• Aucun résultat trouvé

How to show that some rays are maximal transport rays in Monge Problem

N/A
N/A
Protected

Academic year: 2022

Partager "How to show that some rays are maximal transport rays in Monge Problem"

Copied!
23
0
0

Texte intégral

(1)

How to show that some rays are maximal transport rays in Monge problem.

ALDOPRATELLI(*)

ABSTRACT- In this paper we prove a result, extending Lemma 8.1 in [2], which al- lows to proof that a set of segments is the set of the maximal transport rays for a transport problem. This is particularly useful to build non-trivial examples of transport maps, then in particular to provide specific examples (or counter- examples) in mass transportation. We also give some of these examples.

1. Introduction

1.1.The mass transportation problem.

We are interested in the mass transportation problem, first pro- posed by Monge [11] in 1781; in today's language, we can state it as follows. Th e ambient space is th e closure V of an open, bounded and convex subset VR2, and we are given two probability measuresf‡ and f on V. Atransport mapfromf‡ tof is a Borel mapt:V!V suchthat t#f‡ˆf , wh ere th e push -forward t#:M‡(V)! M‡(V) is defined as t#f‡(B):ˆf‡ t 1(B)

for any Borel set BV. If we th ink thatf‡ andf measure respectively the distribution of some mass and the depth of a hole, then any transport map is a «strategy» to move all the mass inside the hole, covering the latter. To any transport map we associate a cost given by

C(t):ˆ

…

V

jt(x) xjdf‡(x);

…1:1†

and the mass transportation problem consists in finding an optimal

(*) Indirizzo dell'A.: Dipartimento di Matematica «F. Casorati», via Ferrata 1, 27100 Pavia, E-mail address: a.pratelli@sns.it

(2)

transport map, that is, a transport map minimizing the cost C. In general, there may be no optimal transport maps, or even no transport maps at all; however, one can consider a sort of relaxation of this problem, as first proposed by Kantorovich([9, 10]) in the '40s: we call transport plan between f‡ and f any probability measure g2 P(VV) such that the two projectionsp1gandp2gofgonVcoincide with f‡ and f respectively. Moreover, to any transport plan g one associates the cost

C(g):ˆ

……

VV

jy xjdg(x;y):

…1:2†

Any transport map t is «naturally» associated to the transport plan gt ˆ(1;t)#f‡ (in fact, it is trivially checked that gt is a transport plan whenevert is a transport map, as well asC(t)ˆ C(gt)). However, there are muchmore transport plans than transport maps, and it can be sh own th at, unlike th e transport maps, th ere always exist transport plans (for instance, f‡f is such a plan), and there always exist also optimal transport plans, that is, transport plans minimizingC(sinceC is easily a weakly* l.s.c. functional and the set of the transport plans is weakly* compact in M‡(VV)). The mass transportation problem is widely studied, and in its more general statements the ambient spaceV is replaced by a subset ofRN, or a manifold, or even a general metric space, and the Euclidean distance j j appearing in (1.1) and (1.2) is replaced by a more general norm, or by a distance, or even by a generic l.s.c. functions: there is a huge literature about this field, for a general reference we only quote [13, 1, 12].

We consider now the maximization problem, sometimes referred to as shape optimization problem, which consists in maximizing the functional

I(u):ˆ

…

V

u(x)d(f‡ f )(x) …1:3†

among the 1-Lipschitz functions u:V!R; any maximizer is called Kantorovich potential (or also optimal shape, depending on the point of view). The following result, stating a deep connection between the two problems, is well known since the work of Kantorovich (see for instance [4, 5, 1]).

THEOREM1.1. inf (1:2)ˆsup (1:3)and both the extremals are reached.

(3)

Moreover, given any Kantorovich potential u and any optimal plang, one has u(x) u(y)ˆ jy xjforg a.e.(x;y)2VV.

As the Theorem above enlightens, the segments on which the mass is moved,i.e.the segmentsxywith(x;y)2sptg, have particular properties.

To understand them, we need some standard notation (see for in- stance [7, 2, 12]): given an optimal transport plang, we say th at

an open oriented segmentxyis atransport rayif (x;y)2sptg;

an open oriented segmentxyis amaximal transport rayif any z2xybelongs to the closure of some transport ray contained intoxy, and xyis not strictly contained into any segment withthe same property;

a pointzis adoubling pointif it belongs to the closure of at least two different maximal transport rays;

T:ˆ

z2R2: zis contained into some maximal transport ray is thetransport set.

It is to be remarked that a maximal transport ray need not to be a transport ray; we also notice the following facts.

REMARK1.2. Thanks to Theorem 1.1, for any Kantorovich potential u and any optimal transport plan g, u is decreasing at maximal slope on any maximal transport ray; this implies, in particular, that different maximal transport rays can not cross (recall that the maximal transport rays are open segments).

By the linearity of the functional (1.2) as well as of the constraints p1gˆf‡;p2gˆf , it is clear that any convex combination of optimal transport plans is still an optimal transport plan. Therefore, the above Remark says, more generally, that two different maximal transport rays can not cross, even if they are obtained from two different optimal trans- port plans. In fact, usually the maximal transport rays do not depend on the choice of an optimal transport plan, see Remark 2.12.

From now on, we will consider the maximal transport rays as oriented following the slope ofu(by Theorem 1.1, this orientation does not depend on the choice of the Kantorovich potential u): therefore, whenever (x;y)2sptgwe have thatxybelongs to some maximal transport ray and x>y. Finally, from the definition it follows also that any doubling point does not belong toT, since it is an endpoint of all the maximal transport rays to the closure of which it belongs; more precisely, it is either the upper

(4)

endpoint of all the maximal transport rays to the closure of which it be- longs, or the lower endpoint of all them.

We conclude this section with an easy but important consequence of Theorem 1.1; first we give a straightforward

COROLLARY1.3. Given a transport plangand a1 Lipschitz function u, one has that bothganduare optimal (for…1:2†and…1:3†respectively) if and only ifC(g)ˆI(u).

Then we can sharpen the above claim as follows:

LEMMA1.4. Given a transport plangand a1 Lipschitz function u, one has that bothgand u are optimal (for …1:2†and…1:3†respectively) if and only ifgis concentrated on the set

(x;y)2VV: u(x) u(y)ˆ jy xj . PROOF. An implication is already given by Theorem 1.1; concerning the other one, takenuandgas in the statement one can evaluate

C(g)ˆ ……

VV

jy xjdg(x;y)ˆ ……

VV

u(x) u(y)dg(x;y)

ˆ ……

VV

u(x)dg(x;y) ……

VV

u(y)dg(x;y)ˆ …

V

u(x)df‡(x) …

V

u(y)df (y)

ˆ …

V

u(x)d(f‡ f )(x)ˆI(u)

recalling that the projections ofgaref. Therefore, the thesis follows by

Corollary 1.3 p

1.2.Discerning the maximal transport rays.

As usual, also in the mass transportation is it of great interest to exhibit examples (and even more counterexamples!); however, except for situa- tions witha broad symmetry, it is not easy to find explicit examples: to be more precise, it is often easy to build examples where one guesses the right optimal transport plan or optimal transport map, but it may be not so easy actually to show their optimality. Therefore, it can be useful a tool allowing to claim in many situations that a particular plan or map is optimal.

(5)

In view of Lemma 1.4, we know that a transport plan is optimal if and only if it has «the right maximal transport rays», that is, if all its maximal transport rays are segments of maximal slope for someKantorovichpo- tential (equivalently, for allthe Kantorovich potentials): we will for sim- plicity refer to these segments as admissible maximal transport rays.

Hence, an useful tool could be a result asserting which are all the ad- missible maximal transport rays: having this at hand, to show that a plan is optimal one could simply check that all its maximal transport rays are admissible. Moreover, as we will discuss in Remark 2.12, the maximal tranpsort rays usually do not depend on the choice of the optimal transport plan.

A tool as the one we are looking for is already present in the literature, namely Lemma 8.1 in [2]:

LEMMA1.5[Horizontal transport rays]. Denote Kˆ[0;1], and let fbe concentrated respectively in KK and[5;6]K. We assume that

f‡([0;1][0;t])ˆf ([5;6][0;t]) 8t2K:

Then the only admissible maximal transport rays are the horizontal segments.

In words, this result says that a transport plan is optimal if and only if all its maximal transport rays are horizontal segments: this Lemma can be applied quite often, basically whenever one wants to build examples where an optimal transport plan moves everything horizontally. For instance, in [2] it was used to build a counterexample giving a negative answer to the following question, that was open: is it true that f‡ Hs with1<s<2 implies the existence of an optimal transport map? The question was in- teresting, since it was already known that the answer was negative for s1 and affirmative forsˆ2, and that the answer was affirmative also for 1<s<2 replacing the Euclidean distancej jin (1.1) by its squarej j2.

Our main result is a generalization of Lemma 1.5, which allows to consider muchmore situations; in fact, as we will discuss before Ex- ample 3.5, our results works in a broad range of cases. Before to state it, we need some notation; first of all, we give the following

DEFINITION1.6. Aray configuration is a family of pairwise disjoint open segments coveringV, each of them with extremes in@V. Given a ray configurationR ˆ fRi;i2Ig, we define its set of doubling pointsas the setDof all the points belonging to the closure of more than one Ri;i2I (of

(6)

course,D @V). Finally, acurve transversal toR (or simply atrans- versal curve) is a C1 injective curve onV intersecting each segment Ri exactly once and whose gradient at any point is not parallel to the direction of the unique ray passing through that point.

It is an easy geometric property that, given any ray configuration R ˆ fRi;i2Ig, it is possible to find a transversal curveH, and it is an open curve withboththe endpoints in@V, so that it dividesVin two parts, that we will refer to arbitrarily as the «left» and the «right» part, each of them intersecting all the Ri's; we equip then each segmentRi withthe orientation going from the left part to the right part of the segment. We give now the following last definitions:

DEFINITION1.7. Given any Borel set SV, we define F(S):ˆ[

Ri: i2I; Ri\S6ˆ ; ;

then, we say that the configuration R is balanced if for any SV the following holds (see also the remark below):

f‡(S)f F(S)

; f (S)f‡ F(S)

: …1:4†

In the above definition, the inequality (1.4) makes sense only ifF(S) is measurable withrespect to bothf‡andf ; in fact, as an easy consequence of the classical Projection Theorem (see for instance [6]) one can deduce more in general thatF(S) is universally measurable, therefore in particular measurable withrespect to bothf‡andf .

We will deduce many fruitful informations about the behaviour off withrespect to the ray configurationRin the balanced case in Lemma 2.5.

We also point out that, in fact, the two possible orientations «left-right» on the raysRiare determined uniquely byR, and not depend on the parti- cular choice of a transversal curveH. Finally, we state

THEOREM A. Let R ˆ fRi;i2Ig be a balanced ray configuration such that there exists two disjoint Borel sets G in V on which f are concentrated with the property that, for all i2I, supRi\G‡ infRi\G . Then, the admissible maximal transport rays are all the segments Ri;i2I and all their open subsegments.

We point out that Lemma 1.5 is the particular case when Vˆ[0;6][0;1] andR ˆ f[0;6] fig;i2[0;1]g, since of course in the hypotheses of Lemma 1.5 the configurationR is balanced.

(7)

In Section 2 we will prove Theorem A, while in Section 3 we will provide some examples using it. In particular, Example 3.5 shows that there are non-trivial situations withpresence of doubling points in the interior of sptf‡, that may be not obvious at first glance.

2. Proof of Theorem A.

This Section is entirely devoted to show Theorem A: to do that, we will exhibit a particular transport plan g and a particular 1-Lipschitz functionu; then, making use of Lemma 1.4, we will check that they are bothoptimal and we will deduce the position of all the admissible max- imal transport rays. In Subsection 2.1 we will defineu, in Subsection 2.2 we will definegand finally in Subsection 2.3 we will give the proof of the Theorem.

First of all, we fix a transversal curveHand we fix arbitrarily on it one of the two possible orientations. Notice that, by the obvious bijection be- tweenHandI, this allows to equipIwithan order, and thatIRwiththis order; therefore, in a purely formal way that will be quite useful in the sequel, we will consider also the indeces 1 on I corresponing to the empty raysR1, but we will also intendR‡1andR 1to be the two closed sets consisting of the single points supH and infH in @V, and we will consider also those two points as elements ofH.

WheneverCandDbelong toH, we denote byCDf the part of the curve H between C and D, that is CDf ˆ

P2H: CPD if CD and CDf ˆ

P2H: CPD ifCD. Therefore, we will write for brevity

CDƒ!

H1 CDf

if C>D;

0 if CˆD;

H1 CDf

if C<D:

8>

><

>>

:

Analogously, wheneverAandBbelong to a sameRi2R (recall that all theRiare oriented segments), we set

ABƒ!

AB if A>B;

0 if AˆB;

AB if A<B:

8>

<

>:

In addition, we will write„D

C

g(s)ds(resp.„B

A

g(s)ds) to denote the integral of

(8)

any functiongon the curveCDf (resp. on the segmentAB) withrespect to H1, multiplied by 1 or 1 if C<Dof C>D(resp. ifA<Bor A>B):

formally, …D C

g(s)ds:ˆCDƒ! …

CDf

g(s)dH1 H(s);

…B A

g(s)ds:ˆABƒ! …

AB

g(s)dH1(s):

…2:1†

2.1.Definition of u.

Here we define the functionu, and we check that it is 1 Lipschitz; we give first a

DEFINITION2.1. Let i:VnD !I the function which associates to any x2VnD the unique iˆi(x)2I such that R(i(x))3;x. Moreover, we set W:VnD !H as

W(x):ˆRi(x)\H;

so thatW(x)is the intersection of the ray passing through x with H. Finally, we letu:VnD !(0;p)be the function so thatu(x)is the angle between the (oriented) segment Ri(x)and the (oriented) tangent to H at Ri(x)\H.

We can finally define the functionuas we claimed: we fix a pointO2H, decidingu(O):ˆ0; then, we set -keep in mind (2.1)-

u(x):ˆ „x

0cos u(s)

ds 8x2H;

u(x):ˆu W(x)

W(x)xƒƒƒ!; 8x2VnH:

8>

<

>: …2:2†

REMARK2.2. By immediate geometrical arguments, the functions W andu are continuous onVnD; notice thatu(x) belongs to the open in- terval(0;p)by definition of transversal curve, and that it is constant on each ray Ri;i2I. By …2:2†, then, we infer that u is continuous on V. Notice that,a priori, u could happen not to be continuously extendable on

@V, in particular on the setD. However, we will show that u is1-Lipschitz, so that it can also be extended to the wholeV, still remaining1-Lipschitz.

The continuity ofuatV stated in the above remark is of course not satisfactory: since we mean to show thatuis a Kantorovichpotential, we have first to check thatuis 1-Lipschitz.

(9)

LEMMA 2.3. The function u defined in…2:2†is1-Lipschitz and it de- creases with slope1in all the rays Ri2R.

PROOF. The second part of the claim is immediate by the defini- tion (2.2), the difficult part is the first one. Assume that this is not true:

then, there must be two pointsAandBinVsuchthat u(A) u(B)>ljA Bj

withsomel>1. Consider now the segment connectingAandB: of course, there is some pointPin this segment with the property that

lim sup

Q!P;Q2AB

ju(Q) u(P)j jQ Pj >l;

…2:3†

we claim thatP=2H. Indeed, by the continuity ofunoticed in Remark 2.2 and by the definition (2.2) ofu, it is clear thatru(S) exists at eachS2H, and moreoverjru(S)j ˆ1 andru(S) is parallel toRi(S), so that (2.3) ensuresP not to be insideH. Without loss of generality, we assumePto be in the right part ofV; take now a pointQclose toPwith the property that

ju(Q) u(P)j>lPQ: …2:4†

we will show that this leads to a contradiction providedQis sufficiently close toP. To do that we give now some definitions, as in Figure 1 where of course we assume that our conventions for left and right, and for up and down are the obvious ones: we setPe:ˆW(P),Qe:ˆW(Q),e:ˆPƒ!eQe

andQb the point of Ri(Q) whose orthogonal projection on Ri(P) is P. Notice thate i(P)6ˆi(Q) by (2.4), since u has exactly slope 1 along each ray Ri, so that e6ˆ0.

Fig. 1. ± Position of points in Lemma 2.3

(10)

Moreover there is a suitable constantr>0 suchthat PQrjej

…2:5†

provided PQis sufficiently small: recalling thatP belongs toV by con- vexity, this follows immediately via a similitude argument sinceRi(P) and Ri(Q) do not intersect inV, and of course the constant rdepends on the angle between the segmentABand the gradient ofHatP, as well as on thee distance betweenPand@V(rbecomes smaller whenPgets closer to@V).

As a consequence, e!0 when Q!P. By (2.2), we evaluate u(P)ˆ

ˆu(P)e PPeandu(Q)ˆu(Q)e QQerecalling thatPandQare on the right ofH(concerningQ, this is of course true providedjQ Pj 1). Therefore, by (2.4) one infers

u(Q)e u(P)e ‡PP Qe Qe>lPQ:

…2:6†

Again by (2.2), we know that

u(Q)e ˆu(P)e …e

Q

eP

cos u(s) ds:

since, as noticed in Remark 2.2, the functionuis continuous, by the defi- nition ofewe deduce

u(Q)e ˆu(P)e ecos(u)‡o(e);

where we writeuin place ofu(P) for shortness. Substituting this estimatee in (2.6), one finds

PP Qe Qe ecos(u)‡o(e)>lPQ:

…2:7†

Again by the continuity ofuatP, we can also clearly estimatee QQeˆQQb‡Qƒ!eQb

; Qƒ!eQb

ˆ ecos(u)‡o(e); PeQb ˆesin(u)‡o(e);

…2:8†

putting then together (2.7) and (2.8), we have PP Qe Qb‡o(e)>lPQ:

…2:9†

To recover a contradiction withjejsufficiently small, we recall that by the triangle inequality

QQb PQ‡PQ;b …2:10†

(11)

moreover by Pitagora's Theorem we can evaluate, using also (2.8) and re- calling thatP6ˆPebecauseP2=H,

PQb ˆ



PPe2‡PeQb2 r

ˆ



PPe2‡ esin(u)‡o(e)2 r

ˆPPe‡o(e) that together with (2.10) gives

QQb PQ‡PPe‡o(e):

…2:11†

In the very same way, the triangle inequality ensures also PPe PQ‡QP;e

…2:12†

and again Pithagora's Theorem, recalling the continuity of u, gives QPeˆQQb‡o(e), so that from (2.12) one has

PPePQ‡QQb‡o(e):

…2:13†

Putting together (2.11) and (2.13), and substituting into (2.9) gives PQ>lPQ‡o(e);

that gives the desired contradiction whenjejis sufficiently small keeping in

mind (2.5). p

REMARK2.4. The function u has been defined only onV; however, by the previous Lemma it can be uniquely extended, remaining1 Lipschitz, to the wholeV.

2.2.Definition ofg.

In this subsection we will define the measuregwiththem we will prove the Theorem. First, let us briefly recall some well-known facts about the disintegration of the measures (for a formal treatment of this subject and for the proof of the claims below one can refer to [3]): given the spacesX andY, and measuresm2 M‡(X),my2 P(X) fory2Yandh2 M(Y), we write mˆmyh if for any Borel set DX the function y7!my(D) is h measurable and one has

m(D)ˆ …

y2Y

my(D)dh(y):

(12)

A particularly interesting situation when such a decomposition of m is possible, is when one is given a functiona:X!Y: in this case, the Dis- integration Theorem ensures that there exists probability measures my2 P(X), each of them concentrated in the correspondinga 1(fyg), such thatmˆmya#m; moreover, these measuresmyare uniquely determined for a#m a.e. y2Y. Finally, if there is a space Z and two functions b:X!Z andc:Z!Y suchthataˆcb, then the decomposition of b#mwithrespect togis given by

b#mˆb#myc# b#m : …2:14†

Now, our aim is to define a transport plangwith the property that each of its maximal transport rays is contained in some of theRi2R. We point out that, without any assumption on the ray configuration R, if such g exists thenR is balanced, so that the balance property is also necessary for the maximal transport rays. To notice this fact, just fix SV and recall that

f‡(S)ˆg

(x;y): x2S g

(x;y): y2F(S) ˆf F(S)

; indeed, by definition of maximal transport rays, whenever (x;y)2sptg and x2S then xy is contained in some ray Ri2R intersecting S (for instance, at x!), so y2F(S). Therefore, the first inequality in (1.4) is verified, and the second can be shown in the very same way.

For anyi2I, now, we consider the «upper set»

Li:ˆ[

ji

Rj;

and we prove some useful consequences of the balance property:

LEMMA2.5. Under the assumptions of Theorem A, take i2I and write Riˆ:AB2R; if both A and B do not belong toD, then f‡(Li)ˆf (Li).

More precisely, if there is no j<i such that A is the left endpoint of Rj(resp.

B is the right endpoint of Rj) then f‡(Li)f (Li)(resp. f (Li)f‡(Li)).

PROOF. DefiningSˆLin fBg, ifAis not the left endpoint of someRj with j<i then by construction F(S)ˆLi; therefore, since by the hy- potheses of the Theorem clearlyB=2G‡, (1.4) gives

f‡(Li)ˆf‡(S)f F(S)

ˆf (Li):

The case whenBis not the right endpoint ofRj for anyj<iis of course

(13)

completely similar; and the first claim, in whichAandBare not doubling

points ofR, is now a consequence. p

REMARK 2.6. For each i2I, at least one of the claims in the above Lemma applies. Indeed, since the rays inR coverVand are disjoint, it is immediately checked that, in case that both A and B belong to D, then either A is the left endpoint only of rays Rj with ji and B is the right endpoint only of rays Rjwith ji, or A is the left endpoint only of rays Rj with ji and B is the right endpoint only of rays Rj with ji.

Take now a point P2G‡\D: by geometrical arguments, there are imin <imax2I, depending onP, suchthatPis the left endpoint of all and only the rays Ri with iminiimax. We claim that the function d:(imin;imax]!Rgiven by

d(j):ˆf (Lj) f‡ Ljn fPg

is decreasing. Indeed, for imin<j1<j2imax we set Sˆ [fRj: j1 j<j2g, for which F(S)ˆLj1nLj2[ fPg; by the hy- potheses,P=2G , hencete also that again by (1.4) one immediately obtains that 0df‡(fPg) in (imin;imax] -we also underline that the fact that the interval is open in the left is fundamental to derive this estimate. Therefore, we extenddto the wholeRby settingd(j)f‡(fPg) ifjiminandd(j)0 ifj>imax, and we define the measuremP2 M‡(I) given bymP:ˆ Dd:

sincedis a bounded decreasing function,mPis a positive measure of mass f‡(fPg) concentrated in [imin;imax]. For anyi2I, we also defineP(n;i) as the point inRiˆABsuchthatAP(n;i)ˆAB=n. Finally, we introduce the measures

fn‡:ˆf‡ G‡nD

‡ X P2D\G‡

imax…(P)

iˆimin(P)

dP(n;i)dmP(i):

…2:15†

The meaning ofmPandfn‡is worthto be understood:mPsays how the part of f‡contained in the pointPshould be distributed among the different rays of Rin order to have a precise mass balance. The measuresfn‡simply perform this distribution for all the doubling points, shifting a corresponding part of mass in the interior of the rays, but remaining arbitrarily close to the dou- bling points in the limitn! ‡ 1. Notice that the definition (2.15) makes sense since the functioni7!P(n;i) is continuous and the setD\G‡ is at most countable. Our last ingredient is the following measure.

(14)

DEFINITION2.7. We define the measurey2 P(I)settingy [i;‡1) for any i2Ras follows:

y [i;‡1) :ˆ

f‡(Di) if the left endpoint ofRiis not left endpoint of anyRjwithj<i;

f (Di) if the right endpoint ofRiis not right endpoint of anyRjwithj<i:

Notice that the two definitions agree where they apply simultaneously thanks to Lemma2:5(recall also Remark2:6).

REMARK 2.8. We point out that, if D is f‡ negligible, then W:VnD !H is in fact defined f‡ a.e., so that it is possible to define the measure W#f‡. Moreover, by Definition2:7 and by construction, it is clearlyyˆW#f‡; in the very same way,yˆW#f if f (D)ˆ0.

Finally, we are ready to prove the fundamental

LEMMA2.9. The measures fn‡ are probability measures weakly* con- verging to f‡; moreover, there are probability measures fn;i‡ concentrated on Rifor n2N, i2I such that one has the disintegration fn‡ˆfn;i‡ y.

Finally, for any i2I one has fn;i‡ * fi‡, and f‡ˆfi‡y with fi‡ con- centrated in Ri. The same statement allows to write f ˆfi ywith fi concentrated in Ri.

PROOF. Sincef‡is a probability measure andfn‡coincides withf‡up to splitting the mass of anyP2G‡\D in the corresponding rays (recall jmPj ˆf‡(fPg)), then alsofn‡is a probability measure. Moreover, since the pointsP(n;i) uniformly converge to the left endpoints of the raysRi, th en the convergence fn‡* f‡ follows. Notice now that by constructionfn‡ is concentrated in VnD, and as noticed in Remark 2.8 it is clear that W#fn‡ˆy. Therefore, we can disintegrate obtainingfn‡ˆfn;i‡ ywith fn;i‡ concentrated inRi.

By construction, for anyi2Ithe sequence of measuresn7!fn;i‡ is the sum of a constant measure and a Dirac mass in a point moving to the left extreme ofRi, so thatfn;i‡ * fi‡withfi‡concentrated inRi; since thefn;i's are probability measures, the Dominated Convergence Theorem ensures that thenfn;i‡ y* fi‡y, so that also the second part of our claim follows.

The result for f can be derived in the same way, replacing in the previous argument and in the definition ofdthe left with the right andf‡

withf . p

(15)

REMARK2.10. We like to mention that the previous claim, i.e. that one can write fˆfiywith probability measures fi‡on Ri, is all we need to perform our construction ofg; therefore, the whole discussions we made in this subsection until now would be completely unnecessary under the assumption thatD is negligible with respect to both f‡ and f , since in that case one could directly decompose f with respect to W. Notice also that by constructionD contains countable many points, so that it would have been sufficient to add to Theorem A the hypothesis of f;being non- atomic. However, since our result is true without this assumption, we decided to present this more involved argument.

Finally, we can give our construction of the transport plang.

LEMMA 2.11. There exists a transport plangbetween f‡and f such that any maximal transport ray ofgis contained in some Ri;i2I.

PROOF. We will define

g:ˆgiy …2:16†

beinggia transport plan betweenfi‡andfi (we writefˆfiyin view of Lemma 2.9); any choice ofgiworks, provided of course thati7!giis a y measurable measure valued map, in the sense of [1]; in other words, since definingg:ˆgiymeans that for anyv2Cb(VV) one sets

hg;vi ˆ … I

hgi;vidy(i);

we need to check that the preceding integral is well defined, that is that i7!hgi;vi is a y measurable real map for any continuous and bounded functionvonVV. To solve easily this problem, it suffices to define

gi:ˆfi‡fi : …2:17†

giis of course a transport plan betweenfi‡andfi , andi7!giis easily seen to be ay measurable measure valued map because so arei7!fi‡andi7!fi by the properties of disintegration; indeed, i7!hgi;vi is clearly y measurable ifvis the characteristic function of a setV1V2whereV1 andV2 are Borel subsets of V, and by a standard density argument one concludes for a generalv.

Having definedgvia (2.16) and (2.17), we need to check that it satisfies our claim; in fact, by (2.14) one hasp1gˆp1giyˆfi‡yˆf‡ and ana-

(16)

logouslyp2gˆf , so thatgis a transport plan betweenf‡andf . More- over,g a.e. pair (x;y) inVV is contained in sptgifor some i2I, then bothxandyare inRi; this immediately implies that all the transport rays are contained in someRi, and then also the maximal transport rays do the same (recall that transport rays and maximal transport rays are open

segments). p

2.3.Proof of the Theorem.

Here we can show Theorem A.

PROOF. of Theorem A:We denote by uthe function defined in (2.2) and bygthe transport plan built in Lemma 2.11: by construction, all the maximal transport rays ofgare contained in some of the raysRi, and on the other hand we know by Lemma 2.3 that udecreases withslope 1 in all these rays (recall also Remark 2.4). Hence, we can apply Lemma 1.4 to derive thatuis a Kantorovichpotential andgis an optimal transport plan.

Recalling the discussions at the beginning of Section 1.2, we know that then the admissible maximal transport rays are exactly those segments on whichudecreases at slope 1. Since the raysRi2Rcover the wholeV, we derive that these segments are exactly all the subsegments of some

Ri2R, and the proof is completed. p

REMARK 2.12. Theorem A says which are the admissible maximal transport rays, and thanks to the discussions in Section1:2 this tells us which of the transport plans are optimal and which are not. However, usually something more is true, that is, that the maximal transport rays are the same for all the different optimal transport plans: indeed, ifD is f negligible (in fact, it suffices thatDis negligible with respect to just one between f‡ and f ), then the maximal transport rays are exactly all the maximal subsegments of the Ri's intersecting the supports of f‡and f . If D is not negligible, the unicity of the maximal transport rays is still true under the assumptions of our Theorem, since as we proved in Section2:2it can be precisely determined how much of each doubling point must be moved along each ray. But in the completely general case, this unicity may not be true, since the different optimal transport plans could split differ- ently the doubling points in the different admitted directions. For instance, in Example3:2the maximal transport rays associated tog0are the vertical segments AC and DB, the maximal transport rays associated tog1are the horizontal segments AB and DC, and the maximal transport rays asso- ciated togswith0<s<1are all the four segments.

(17)

3. Examples.

In this section, we give several examples to comment our Theorem and to show some of its possible applications. First of all, we give an easy ex- ample to enlighten the necessity of introducing the setsGin the claim of the Theorem.

EXAMPLE3.1. Let f‡ andf be concentrated in two polygons as in Figure 2.a); more precisely, letfbe absolutely continuous withrespect to the Lebesgue measureL, withdensities given by

fL‡(x; y):ˆ 1 in[0;1][0;1=2];

1=2 in[0;2][1=2;1]:

(

fL(x; y):ˆ 1=2 in[1;3][0;1=2];

1 in[2;3][1=2;1]:

(

Then one can apply Theorem A with the horizontal rays and with GˆsptfnR0, beingR0 :ˆ[0;3] f1=2g, and it follows that the admis- sible maximal transport rays are the horizontal segments (thus this could have been recovered also through a stronger version of Lemma 1.5).

Notice that the condition supRi\G‡infRi\G would have not been satisfied replacing G withsptf: this shows the importance of introdu- cing the setsGin our statement also for very simple situations.

Now, consider once more the meaning of the Theorem: roughly speaking, if one has the insight of which are the maximal transport rays, it allows to show it formally, and this seems to be the easiest way to give explicit examples of optimal transport plans (or maps), except than in trivial cases. In other words, there is a certain class of pairs f (and of corresponding setsV) for which the Theorem can be applied to reveal the maximal transport rays -of course, after that one has guessed them. This class was very little after Lemma 1.5 in [2], since it covered only some

Fig. 2. ± Examples 3.1, 3.2 and 3.3

(18)

situations when all the transport rays are horizontal; with our result at hand, this class has become much larger, as we will try to show mainly through Examples 3.4 and 3.5. But of course this class can not coverallthe possible configurationsf, as we can understand withthe following two examples.

EXAMPLE 3.2. Consider the situation when f‡ˆ1=2(dA‡dD) and f ˆ1=2(dB‡dC) and the four pointA; B; C andDare the vertexes of a square as in Figure 2.b). Then it is clear that the transport plan are exactly the measures

gsˆs dAdB‡dDdC

‡(1 s) dAdC‡dDdB

with0s1. Observe that, for any convex setVcontaining the square, it is not possible to find a ray configuration satisfing the conditions of the Theorem: in particular, there are balanced configurations (for instance, the one made by all horizontal lines), but the condition supRi\G‡ infRi\G can not be satisfied because whenever A is on the left of its ray then D is on the right of its one. Thus, the Theorem can not be applied.

A less critical situation in which the Theorem still can not be applied is the following one:

EXAMPLE 3.3. Consider, as in Figure 2.c), the situation when f‡ and f are (up to the constant 1=2 to ensure jfj ˆ1) the Lebesgue measure restricted on the sets [0;1][[3;4]

[0;1] and [1;3][0;1]

respectively: then, it is easy to understand that the maximal transport rays are all the open segments (0;2) fsgand (2;4) fsgfor0s1, so one can not find any ray configuration with segments having the endpoints in @V.

Now we give a simple example, to show a situation that can be handled with Theorem A and could have not been solved with Lemma 1.5 (this

Fig. 3. ± Example 3.4

(19)

example has been already presented in [8] because it admits an optimal transport map which is not continuous).

EXAMPLE 3.4. Let V:ˆ[0;3][0;1], and let R :ˆ fRi; i2[0;1]g [ [ fRei; i2[0;1]g, being Ri (resp. Rei) the open segment connecting (0;0) and (3; i) (resp. (0; i) and (3;1)); Figure 3 sh ows some of th ese rays. We let f‡ and f be concentrated on [0;1][0;1] and [2;3][0;1] and abso- lutely continuous withrespect to the Lebesgue measure, withdensities given by

fL‡(x; y):ˆ C‡(i) in[0;1][0;1]\Ri; Ce‡(i) in[0;1][0;1]\Rei; (

fL(x; y):ˆ C (i) in[2;3][0;1]\Ri; Ce (i) in[2;3][0;1]\Rei (

and where the Borel functions C;Ce:[0;1]!R‡ can be arbitrarily chosen subject to the constraint

C‡5C ; Ce 5Ce‡:

…3:1†

We claim that Theorem A can be applied to this situation, and hence that the maximal transport rays are exactly theRiand theRei. To sh ow th is assertion, one has only to check that the mass balance condition holds, then by symmetry we limit ourselves to show thatf‡(T)ˆf (T) for th e triangle T withvertexes (0;0), (3;0) and (3; i) for a generici2[0;1]. A trivial calculation ensures that f‡(T)ˆ1=6„i

0C‡(t)dt and f (T)ˆ

ˆ5=6„i

0C (t)dt, so that thanks to (3.1), our claim follows.

We give now our last example, which will be a more involved appli- cation of Theorem A and it will also give an interesting counterexample.

To show its meaning, we first underline that our Theorem can not clearly be applied whenever there is some doubling point in the interior of the support off‡or off , as it happens in Example 3.3. From this example and other simple situations, it could seem reasonable that wheneverf are sufficiently regular and withsupports convex and disjoint, there may not be doubling points in the interior of the supports: notice that this would imply that the Theorem can be applied in all these situations. On the countrary, in the example below we present a situation where the supports offare convex and disjoint andfare regular, but there is a doubling point in the interior of sptf‡. We remark that as a consequence

(20)

the Theorem can not be applied in this situation: nevertheless, we will prove our claim witha careful application of our result in some subset of the supports.

EXAMPLE 3.5. Let f‡ and f , as in Figure 4.a), be two absolutely continuous measures concentrated on the squares [0;2][ 1;1] and [4;6][ 1;1], whose densities (up to the rescaling factor4ˆ L(sptf‡)ˆ L(spt(f )) are given by

fL‡(x;y):ˆ1; fL(x;y):ˆ 2 e in [4;6] [ 1; 1=2][[1=2;1]

; e in [4;6][ 1=2;1=2];

(

whereeis a small positive number. An immediate symmetry argument shows that whenever g is an optimal transport plan and (x;y)2sptg, if x2>0 th en y2 >0 and if x2<0 th en y2<0. Therefore, we con- sider the transport problem associated to ~f‡:ˆf‡ fx2 0g and

~f :ˆf fx2 0g, and the admissible maximal transport rays for the original problem will be all the admissible maximal transport rays for the restricted problem and all their symmetric images with respect to the axis fx2 ˆ0g. We will apply Theorem A to the restricted config- uration. To this aim, for any 0s3 and any 0t1 we define th e points xs2@spt~f‡ and yt 2@spt~f as

xs:ˆ 0;1 s

if 0s1;

s 1;0

if 1s3; yt :ˆ 6;1 t

; (

notice that the set of the points fxs;0s3g is the left and the bottom part of the boundary of spt~f‡, wh ile th e set of th e points fyt;0t1=2g is the right part of the boundary of spt~f . We also define the functionW:[0;3][0;1]!R2 as follows: the first (resp. the second) component of W(s;t) is the measure w.r.t. ~f‡ (resp. to ~f ) of the region above the segment connectingxs and yt. We state now th e

Fig. 4. ± Example 3.5

(21)

CLAIM.There is a strictly increasing continuous functionst:[0;1=2]!

![0;3]witht(0)ˆ0with the property thatW1(t(t);t)ˆW2(t(t);t)for any 0t1=2

Before to show the claim, notice that an immediate consequence is that t(1=2)<3 sinceW1(3;1=2)ˆ2<W2(3;1=2): consistently, in Figure 4.b) we drawn xaˆt(ya), xbˆt(yb) and vˆt(w) withw(6;1=2). Extend now the functiontto (1=2;1] settingt(t):ˆt(1=2) for 1=2<t1, and consider the ray configurationR made by all the segmentsxt(t)yt for 0t1 (as the segmentvyc in the figure): since, for anyt>1=2, in the triangle with vertexes v;w and (6;0) the densities of ~f‡ and ~f (where positive) are costantly 1 ande, by the Claim it follows thatR is balanced. Hence, we can clearly apply the Theorem, finding that the elements ofR are exactly the maximal transport rays for the problem with data~f‡ and~f . As noticed before, a symmetry tells us also which are the maximal transport rays for the original problem: therefore, we found that the point v is a doubling point inside the interior of sptf‡! We remark also thatvclearly converges to (2;0) whene!0; moreover, the discontinuity of (the density of)f on the lines fx2ˆ 1=2g is clearly not important in this example, and one could easily replacefwithprobability measures having smoothdensities.

We conclude now this example showing the claim.

PROOF OF THECLAIM: One could easily obtain the proof through trivial but very boring calculations; nevertheless, we prefer to give now a simple ab- stract proof. First of all, notice that for any 0<t1=2 one has clearly W1(0;t)<W2(0;t) butW1(3;t)ˆ2>W2(3;t): therefore, by continuity one has that for anytthere exists someswithW1(s;t)ˆW2(s;t) (also fortˆ0 since of courseW(0;0)ˆ(0;0)). We want to show that for anytthere is a unique such s:ˆt(t), and for anys there is at most a unique t1=2 suchthat W1(s;t)ˆW2(s;t): as a consequence, the claim will immediately follow.

Concerning the uniqueness ofs for a givent, assume by contradiction the existence of 0t1=2 and 0s1<s23 withW1(si;t)ˆW2(si;t) for bothiˆ1;2. Assume first thats21, and denote byTthe triangle with vertexesxs1;xs2 andyt: it should be~f‡(T)ˆ~f (T), but by similitude one immediately hasL T\spt~f‡

>2L T\spt~f

, and since the density of

~f‡ on T\spt~f‡ is costantly 1 while the density of ~f‡ onT\spt~f‡ is everywhere less than 2, one finds the contradiction, then it must bes2>1.

We can then assume the existence of 1s<s23 with W1(s;t)W2(s;t);W1(s2;t)ˆW2(s2;t): …3:2†

(22)

indeed, ifs11 the choices:ˆs1 clearly works and the «» above is in fact «ˆ», while ifs1<1 then the same similitude argument as before en- sures that the choices:ˆ1 works and the «» is a «>». We will obtain that (3.2) is absurd by showing that W1(;t) and W2(;t) are respectively concave and convex in the interval [s;3] and recalling that W1(3;t)ˆ

ˆ2>W2(3;t). To obtain this, just call Ts the triangle with vertexes yt,xs

andxs for a genericss3: then the concavity ofW1 follows by the ob- vious geometric fact thats7!L Ts\spt~f‡

ˆ~f‡ Ts

is (strictly) concave on [s;3]. On the other hand, the convexity ofW2is true since in [s;3] the map

s7!L Ts\ fp2R2: the density of ~f atpis 2 eg is linear, while the map

s7!L Ts\ fp2R2: the density of ~f atp iseg is (strictly) convex.

Finally, concerning the uniqueness oftfor a givens, assume that there exists s2[0;3] and 0t1<t21=2 with W1(s;ti)ˆW2(s;ti) for both iˆ1;2. The conclusion follows in a very similar way as before: we have that W1(s;t1)ˆW2(s;t1), and clearly W1(s;1)W2(s;1) (the inequality is strict if and only if s<1). The absurd (which concludes the proof of the claim and the example) is given by the fact that on [t1;1] the functionW1(s;) is linear, whileW2(s;) is strictly concave on [t1;1=2] and linear on [1=2;1].

REFERENCES

[1] L. AMBROSIO,Lecture Notes on Optimal Transport Problems,inMathema- tical Aspects of Evolving Interfaces, Lecture Notes in Mathematics, LNM 1812, Springer (2003), pp. 1-52.

[2] L. AMBROSIO- A. PRATELLI,Existence and stability results in the L1theory of optimal transportation, inOptimal Transportation and Applications, Lec- ture Notes in Mathematics, LNM1813, Springer (2003), pp. 123-160.

[3] L. AMBROSIO- N. FUSCO- D. PALLARA,Functions of Bounded Variation and Free Discontinuity Problems, Oxford University Press (2000).

[4] G. BOUCHITTEÂ- G. BUTTAZZO,Characterization of optimal shapes and masses through Monge-Kantorovich Equation, J. Eur. Math. Soc.,3(2001), pp. 139- [5] G. B168.OUCHITTEÂ- G. BUTTAZZO P. SEPPECHER,Shape optimization solutions via Monge-Kantorovich equation, C.R.Acad. Sci. Paris,324-I(1997), pp. 1185- 1191.

[6] C. CASTAING - M. VALADIER, Convex analysis and measurable multifunc- tions, Lecture Notes in Mathematics,580, Springer (1977).

(23)

[7] L.C. EVANS - W. GANGBO, Differential equations methods for the Monge- Kantorovich mass transfer problem, Memoirs of the A.M.S., Vol.137, Number 653, (1999).

[8] I. FRAGALAÁ, M.S. GELLI- A. PRATELLI,Continuity of an optimal transport in Monge problem, to appear on JMPA.

[9] L.V. KANTOROVICH,On the transfer of masses, Dokl. Akad. Nauk. SSSR,37 (1942), pp. 227-229.

[10] L.V. KANTOROVICH,On a problem of Monge, Uspekhi Mat. Nauk.,3(1948), pp.

225-226.

[11] G. MONGE, Memoire sur la Theorie des DeÂblais et des Remblais, Hist. de l'Acad. des Sciences de Paris (1781).

[12] A. PRATELLI, Existence of optimal transport maps and regularity of the transport density in mass transportation problems, Ph.D. Thesis, Scuola Normale Superiore, Pisa, Italy (2003). Avalaible on http://cvgmt.sns.it/ . [13] S.T. RACHEV- L. RUÈSCHENDORF,Mass Transportation Problems, Springer-

Verlag (1998).

Manoscritto pervenuto in redazione il 25 novembre 2004.

Références

Documents relatifs

In a three-dimensional chaotic system the separation be- tween two initially adjacent field lines first increases expo- FIG. The symbols and vertical dashed lines are as in Fig.

Then files may be copied one at a time (or with a wild card transfer) to the dual density diskette.. Using the Filer under the UCSD O/S, do an E)xtended listing of the files on

We give then locally necessary and sufficient conditions (Theorem 3.2 for interior vertices and Theorem 2.2 for vertices on straight parts of the boundary) on a function ^: E~^ —&gt;

To prove that the sequence (a n ) n≥0 converges in the weak ∗ topol- ogy of A , since this sequence is bounded, it is enough to prove that it has only one weak ∗ cluster

Theorem 3.5: Any possibly discontinuous PWA function, defined over a polyhedral partition of a polyhedron, can be equivalently obtained as a selection among the optimal solu- tions of

In the global theory with irreducible automorphic representations, if we fix the archimedean data (the “weight”) then the space of invariant vectors under a compact open subgroup of

A new abstract characterization of the optimal plans is obtained in the case where the cost function takes infinite values.. It leads us to new explicit sufficient and necessary

In Section 4 we choose one such finite plan π 0 which need not be an optimal plan and prove a dual attainment result for a modified Monge–Kantorovich problem which is restricted to