• Aucun résultat trouvé

2.1. The branched transportation energy

N/A
N/A
Protected

Academic year: 2022

Partager "2.1. The branched transportation energy"

Copied!
27
0
0

Texte intégral

(1)

DOI:10.1051/cocv/2015049 www.esaim-cocv.org

UNIFORM ESTIMATES FOR A MODICA–MORTOLA TYPE APPROXIMATION OF BRANCHED TRANSPORTATION

Antonin Monteil

1

Abstract.Models for branched networks are often expressed as the minimization of an energyMαover vector measures concentrated on 1-dimensional rectifiable sets with a divergence constraint. We study a Modica–Mortola type approximationMεα, introduced by Edouard Oudet and Filippo Santambrogio, which is defined over H1 vector measures. These energies induce some pseudo-distances between L2 functions obtained through the minimization problem min{Mεα(u):∇ ·u=f+−f}. We prove some uniform estimates on these pseudo-distances which allow us to establish a Γ-convergence result for these energies with a divergence constraint.

Mathematics Subject Classification. 49J45, 90B06, 90B18.

Received March 11, 2015. Accepted September 30, 2015.

1. Introduction

Branched transportation is a classical problem in optimization: it is a variant of the Monge–Kantorovich optimal transportation theory in which the transport cost for a massmper unit of length is not linear anymore but sub-additive. More precisely, the cost to transport a massmon a lengthlis considered to be proportional tomαl for someα∈]0,1[. As a result, it is more efficient to transport two masses m1andm2together instead of transporting them separately. For this reason, an optimal pattern for this problem has a “graph structure”

with branching points. Contrary to what happens in the Monge–Kantorovich model, in the setting of branched transportation, an optimal structure cannot be described only using a transport plan, giving the correspondence between origins and destinations, but we need a model which encodes all the trajectories of mass particles.

Branched transportation theory is motivated by many structures that can be found in the nature: vessels, trees, river basins. . . Similarly, as a consequence of the economy of scale, large roads are proportionally cheaper than large ones and it follows that the road and train networks also present this structure. Surprisingly the theory has also had theoretical applications: recently, it has been used by Bethuel in [4] so as to study the density of smooth maps in Sobolev spaces between manifolds.

Branched transportation theory was first introduced in the discrete framework by Gilbert in [14] as a gener- alization of the Steiner problem. In this case an admissible structure is a weighted graph composed of oriented edges of length li on which some mass mi is flowing. The cost associated to it is then

ilimαi and it has to be minimized over all graphs which transport some given atomic measure to another one. More recently, the branched transportation problem was generalized to the continuous framework by Xia in [23] by means

Keywords and phrases.Branched transportation networks,Γ-convergence, phase field models.

1 Laboratoire de Math´ematiques d’Orsay, Universit´e Paris-Sud 11, Bˆat. 425, 91405 Orsay, France.antonin.monteil@gmail.com

Article published by EDP Sciences c EDP Sciences, SMAI 2016

(2)

of a relaxation of the discrete energy (see also [24]). Then, many other models and generalizations have been introduced (see [15] for a Lagrangian formulation, see also [1–3] for different generalizations and regularity properties.). In this paper, we will concentrate on the model with a divergence constraint, due to Xia. However, this is not restrictive since all these models have been proved to be equivalent (see [3,19]).

In this model, a transport path is represented as a vector measure u on some open set Ω Rd such that

∇·u=μ+−μfor two probability measuresμ+andμ. Then the energy ofuis defined asMα(u) = MθαdH1 ifuis a vector measure concentrated on a rectifiable 1-dimensional setM on whichuhas multiplicityθ w.r.t.

the Hausdorff measure (see [3] for more details). In this framework,umust be considered as the momentum (the mass θtimes the velocity) of a particle at some point. Then (∇ ·u)(x) represents the difference between incoming and outcoming mass at each pointx.

In this paper, we are interested in some approximation of branched transportation proposed by Oudet and Santambrogio few years ago in [18] and which has interesting numerical applications. This model was inspired by the well known scalar phase transition model proposed by Modica and Mortola in [16]. Givenu∈H1(Ω,Rd) for some bounded open subsetΩ⊂Rd, Oudet and Santambrogio introduced the following energy:

Mεα(u) =εγ1

Ω

|u|β+εγ2

Ω

|∇u|2,

whereβ (0,1) andγ1, γ2>0 are some exponents depending onα(see (2.4)). Ifudoes not belong to the set H1(Ω), the value ofMεα is taken as +.

We recall the heuristic which shows whyMεαis an approximation ofMα(see [18]): assume thatμ (resp.μ+) is a point source atS1(resp.S2) with massm. Then, it is clear that the optimal path forMαbetween these two measures is the oriented edgeS= (S1, S2) of lengthlwith a massmflowing on it. We would like to approximate this structure, seen as a vector measureuconcentrated onS, by someH1vector fieldsv which are more or less optimal forMεα. What we expect is thatv looks like a convolution ofuwith a kernelρdepending onεandm:

v=u∗ρR for someR=R(ε, m), where

ρR(x) =Rdρ R−1x

(1.1) for some fixed smooth and compactly supported radial kernelρ∈ Cc (Rd). Then the support ofv is like a strip of width R around S so that |v| is of the order of m/Rd−1 and |∇v| is of the order of m/Rd. This gives an estimate ofMεα(v) like

Mεα(v)εγ1Rd−1

m/Rd−1β

l+εγ2Rd−1

m/Rd2

l. (1.2)

With our choice for the exponentsγ1,γ2 andβ, the optimal choice forRis

R=εγm1−γd−1, (1.3)

where

γ= 2

2d−β(d−1) = γ2

d+ 1· (1.4)

This finally leads toMεα(v)mα as expected.

It was proved in [18] that, at least in two dimensions, the energy sequence (Mεα)ε>0 Γ-converges to the branched transportation functionalc0Mα for some constantc0 and for some suitable topology (see Thm. 2.1, p. 313). This result has been interestingly applied to produce a numerical method. However, rather than aΓ- convergence result onMεαwe would need to deal with the functionalsMαε, obtained by adding a divergence con- straint: it should be shown thatMαε(u) :=Mεα(u)+I∇·u=fε Γ-converges toc0Mα(u) :=c0Mεα(u)+I∇·u=μ+μ, where fε ∈L2 is some suitable approximation of μ+−μ and IA(u) is the indicator function in the sense of convex analysis that is 0 whenever the condition is satisfied and +∞otherwise. Even if this property was not

(3)

proved in [18], the effectiveness of the numerical simulations made the authors think that it actually holds true.

Note that an alternative using a penalization term was proposed in [20] to overcome this difficulty.

In Section2we recall Xia’s formulation of branched transportation and its approximationMεαintroduced by Oudet and Santambrogio. The longest part of this paper, Section3, is devoted to a local estimate which gives a bound on the minimum valuedαε(f+, f) := min{Mεα(u) : ∇·u=f}depending onfL1,fL2and diam(Ω) (see Thm.3.2, p.314). In Section4, we deduce a comparison betweendαε and the Wasserstein distance with an

“error term” involving theL2 norm off+−f. As an application of this inequality, in the last section, we will prove the followingΓ-convergence result which was lacking in [18]

Theorem 1.1. Let (fε)ε>0 ⊂L2(Ω) be a sequence weakly converging to μ as measures when ε→ 0. Assume that the sequence(fε)ε>0 satisfies

Ω

fε(x) dx= 0 and εγ2fε2L2−→

ε→00.

There exists a constant c0 > 0 such that the functional sequence (Mαε)ε>0 Γ-converges to c0Mα as ε 0.

Moreover c0 is the minimum value for the minimizing problem (5.2).

This answers the open question 1 in [18,20] and validates their numerical method.

2. Mathematical setting

2.1. The branched transportation energy

In all what follows, we will use the model proposed by Xia (see [23,24]):

Letd≥1 be an integer andΩbe some open and bounded subset ofRd. Let us denote byMdiv(Ω) the set of finite vector measures onΩsuch that their divergence is also a finite measure:

Mdiv(Ω) :=

umeasure onΩvalued in Rd : uMdiv(Ω)<+∞

, whereuMdiv(Ω):=|u|(Ω) +|∇ ·u|(Ω) with

|u|(Ω) := sup

Ω

ψ·du : ψ∈ C(Ω,Rd),ψL1

and, similarly,

|∇ ·u|(Ω) := sup

Ω

∇ϕ·du : ϕ∈ C1(Ω,R), ϕL1

.

In all what follows,∇ ·u has to be thought in the weak sense,i.e. ϕ∇ ·u=− ∇ϕ·dufor allϕ∈ C1(Ω).

Since we do not askϕ to vanish at the boundary,∇ ·umay contain possible parts on∂Ω which are equal to u·nwhen u is smooth, wheren is the external unit normal vector to∂Ω. In other words, ∇ ·uis the weak divergence of u1Ω in Rd, where 1Ω is the classical indicator function ofΩ, equal to 1 onΩ and 0 elsewhere.

From now on, the notation1X for the classical indicator function of a setX and IX for the indicator function in the sense of convex analysis (equal to 1 inside and +∞outside) will be used.Mdiv(Ω) is endowed with the topology of weak convergence onuand on its divergence:i.e.unMdiv(Ω)

−→ uifun uand∇ ·un∇ ·uweakly as measures.

Given 0 < α < 1, the energy of branched transportation can be represented as follows for measures u Mdiv(Ω):

Mα(u) = MθαdH1 ifucan be written asu=U(M, θ, ξ),

+ otherwise, (2.1)

(4)

whereU(M, θ, ξ) is the rectifiable vector measureu=θξ·H1|M with densityθξwith respect to theH1-Hausdorff measure on the rectifiable setM. The real multiplicity is a measurable functionθ:M R+and the orientation ξ:M →Sd−1Rd is such thatξ(x) is tangential toM forH1-a.e.x∈M.

Given two probability measuresμ+ andμ onΩ, the problem of branched transportation consists in mini- mizingMα under the constraint∇ ·u=μ+−μ:

inf

Mα(u) : u∈ Mdiv(Ω) and ∇ ·u=μ+−μ

. (2.2)

Note that, ifμ±(∂Ω) = 0, the divergence constraint implies a Neumann condition onu:u·n= 0 on∂Ω.

2.2. Functionals M

εα

For the minimum value in (2.2) to be finite whatever μ+ andμ in the set of probability measures, we will requireαto be sufficiently close to 1. More precisely, we make the following assumption:

11

d < α <1. (2.3)

Xia has shown in [23] that, under this assumption, there exists at least one vector measureu∈ Mdiv(Ω) such thatMα(u)<+∞.

We are interested in the following approximation of Mα which was introduced in [18]: for allu∈ Mdiv(Ω) and for all open subsetω⊂Ω,

Mεα(u, ω) :=

⎧⎨

εγ1

ω

|u(x)|βdx+εγ2

ω

|∇u(x)|2dx ifu∈H1(ω)

+ otherwise,

(2.4) whereβ,γ1andγ2 are three exponents depending onαanddthrough:

β= 22d+ 2αd 3−d+α(d−1) and

γ1= (d1)(1−α) and γ2= 3−d+α(d−1).

Note that inequality 11/d < α <1 implies that 0< β <1. Whenω=Ω, we simply write Mεα(u, Ω) =:Mεα(u).

We point out the 2-dimensional case whereMεαrewrites as Mεα(u) =εα−1

Ω

|u(x)|βdx+εα+1

Ω

|∇u(x)|2dx, (2.5) whereβ =4αα+1−2.

Given two densities f+, f ∈L2+(Ω) :={f ∈L2(Ω) : f 0} such that f+ = f, we are interested in minimizingMεα(u) under the constraint∇ ·u=f+−f:

inf

Mεα(u) : u∈H1(Ω) and ∇ ·u=f+−f

. (2.6)

The classical theory of calculus of variation shows that this infimum is actually a minimum. A natural question that arises is then to understand the limit behavior for minimizers of these problems whenεgoes to 0. A classical tool to study this kind of problems is the theory ofΓ-convergence which was introduced by De Giorgi in [12]. For the definition and main properties ofΓ-convergence, we refer to [8,11]. In particular, ifMεαΓ-converges to some energy functionalM0αand if (uε) is a sequence of minimizers forMεαadmitting a subsequence converging tou, then,uis a minimizer forM0α. By construction ofMεα, we expect that, up to a subsequence,Mεα Γ-converges toc0Mα. In the two dimensional case, we have the followingΓ-convergence theorem proved in [18].

(5)

Theorem 2.1. Assume that d= 2 and α∈ (1/2,1). Then, there exists a constant c >0 such that (Mεα)ε>0

Γ-converges tocMα inMdiv(Ω)whenεgoes to0.

Nevertheless, this does not imply theΓ-convergence ofMεα(u) +I∇·u=f+f toMεα(u) +I∇·u=f+f. Indeed, theΓ-convergence is stable under the addition of continuous functionals but not l.s.c. functionals. Consequently, we cannot deduce, from this theorem, the behavior of minimizers for (2.6). For instance, it is not clear that there exists a recovery sequence (uε), i.e. uε converges tou in Mdiv(Ω) and Mεα(uε) converges toMα(u) as ε→0, with prescribed divergence∇ ·uε=f+−f. To this aim, we require some estimates on these energies and this is the purpose of this paper.

2.3. Distance of branched transportation

We remind our hypothesis 11/d < α < 1. In [23], Xia has remarked that, as in optimal transportation theory,Mα induces a distancedα on the spaceP(Ω) of probability measures onΩ:

dα+, μ) = inf

Mα(u) : u∈ Mdiv(Ω) such that ∇ ·u=μ+−μ ,

for all μ+, μ ∈ P(Ω). Thanks to our assumption α >11/d, dα is finite for all μ± ∈ P(Ω) and it induces a distance on the set P(Ω) which metrizes the topology of weak convergence of measures. Actually,dα has a stronger property which is a comparison with the Wasserstein distance.

Proposition 2.2. Letμ+andμ be two probability measures onΩ. We denote byWp the Wasserstein distance associated to the cost (x, y)→ |x−y|p for p≥1. Then, one has

W1+, μ)≤dα+, μ)≤C W1+, μ)1−d(1−α), for a constant C >0 only depending ond,αand the diameter of Ω.

We refer to [17] for a proof of this property (see also [3,9] for an alternative proof) and [21,22] for the definition and main properties of the Wasserstein distance. In the same way, we definedαε as follows:

dαε(f+, f) = inf

Mεα(u) : u∈H1(Rd) such that ∇ ·u=f+−f

, (2.7)

wheref+, f∈L2+(Ω) satisfy Ωf+= Ωf. Althoughdαis a distance, it is not the case fordαε which does not satisfy the triangular inequality. Actually, because of the second term involving|∇u|2,Mεα is not subadditive.

However, foru1, . . . , un inMdiv(Ω), the inequality |∇u1+. . .+∇un|2≤n{|∇u1|2+. . .+|∇un|2}implies Mεα

n

i=1

ui

≤n n

i=1

Mεα(ui).

In particular, dαε is a pseudo-distance in the sense that the three properties in the following proposition are satisfied.

Proposition 2.3. Let f+,f andf1,. . . , fn beL2 densities,i.e.L2 non negative functions whose integral is equal to 1. Then one has

1. dαε(f+, f) = 0implies f+=f, 2. dαε(f+, f) =dαε(f, f+), 3. dαε(f0, fn)≤n

dαε(f0, f1) +dαε(f1, f2) +· · ·+dαε(fn−1, fn) .

(6)

3. Local estimate

We remind our assumption (2.3) which insures thatdα+, μ) is always finite. Our goal is to prove thatdαε enjoys a property similar to the following one.

Proposition 3.1. Let Q0 = (0, L)d Rd be a cube of side length L > 0. There exists some constant C > 0 only depending on dandαsuch that for all non negative Borel finite measureμof total mass θ >0,

dα(μ, θδ0)≤C θαL, whereδ0 is the Dirac measure at the point cQ0, the center ofQ0.

Since dαε(f+, f) is only defined onL2 functions f±, to do so, we first have to replaceθδ0 by some kernel which concentrates at the origin whenε goes to 0. Letρ∈ Cc1(B,R+) be a radial non negative function such that Rdρ= 1, whereB⊂Rd is the unit ball centered at the origin, and defineρθ,ε:=ρR as in (1.1), where

R=:Rθ,ε=εγθ1−γd−1.

Here, we recall that R and γ = dγ+12 were introduced in (1.4). Let Q be a cube inRd centered at some point cQ Rd and f L2+(Q) be a density such that Qf =: θQ. Then, we will denote by ρQ the kernel θρθ,ε

refocused atcQ with a small abuse of notation (indeed,ρQ also depends onf):

ρQ(x) =θQρθQ(x−cQ).

The main result of this section is the following theorem

Theorem 3.2 (Local estimate). Let us set Q0 = (0, L)d for some L > 0. There exists C >0 only depending onα,ρanddsuch that for all f ∈L2+(Q0)with Q

0f =θ, we have

IfsuppρQ0⊂Q0 then, there existsu∈H01(Q0)such that∇ ·u=f−ρQ0 and dαε(f, ρQ0)≤Mεα(u)≤C

θαL+εγ2f2L2

and uL1 ≤C L θ.

Otherwise, there existsu∈H01(Q0)such that

dαε(f, ρQ0)≤Mεα(u)≤Cεγ2f2L2 and uL1 ≤C L θ, whereQ0= 2 suppρQ0 :=B(cQ0,2Rθ,ε).

Remark 3.3. The Dirichlet term, εγ2f2L2, in the estimates above is easily understandable. Indeed, if ε is very large so that one can get rid of the first term in the energyMεα, then, one can use a classical Dirichlet type estimate, that is Theorem3.4below. On the contrary, forεvery small, the Γ-limit result onMεαtells us that these energies are close toMα so that it is natural to hope a similar estimate as the one forMα: that is to say an estimate from above byθαL(see [3]).

The main difficulty to prove Theorem3.2 is the non subadditivity of the pseudo-distances λαε. Indeed, our proof is based on a dyadic construction used by Xia in [23] to prove Proposition3.1(see also [3]). This gives a singular vector measureuwhich is concentrated on a graph. SinceMεα contains a term involving theL2 norm of∇u, we have to regularizeuby taking a convolution with the kernelρθ,εon each branch of the graph (θbeing the mass traveling on it). Unfortunately in this way, two different branches are no longer disjoints.

It is useful to see that we have a first candidate for the minimization problem (2.7). This candidate is of the form u=∇φ, where φis the solution of the Dirichlet problem

Δφ=f+−fin Q,

φ = 0 on ∂Q. (3.1)

(7)

Then u = ∇φ satisfies ∇ ·u = f+−f in Q and u(x) Rn a.e. on ∂Q where n stands for the external unit normal vector to∂Q. Alternatively, one could consider Neumann homogeneous boundary conditions forφ rather than Dirichlet boundary conditions. Then, one would obtainu(x)·n= 0 a.e. on∂Q. Theorem3.4below gives a better result in the sense that the candidateuvanishes at the boundary:

Theorem 3.4. Let Q0= (0, L)d be a cube of side length L >0. There exists C >0 only depending on dsuch that for allf ∈L20(Q0), there exists u∈H01(Q0,R2)solving ∇ ·u=f and satisfyinguL1(Q0)≤CLfL1(Q0)

together with

uH10(Q0):=

Q0

|∇u|2 1/2

≤CfL2(Q0),

whereL20(Q0) =

f ∈L2(Q0) : Q

0f(x) dx= 0

.

For a proof of this theorem, see, for instance, Theorem 2 in [7]: the only difference with Theorem3.4is that we add the estimate uL1(Q0) ≤CLfL1(Q0) which can be easily obtained following the proof of Bourgain and Brezis. The corresponding property formulated on a Lipschitz bounded connected domain Ω is also true (see Thm. 2’ in [7]) except that the constantCcould depend onΩin this case.

Of course, this candidate is usually not optimal for (2.7) and this does not allow for a good estimate because of the first term in the definition ofMεα. For this reason, we have to use the dyadic construction of Xia up to a certain level (“diffusion level”) from which we simply use Theorem3.4.

3.1. Dyadic decomposition of Q

0

and “diffusion level” associated to f

Let us call “dyadic descent” ofQ0= (0, L)d the setQ=

j≥0Qj, whereQj is thejth “dyadic generation”:

Qj=

(x1, . . . , xd) + 2jQ0 : xi ∈ {k2jL : 0≤k≤2j1} for i= 1, . . . , d . Note that Card(Qj) = 2jd. For eachQ∈ Q, let us define

• D(Q): the descent ofQ, the family of all dyadic cubes contained inQ.

• A(Q): the ancestry ofQ, the family of all dyadic cubes containingQ.

• C(Q): the family of children ofQcomposed of the 2d biggest dyadic cubes strictly included inQ.

F(Q): the father ofQ, the smallest dyadic cube strictly containingQ.

We now remind the dyadic construction described in [23] which irrigatesf from a point source. We first introduce some notations: fix a functionf∈L2+(Q0) with integralθand letQ∈ Qbe a dyadic cube centered atcQRd. Then we introduceθQ the mass associated to the cubeQas

θQ =

Q

f.

IfθQ= 0, we also define the kernel associated toQthrough

ρQ(x) =ρR(x), (3.2)

whereρR is defined in (1.1) for

R=RQ :=εγθ

1−γd−1

Q , γ= γ2 d+ 1· Here γwas defined in Define also the weighted recentered kernel by

ρQ(x) =θQρQ(x−cQ) (3.3)

(8)

ifθQ= 0 andρQ(x) = 0 otherwise. Lastly, we introduce the point source associated to the cubeQas SQ:=θQ×Dirac measure at pointcQ.

We are now able to construct a vector measureX such thatMα(X)<+∞. First define the measuresXQ as below:

XQ=

Q∈C(Q)

θQ nQ H1|[cQ,cQ], (3.4)

wherenQ = cQ −cQ

cQ−cQ. Then, we have

∇ ·XQ=

Q∈C(Q)

SQ− SQ

and the energy estimate

Mα(XQ)2d−2θαQdiam(Q), where diam(Q) stands for the diameter ofQ. Finally, the measureX =

Q∈QXQsolves∇ ·X=f− SQ0 and satisfies

Mα(X)≤Cθαdiam(Q0).

Indeed, it is enough to apply the following lemma withλ=α:

Lemma 3.5. Let Q∈ Q andλ∈]11/d,1]. There exists a constantC=C(λ, d)such that

Q∈D(Q)

θQλdiam(Q)≤CθλQdiam(Q).

Proof. Let j0 0 be such that Q ∈ Qj0. The definition of D(Q), the Jensen inequality and the fact that d−1−λd <0 give

Q∈D(Q)

θλQdiam(Q) =

j≥0

2jdiam(Q)

Q∈D(Q)∩Qj0+j

θλQ

diam(Q)

j≥0

2j2jd

⎝2jd

Q∈D(Q)∩Qj0+j

θQ

λ

≤θQλdiam(Q)

j≥0

2j(d−1−λd)

≤CθλQdiam(Q).

Now, the idea is to replace each term in (3.4) by its convolution with the kernelρQ. Unfortunately, this will make appear extra divergence terms around each node. We have to modifyX so as to make this extra divergence vanish using, for instance, Theorem3.4. Furthermore, we cannot follow the construction for all generationsj 1, otherwise the “enlarged edges” (convolution of a dyadic edge and the kernelρθ,ε) may overlap. This is the reason why we introduce the following definition:

Definition 3.6(“Diffusion level”). For a cubeQ0andf ∈L2+(Q0) we define the setD(Q0, f) orD(f)⊂ Qas the maximal element for the inclusion in the set

Λ={D⊂ Q : ∀Q∈D, A(Q)∪ C(F(Q))⊂D and suppρQ⊂Q}.

(9)

IfΛ=∅, that is suppρQ0 Q0, we take the conventionD(f) =∅. For all x∈Q0, define also the “generation index” ofxassociated to f as

j(f, x) = max{j : ∃Q∈ D(f)∩ Qj, x∈Q} ∈N∪ {±∞}, where the convention max() =−∞has been used.

In this way, each cube in D(f) contains the support of its associated kernel. Moreover, if Qis an element of D(f), then all its ancestry and its brothers (i.e. elements of the set C(F(Q))) are elements ofD(f). D(f) can be constructed by induction as follows: first takej= 0 andD(f) =∅. If suppρQ0⊂Q0 then addQ0 to the set D(f) andj is replaced by j+ 1. For all cubes Qin Λ∩ Qj−1: if all cubes Q ∈ C(Q)⊂ Qj are such that their associated kernels are supported onQthenD(f) is replaced byD(f)∪ C(Q). IfD(f) has been changed at this stagej is replaced byj+ 1 and the preceding step is reiterated. This process is repeated forj≥1 until it fails.

Let Dmin(f) be the set of all cubes in D(f) which are minimal for the inclusion. If Dmin(f) = , we also define

D(f) =

Q∈Dmin(f)

Q.

Note that this is actually a disjoint union: two distinct cubes in Dmin(f) are disjoint. Indeed, for Q, Q Dmin(f)⊂ Q, either Q∩Q= orQandQ are comparable:Q⊂Q or Q ⊂Q. In the last case, sinceQand Q are minimal, we deduce thatQ=Q.

Moreover, it is not difficult to see that, if Dmin(f)=∅, then D(f) = {x∈Q0 : j(f, x) is finite} and also that f(x) = 0 wheneverj(f, x) = +∞, where f is the precise representative off (i.e. the limit of the mean values of f on small cubes). Indeed, assume thatQ ∈ D(f) is a cube of side length LQ. Then, by definition, suppρQ⊂Qand for some constantCdepending onρand forν= 1−d−1γ, one hasεγθQν ≤CLQ and so

Q

f :=LQdθQ≤εγ/νL1Qd.

Since 1/ν−d= (d−1)(dαd+1d+1) is positive, we deduce thatLQ cannot be arbitrarily small if there exists x∈Q such thatf(x)>0. Moreover, iff(x)≥ηa.e. for someη >0, then there exists some constantCη >0 depending onη,ε,dandαsuch that

∀Q∈ D(f), LQ≥Cη. (3.5) In particular, one can deduce thatDmin(f) = if and only ifD(f) = or f(x) = 0 a.e. Indeed, ifD(f) =∅, then it is clear that Dmin(f) =. Conversely, assume thatDmin(f)=(i.e.Q0∈ D(f)) and that there exists x Q0 such that f(x) > 0. Since

Q∈D(f)∂Q is negligible for the Lebesgue measure, one can assume that x

Q∈D(f)Q. Then 0 j(f, x) < + and so there exists a minimal cube Q ∈ D(f) containing x. Then Q ∈ Dmin(f). Indeed, ifQ ∈ D(f) and Q Q, then A(Q) ⊂ D(f) and there exists Q ∈ A(Q) such that QQandx∈Q which is a contradiction.

We are now able to define two approximations off which are useful for our problem. The first is a dyadic approximation off by an atomic measure,

Λεf = Q∈Dmin(f)SQ ifDmin(f)=∅,

SQ0 otherwise,

where we recall the definition ofSQ :=θQδcQ. We also define an approximation inH1(Q0),

λεf =

⎧⎨

Q∈Dmin(f)

ρQ ifDmin(f)=∅, ρQ0 otherwise,

(10)

where ρQ is defined in (3.3). The following result shows in which sense λεf is an approximation of f and justifies the term “diffusion level”. Indeed, this proposition indicates that we get a good estimate by using a local diffusion fromλεf tof, i.e. minimizing Q|∇u|2 over the constraint∇ ·u=λεf−f for allQ∈ Dmin(f) (see Thm.3.4).

Proposition 3.7. There exists a constant C >0 depending on dandρsuch that for all f ∈L2+(Q0), dαεεf, f) +dαεf, f)≤C εγ2f2L2(Q0).

More precisely, ifsuppρQ0⊂Q0, there existsu∈H01(Q0)such that ∇ ·u=f−λεf as well as Mεα(u)≤C εγ2f2L2 and uL1 ≤C diam(Q0)fL1.

If suppρQ0 Q0 the same estimates hold but the condition u∈ H01(Q0) has to be replaced by u H01(Q0), whereQ0 is a cube containingQ0 andsuppρQ0.

Proof. First assume that suppρQ0 ⊂Q0 i.e.Q0∈ D(f). IfDmin(f) =∅, thenf(x) = 0 a.e. and the proposition is trivial. Hence, one can assume that Dmin(f)=∅. Thenf is supported onD(f) andDmin(f) =:{Qi}iI is a finite or countable partition ofD(f). Denote for simplicityDi:= diam(Qi),fi:=f1Qi (restriction off toQi), θi:=θQi andρi:=ρQi =θiρRi fori∈I, where

Ri:=RQi =εγθ

1−γd−1

i .

SinceQi is minimal inD(f), we deduce that, for some constantsC, C >0,

CRi≤Di≤CRi. (3.6)

Indeed, the first inequality follows from the fact that suppρi⊂Qi and diam(suppρi) =cRi for some constant c depending on ρ. For the second inequality observe that, since Qi is minimal, there exists Q ∈ C(Qi) such that suppρQ Q and hence RQ cdiam(Q) = c/2Di for some constant c > 0 depending on ρ. Since θQ≤θQi=θi, one hasRQ≤Ri and the second inequality follows.

Now, Theorem 3.4 allows us to find ui H01(Qi) such that ∇ ·ui = gi, uiH1(Qi) CgiL2(Qi) and uiL1(Qi)≤C DigiL1(Qi), wheregi:=fi−ρi. Sinceui vanishes at∂Qi, one can extenduiby 0 out ofQi to get a function inH1(Rd): for the sake of simplicity, this function is still denoted byui. Consequently,u=

iui

belongs toH01(Q0) and∇ ·u=f −λεf. It remains to estimateMεα(u) anduL1(Q0). First of all, uL1(Q0)

i

uiL1(Qi)≤Cdiam(Q0)

i

giL1(Qi)

and the inequalitygiL1(Qi)i leads touL12C diam(Q0)fL1 as required.

Let us compute theL2-norm ofρi:

ρi2L2(Qi)=θi2ρRi2L2(Qi)=θ2iRidρ2L2(Qi)=2iRi d. By a Cauchy–Schwarz inequality,

θi2=

Qi

fi

2

≤ |Qi|fi2L2(Qi)=Ddifi2L2(Qi) (3.7) which, together with (3.6), gives

ρi2L2(Qi)≤C Rdifi2L2(Qi)Ri d=Cfi2L2(Qi).

Références

Documents relatifs

From the preceding Theorem, it appears clearly that explicit expressions of the p-optimal coupling will be equivalent to explicit expressions of the function f

Contrary to what happens in the Monge-Kantorovich model, in the setting of branched transportation, an optimal structure cannot be described only using a transport plan, giving

Among our results, we introduce a PDE of Monge–Kantorovich type with a double obstacle to characterize optimal active submeasures, Kantorovich potential and optimal flow for the

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

In particular, a topological space is arcwise connected (resp. locally arcwise connected), if and only if every couple of points in is connected by some path [a , b] ⊂ (resp.

In two spatial dimensions we propose a phase field approximation of this street network using N phase fields and a function approximating the mass flux through the network.. We

We present a general method, based on conjugate duality, for solving a con- vex minimization problem without assuming unnecessary topological restrictions on the constraint set..

As regards the most basic issue, namely (DG) pertaining to the question whether duality makes sense at all, this is analyzed in detail — building on a lot of previous literature —