980). 16