• Aucun résultat trouvé

Anisotropic Minkowski Content and Application to minimizers of free discontinuity problems

N/A
N/A
Protected

Academic year: 2021

Partager "Anisotropic Minkowski Content and Application to minimizers of free discontinuity problems"

Copied!
19
0
0

Texte intégral

(1)

HAL Id: hal-01222602

https://hal.archives-ouvertes.fr/hal-01222602

Preprint submitted on 4 Nov 2015

HAL

is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire

HAL, est

destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

Anisotropic Minkowski Content and Application to minimizers of free discontinuity problems

David Vicente

To cite this version:

David Vicente. Anisotropic Minkowski Content and Application to minimizers of free discontinuity

problems. 2015. �hal-01222602�

(2)

Anisotropic Minkowski Content and Application to minimizers of free discontinuity problems

David Vicente October 30, 2015

Abstract

This paper deals with a generalization of a result of Geometric Measure Theory: if a rectifiable set is closed, then its Minkowski content is equal to its Hausdorff measure. This result is generalized to an anisotropic and continuous perturbation of the Hausdorff measure. So, an adapted Minkowski content is introduced and the proof that it is equal to the perturbed Hausdorff measure under the same hypothesis of closure and rectifiability is given. This result is applied to the jump set of a minimizer of a free boundary problem for the case where the Hausdorff measure of the boundary is perturbed by an anisotropic and continuous function and a rigorous setting is given for the approximation of an anisotropic version for the Mumford-Shah functional.

Introduction

ForS⊂Rn, the (n−1)-dimensional upper and lower Minkowski contents ofS are defined by M?(S) = lim sup

ρ→0+

Ln({x: dist(x, S)< ρ})

2ρ , M?(S) = lim inf

ρ→0+

Ln({x: dist(x, S)< ρ})

2ρ ,

whereLn is the Lebesgue measure. IfM?(S) =M?(S), their common value, denoted byM(S), is called the (n−1)-dimensional Minkowski content. In [1] (Theorem 3.2.39), the following result is given.

Theorem 0.1 (Federer). If S is a closed and (n−1)-rectifiable subset ofRn, then M(S) =Hn−1(S),

whereHn−1 is the(n−1)-dimensional Hausdorff measure.

In this article, we focus on a class of anisotropic perturbation of Hn−1 defined on (n−1)- rectifiable surfacesS⊂Rn by

Z

S

hMν, νi1/2dHn−1, (0.1) whereM:Rn→S+n(R) is a continuous field of symmetric definite positive matrices,νis an unitary vector orthogonal to the tangent plane ofS. AsS is rectifiable, thenνis well definedHn−1-almost everywhere inS. In order to get an approximation of (0.1) which involves the Lebesgue measure, as in Theorem0.1, for any (x,v)∈Rn×Rn, we consider the following Riemannian metric:

φ(x,v) =hM−1(x)v,vi1/2 and for any (x, y)∈(Rn)2its associated integrated distance:

distφ(x, y) = inf Z 1

0

φ

γ,dγ dt

dt: γW1,1([0; 1];Rn), γ(0) =x, γ(1) =y

, distφ(x, S) = inf{distφ(x, y):yS}.

(0.2) Then, the associated anisotropic Minkowski (n−1)-dimensional upper and lower contents are defined by the limits

M?M(S) = lim sup

ρ→0+

Ln({x: distφ(x, S)< ρ})

2ρ , M?M(S) = lim inf

ρ→0+

Ln({x: distφ(x, S)< ρ})

2ρ .

(3)

IfM?M(S) =M?M(S), we call their common value the (n−1)-dimensional anisotropic Minkowski contentMM(S). In this paper, we prove the following

Theorem 0.2. If S is a closed and (n−1)-rectifiable subset of Rn and M : RnS+n(R) is continuous, then we have

MM(S) =Z

S

hMν, νi1/2dHn−1,

whereν is an unitary vector, normal toS.

A comparable result is given in [2] (Theorem 6.1) forS=∂EandEis a set of finite perimeter.

In [3], the study focuses on anisotropic outer Minkowski content for the same class of sets. In our case,S is not necessary the boundary of a set. This generalization is of interest for the case where S is the jump setJu of a functionuwith bounded variation as we will see in last section.

Our result allows us to extend a regularity result for free boundary problems to a larger class of energies which involve a continuous perturbation of the Hausdorff measure. More precisely, as it has been introduced in [4], u ∈ SBV(Ω) is an almost quasi-minimizer of a free discontinuity problem if there exist σ ≥ 1, α > 0 and cα > 0 such that: if v ∈ SBV(Ω), B(x, r) ⊂ Ω and [w6=v]⊂B(x, r), then we have

Z

B(x,r)

|∇w|2dx+Hn−1(JwB(x, r))≤ Z

B(x,r)

|∇v|2dx+σHn−1(JvB(x, r)) +cαrn−1+α, whereJu is the jump set ofu. The difference of this definition with respect to the classical notion ofquasi-minimizer consists in the coefficientσ≥1 which allows bounded perturbation ofHn−1. We prove the following

Theorem 0.3. LetM:RnS+n(R)be continuous anduSBV(Ω)be an almost quasi-minimizer of a free discontinuity problem, then we have

MM(Ju) =Z

Ju

hMνu, νui1/2dHn−1.

This result is applied to the Mumford-Shah model in an anisotropic setting. Indeed, a minimizer this model is an almost quasi-minimizer of a free boundary problem and the approximation result of Theorem 3.1is satisfied. In a joint paper [5], this is the key tool to perform a Γ-convergence approximation of our model.

In section1, we recall some classical results of Geometric Measure Theory. Section2is devoted to the proof of Theorem 0.2. In section 3, we give the proof of Theorem 0.3 and we apply this result to give a rigorous setting for the approximation of an anisotropic Mumford-Shah model.

1 Geometric Measure Theory Framework

We adopt the notations:

• Ω for an open and bounded subset ofRn,

• hv1,v2i ∈Rfor the canonical scalar product ofv1,v2∈Rn,

• Vn−1

i=1vi∈Rn for the canonical vectorial product ofv1, . . . ,vn−1∈Rn,

• |v|for the euclidean norm of v∈Rn,

B(x, r) for the open ball inRn with radiusrand centerx,

• Mn(R) for the space of n×nreal matrices,

• kMkfor an induced norm of M∈Mn(R),

• S+n(R)⊂Mn(R) for the subset of symmetric definite positive matrices,

• GLn(R)⊂Mn(R) for the subset of invertible matrices,

• On(R)⊂GLn(R) for the subgroup of orthogonal matrices,

• B(Ω) the class of Borelian subsets of Ω,

• Ln for the Lebesgue measure inRn,

• Hk for thek-dimensional Hausdorff measure.

(4)

The geometric measure theory framework is mainly extracted from [1] and [2].

Definition 1.1 (Federer, 3.2.1). Let f maps a subset ofRn−1 ontoRn, the (n−1)-dimensional Jacobian is defined by

Jn−1(f)(a) =

^n−1

i=1

∂f

∂xi(a) wheneverf is differentiable at a.

Theorem 1.1 (Area Formula, Federer, 3.2.3). Supposef :Rn−1→Rn is Lipshitzian. If uis an Ln−1-integrable function, then

Z

Rn−1

u(x)Jn−1(f)(x)dLn−1(x) =Z

Rn

X

x∈f−1{y}

u(x)dHn−1(y). We are interested in the class of subset of Rn which are (n−1)-rectifiable.

Definition 1.2(Federer, 3.2.14). The setE⊂Rnis(n−1)-rectifiable if there exists a Lipshitzian function f :Rn−1→Rn mapping some bounded subset ofRn ontoE.

The following result is useful to get an univalent parametrization of a rectifiable set.

Theorem 1.2 (Federer, 3.2.4). For everyLn−1 measurable set A, there exists a Borel set BA∩ {x:Jn−1(f)(x)>0}

such that f B is univalent andHn−1(f(A)\f(B)) = 0.

We say thatϕ: Ω×Rn →R+is a Finsler metric ifϕis continuous and if there existsλ,Λ>0 such that, for any (x,v, t)∈Ω×Rn×R, it satisfies

φ(x, tv) =|t|φ(x,v), λ|v| ≤φ(x,v)≤Λ|v|.

We define the integrated distance associated toϕas

∀(x, y)∈(Rn)2, distϕ(x, y) = infZ 1 0

ϕ

γ,dγ dt

dt: γW1,1([0; 1];Rn), γ(0) =x, γ(1) =y

.

For anyE⊂Rn, we define the associated diameter as

diamϕ(E) = sup{distϕ(x, y):x, yE}

and the associatedk-dimensional Hausdorff measure by Hkϕ(E) = lim

δ→0+inf (ωk

2k X

i∈I

(diamϕ(Ai))k:E⊂[

i∈I

Ai,∀i∈I,diamϕ(Ai)≤δ )

whereωk is the volume of the unit ball inRk.

Theorem 1.3 (Bellettini). Let ϕbe a Finsler metric, then, for all Borel set E⊂Rn, we have Hnϕ(E) =Z

E

Hn(B(0,1)) Hn(Bϕ(x,1))dx, whereBϕ(x,1) is the unit ball centered atxfor the metricϕ.

2 Proof of Theorem 0.2

The proof is divided in two steps. First, we assume in section2.1that the metric does not depend onx. Then, in section2.2, we remove this assumption and we prove the Theorem0.2in the general setting.

(5)

2.1 Homogeneous Case

In this section, we assume that the metric is homogeneous. So, we denote byM0its common value and, for anyv∈Rn, we set

φ0(v) =hM−10 v,vi1/2, (2.1)

so that (Rn, φ0) is an Euclidean space. We prove the following

Theorem 2.1. If S is a closed and (n−1)-rectifiable subset of Rn and M0S+n(R). Then we have

MM0(S) =Z

S

hM0ν, νi1/2dHn−1,

whereν is an unitary and normal vector toS.

We decompose the proof in three Lemmas. AsM−10 is symmetric positive definite,Rnmay be viewed as an euclidean space according to the scalar producthM−10 ·,·i. So, Theorem0.1may be directly applied. The three following Lemmas consist in the change of variable fromRn endowed with the scalar producthM−10 ·,·ito Rn endowed with the canonical scalar product.

Lemma 2.1. ForM0S+n(R)0 defined by (2.1) and for anyE∈ B(Rn), we have Ln(E) =p

det(M0)Hnφ0(E).

Proof. AsM0is a symmetric positive definite matrix, there existsP ∈On(R) and (λi)i=1...nsuch thatλi>0 for anyi∈ {1, . . . , n} and

M0=P

λ1 0 · · · 0 0 λ2 ... ...

... ... ... 0 0 · · · 0 λn

P−1. (2.2)

We denote by (ci)i=1...n the canonical basis of Rn, then (√

λiP ci)i=1...n is an orthonormal basis for the scalar producthM−10 ·,·i. So, the linear applicationL:Rn→Rn characterized byL(ci) =

λiP ci for any i ∈ {1, . . . , n}, is an isomorphism which satisfies L(B(0,1)) = Bφ0(0,1) and det(L) =p

det(M0). It gives

Ln(Bφ0(0,1)) = Z

B(0,1)

|det(L)|dx,

= p

det(M0)Ln(B(0,1))

According to isodiametric inequality (see 2.10.35 in [1]), we have Hn = Ln and then, for any E∈ B(Rn), Theorem1.3yields

Hnφ

0(E) = Ln(E) pdet(M0). which concludes the proof of Lemma2.1.

Setting ei = √

λiP ci, then (ei)i=1,...,n is an orthonormal basis of Rn for the scalar product hM−10 ·,·i. In the Euclidean space Rn, φ0), the vectorial product of (vi)i=1,...,n−1, denoted by Vn−1

φ0,i=1vi, is characterized by the following equality

hM−10 v1, e1i . . . hM−10 vn−1, e1i hM−10 w, e1i

... ... ...

... ... ...

hM−10 v1, eni . . . hM−10 vn−1, eni hM−10 w, eni

=

M−10

^n−1

φ0,i=1vi

,w

, (2.3)

which must be true for anyw∈Rn.

Lemma 2.2. If M0S+n(R) andφ0 defined by (2.1), then we have

^n−1

φ0,i=1vi=M0(n−1i=1vi) pdet(M0) .

(6)

Proof. AsP ∈On(R), then (P ci)ni=1 is an orthonormal basis for the usual scalar product and, for anyw∈Rn, we have

hv1, P c1i . . . hvn−1, P c1i hw, P c1i

... ... ...

... ... ...

hv1, P cni . . . hvn−1, P cni hw, P cni

=

n−1i=1vi,w .

According to (2.2), for anyv∈Rn, we have hv, P cii=p

λihM−10 v,p λiP cii and then

pλ1. . . λn

hM−10 v1, e1i . . . hM−10 vn−1, e1i hM−10 w, e1i

... ... ...

... ... ...

hM−10 v1, eni . . . hM−10 vn−1, eni hM−10 w, eni

=

M−10 M0n−1i=1vi

,w .

According to (2.3), it concludes the proof of Lemma2.2.

As for the vectorial product, we may define an anisotropic (n−1)-dimensional Jacobian by Jφn−10 (f)(a) =

^n−1

φ0,i=1

∂f

∂xi(a)

wheneverf is differentiable ata. The following result is a straightforward consequence of Lemma 2.2.

Lemma 2.3. Let M0S+n(R) be fixed andφ0 defined by (2.1), then we have Jφn−10 (f)(a) = 1

pdet(M0)

M0

^n−1

i=1

∂f

∂xi(a) ,^n−1

i=1

∂f

∂xi(a)1/2 .

Now we give the proof of Theorem2.1.

Proof. LetS be a (n−1)-rectifiable closed set. AsL, introduced in the proof of Lemma2.1, is an automorphism, thenSis also (n−1)-rectifiable and closed forRnendowed with the scalar product hM−10 ·,·i. According to Lemma2.1, we have

Ln({x: distφ0(x, S)< ρ})

2ρ =p

det(M0)Hnφ0({x: distφ0(x, S)< ρ})

2ρ . (2.4)

We may apply Theorem0.1in the euclidean space (Rn,hM−10 ·,·i), it gives

ρ→0lim+ Hnφ

0({x: distφ0(x, S)< ρ})

2ρ =Hn−1φ

0 (S). (2.5)

As S is rectifiable, there exists a Lipschitzian function f : Rn−1 → Rn and a bounded subset A⊂Rn−1such thatSf(A). According to Corollary1.2, there exists a Borel set

BA∩ {x:Jφn−10 (f)(x)>0}

such thatf B is univalent andHn−1(f(A)\f(B)) = 0. We denote byC=f−1(S)∩B and then Hn−1φ

0 (f(C)) =Hn−1φ

0 (S). Asf Cis univalent, according to Area formula1.1withu=1C, we have Hn−1φ

0 (S) =Z

C

Jφn−10 (f)(x)dx. (2.6)

(7)

For anyxC we haveJφn−10 (f)(x)6= 0 which implies that ∂x∂f1(x), . . . ,∂x∂f

n−1(x) is free and then Jn−1(f)(x)6= 0. According to Lemma2.3, we may write

Jφn−10 (f)(x) = Jφn−10 (f)(x)

Jn−1(f)(x)Jn−1(f)(x),

= 1

pdet(M0)

* M0

 Vn−1

i=1

∂f

∂xi(x)

Vn−1 i=1

∂f

∂xi(x)

, Vn−1

i=1

∂f

∂xi(x)

Vn−1 i=1

∂f

∂xi(x)

+1/2

Jn−1(f)(x).

We set

u(x) =

* M0

 Vn−1

i=1

∂f

∂xi(x)

Vn−1 i=1

∂f

∂xi(x)

, Vn−1

i=1

∂f

∂xi(x)

Vn−1 i=1

∂f

∂xi(x)

+1/2 .

According to Lemma 3.2.25 in [1], we haveu(x) =hM0ν(f(x)), ν(f(x))i1/2 where ν(f(x)) is an unitary and orthogonal vector tof(C) atf(x). Applying Area formula1.1, this time inRnendowed with its canonical euclidean structure, gives

Z

C

Jφn−10 (f)(x)dx = 1 pdet(M0)

Z

C

u(x)Jn−1(f)(x)dx,

= 1

pdet(M0) Z

f(C)

hM0ν, νi1/2dHn−1,

= 1

pdet(M0) Z

S

hM0ν, νi1/2dHn−1. According to (2.4), (2.5) and (2.6) we may conclude

ρ→0lim+

Ln({x: distφ0(x, S)< ρ})

2ρ =Z

S

hM0ν, νi1/2dHn−1.

2.2 Inhomogeneous setting

In this section, we remove the homogeneity assumption and, using a piecewise constant approxi- mation ofM, we prove Theorem0.2. The proof is divided in three Propositions.

Proposition 2.1. LetMbe as in Theorem0.2. Then, for any compact neighborhoodCofS, there existsλ,Λ>0, such that the following ellipticity condition is satisfied for any(x,v)∈C×Rn

λ|v|2≤ hM(x)v,vi ≤Λ|v|2.

Proof. As M(x) is symmetric positive definite for any xC, there exists λ(x),Λ(x) > 0 such that, for anyv∈Rn, we have

λ(x)|v|2≤ hM(x)v,vi ≤Λ(x)|v|2.

As M is continuous, the previous inequalities remain true in any compact C with (λ,Λ) only depending onC.

Proposition 2.2. Let S andM be as in Theorem 0.2. Then, we have M?M(S)≤

Z

S

hMν, νi1/2dHn−1. Proof. We may assume thatR

ShMν, νi1/2dHn−1 is finite, otherwise the result is ensured. LetC be a compact neighborhood ofS. Letδ >0 be fixed. AsMis continuous, there existsη >0 such that

|x−y| ≤η ⇒ kM(x)−M(y)k ≤δ (2.7)

(8)

for any (x, y)∈C2. According to Proposition 2.1, there existsλ >0 such that λ1/2Hn−1(S)≤

Z

S

hMν, νi1/2dHn−1

and thenHn−1(S) is also finite. For t ∈R andi ∈ {1, . . . , n}, we set Πit ={x∈C:hx, eii=t}. Thus, fork ∈Nfixed,

t∈R:Hn−1(S∩Πit)> 1k is finite and then, for anyi ∈ {1, . . . , n}, the set

t∈R:Hn−1(S∩Πit)>0 is at most countable. So, if we consider the cubes with its faces orthogonal to the vectors of the orthogonal basis, there exists a partition K of C by cubes with diameter less thanη, such that for anyK∈ K, we have

Hn−1(S∂K) = 0. (2.8)

For anyK∈ K, we fixaKKand, for any (x,v)∈K×Rn, we set (

fM(x) =M(aK),

φe(x,v) =h(M(aK))−1v,vi. (2.9) ForK∈ Kandr >0, we setKr={x: dist(x, ∂K)≤r}. We have the following decomposition

M?M(S)≤ X

K∈K

M?M(SKr) + X

K∈K

M?M(S∩(K\Kr)). (2.10) InClaim 1 andClaim 2, we will determine an upper bound for the two previous sums.

Claim 1: We have X

K∈K

M?M(SKr)≤λn2Λ1/2 X

K∈K

Hn−1(SKr).

According to Proposition2.1, there existsλ >0 and Λ>0 such that, for any (x,v)∈C×Rn, we have

Λ12|v| ≤ hM(x)−1v,vi1/2λ12|v|. (2.11) Let (x, y)∈C2 andγ∈W1,1([0; 1];Rn) such thatγ(0) =xandγ(1) =y, then we have

Λ12 Z 1 0

|γ|˙ dt≤ Z 1

0

φ(γ,γ˙)dtλ12 Z 1

0

|γ|˙ dt.

According to (0.2) and (2.11), for any (x, y)∈C2, it gives

Λ12dist(x, y)≤distφ(x, y)≤λ12dist(x, y) (2.12) and then, for anyxC, the following inclusions are satisfied

B(x, λ1/2)⊂Bφ(x,1)⊂B(x,Λ1/2). For anyxC, , asHn(B(x, t)) =tnHn(B(0,1)), we get

Λn2 ≤ Hn(B(0,1))

Hn(Bφ(x,1)) ≤λn2 and with Theorem1.3, for any E∈ B(C), it yields

Λn2Hn(E)≤ Hnφ(E)≤λn2Hn(E).

Moreover, (2.12) implies {x: distφ(x, SKr)< ρ} ⊂ {x: dist(x, SKr) <Λ1/2ρ} and then we have

M?M(SKr) = lim sup

ρ→0+

Hnφ({x: distφ(x, SKr)< ρ})

2ρ ,

λn2 lim sup

ρ→0+

Hn({x: dist(x, SKr)<Λ1/2ρ})

2ρ ,

λn2Λ1/2M?(SKr).

(9)

AsSis closed and (n−1)-rectifiable, so isS∩Kr. We may apply Theorem0.1to getM?(S∩Kr) = Hn−1(SKr) andClaim 1 is proved.

Claim 2: We have

X

K∈K

M?M(S∩(K\Kr))≤(1 +θ0δ)Z

S

hMν, νi1/2dHn−1,

whereθ0 is a constant which depends only on the restriction of M inC.

Let C be a compact neighborhood of S. As M is continuous, then there exists a constant m >0 which depends only onCsuch that, for any (x, y)∈C2, we have

kM−1(x)−M−1(y)k ≤mkM(x)−M(y)k.

Let (x, y)∈ C2 and γ ∈W1,1([0; 1];C) such that γ(0) =x andγ(1) =y, according to (2.7), we have the inequalities

Z 1 0

hM−1(γ) ˙γ,γi˙ 1/2dt− Z 1

0

hfM−1(γ) ˙γ,γi˙ 1/2dt

≤ Z 1

0

|h(M−1(γ)−Mf−1(γ)) ˙γ,γi|˙ hM−1(γ) ˙γ,γi˙ 1/2+hM−1(γ) ˙γ,γi˙ 1/2dt,

≤ Z 1

0

kM−1(γ)−Mf−1(γ))khγ,˙ γi˙ hM−1(γ) ˙γ,γi˙ 1/2+hM−1(γ) ˙γ,γi˙ 1/2dt,

−1

Z 1 0

hγ,˙ γi˙ 1/2dt,

32

Z 1 0

hfM−1(γ) ˙γ,γi˙ 1/2dt.

We denote m

32 byθ0, it gives

Z 1 0

hM−1(γ) ˙γ,γi˙ 1/2dt

≤(1 +θ0δ)Z 1 0

hfM−1(γ) ˙γ,γi˙ 1/2dt, so that

{x: distφ(x, S∩(K\Kr))< ρ} ⊂n x: dist

eφ(x, S∩(K\Kr))<(1 +θ0δ)ρo

, (2.13)

and then

M?M(S∩(K\Kr))≤(1 +θ0δ)M?

Me(S∩(K\Kr)). (2.14) As φe is an homogeneous metric in a neighborhood of K\Kr as in Section 2.1, we may apply Theorem2.1to obtain

M?

Me(S∩(K\Kr)) =Z

S∩(K\Kr)

hMν, νi1/2dHn−1 and

M?

Me(S∩(K\Kr))≤ Z

S∩K

hMν, νi1/2dHn−1. According to (2.14), we get

X

K∈K

M?M(S∩(K\Kr))≤(1 +θ0δ)Z

S

hMν, νi1/2dHn−1, so it concludes the proof ofClaim 2.

ApplyingClaim 1 andClaim 2 to the decomposition (2.10), we have M?M(S)≤λn2Λ1/2 X

K∈K

Hn−1(SKr) + (1 +θ0δ)Z

S

hMν, νi1/2dHn−1 Taking the limit r→0+gives

(10)

M?M(S)≤λn2Λ1/2 X

K∈K

Hn−1(S∂K) + (1 +θ0δ)Z

S

hMν, νi1/2dHn−1. Applying (2.8), and then taking the limit asδ→0+concludes the proof of Proposition2.2

M?M(S)≤ Z

S

hMν, νi1/2dHn−1.

Proposition 2.3. Let S andM as in Theorem 0.2. Then, we have M?M(S)≥

Z

S

hMν, νi1/2dHn−1.

Proof. Let δ > 0, η > 0 and K as in (2.8) and (2.9). For r > 0 and K ∈ K, we still set Kr={x: dist(x, ∂K)≤r}. AsK is finite and

(K, L)∈ K2, K6=L ⇒ dist(Kr, Lr)>0, then we have

M?M(S)≥ X

K∈K

M?M(S∩(K\Kr)) (2.15)

With the same proof as for (2.13), we have {x: dist

φe(x, S∩(K\Kr))< ρ} ⊂ {x: distφ(x, S∩(K\Kr))<(1 +θ0δ)ρ}

and then

Ln({x: dist

φe(x, S∩(K\Kr))< ρ})

2ρ ≤Ln({x: distφ(x, S∩(K\Kr))<(1 +θ0δ)ρ})

2ρ .

Passing to the lim inf withρ→0+ gives M?

Me(S∩(K\Kr))≤(1 +θ0δ)M?M(S∩(K\Kr)).

As fM is constant in a neighborhood of K\Kr as in Section 2.1, we may apply Theorem2.1 to obtain

M?

Me(S∩(K\Kr)) =Z

S∩(K\Kr)

hfMν, νi1/2dHn−1. According to (2.15), we get

M?M(S) ≥ (1 +θ0δ)−1 X

K∈K

Z

S∩(K\Kr)

hfMν, νi1/2dHn−1,

≥ (1 +θ0δ)−1Z

S∩Cr

hfMν, νi1/2dHn−1, where we have setCr=S

K∈K(K\Kr). Remark that r1< r2Cr2Cr1, [

r>0

Cr=C\ [

K∈K

∂K

!

so, we have

r→0lim+ Z

S∩Cr

hfMν, νi1/2dHn−1=Z

S\∪∂K

hfMν, νi1/2dHn−1. AsHn−1(∂K) = 0 for anyK∈ K(2.8), then

r→0lim+ Z

S∩Cr

hfMν, νi1/2dHn−1=Z

S

hfMν, νi1/2dHn−1. We deduce that

M?M(S)≥(1 +θ0δ)−1Z

S

hfMν, νi1/2dHn−1 and passing to the limit forη→0+ and then forδ→0+concludes the proof.

(11)

3 Regularity result for almost quasi-minimizers of a free boundary problem

The standard functional framework is the theory of functions with bounded variation and in particular thespecial set of functions with bounded variations SBV that may be found in [6].

3.1 Definition and main result

In [7], it is proved that ifu∈SBV(Ω) is a minimizer of the well known Mumford-Shah functional E(u) =Z

(ug)2dx+Z

|∇u|2dx+Hn−1(Ju),

whereJu is the jump set ofu, then we haveHn−1(Ju) =M(Ju). We want to get the same result when Hn−1 is perturbed by an anisotropic and continuous function. So, for u ∈ SBV(Ω) and B(x, r)⊂Ω, we set

F(u, σ, B(x, r)) =Z

B(x,r)

|∇u|2dx+σHn−1(JuB(x, r)), In [4], the following definition is introduced.

Definition 3.1. For σ ≥1, α > 0 and cα > 0, we say that wSBV(Ω) is a (σ, α, cα)-almost quasi-minimizer of a free discontinuity problem, if there existsα >0andcα≥0such that for any vSBV(Ω), we have

[w6=v]⊂B(x, r), B(x, r)⊂Ω ⇒ F(w,1, B(x, r))≤F(v, σ, B(x, r)) +cαrn−1+α. The main result we introduce in this section is the following.

Theorem 3.1. Let M: Ω→S+n(R)be continuous anduSBV(Ω)be an almost quasi-minimizer of a free discontinuity problem (3.1), then we have

MM(Ju) =Z

Ju

hMνu, νui1/2dHn−1.

3.2 Proof of Theorem 3.1

To prove Thorem 3.1, we need the two following regularity results for the jump set of almost quasi-minimizers which are extracted from [4].

Theorem 3.2. Let ube an almost quasi-minimizer of a free discontinuity problem, then Hn−1 Ju\Ju

= 0.

Theorem 3.3. There exist constantsβ, ρ0 such that for every(σ, α, cα)almost quasi-minimizer u, for everyxJu and for every 0< ρ < ρ0 such that B(x, ρ)⊂Ω, we have

Hn−1(JuB(x, ρ))≥βρd−1. First, we prove the following

Lemma 3.1. Let ube an almost quasi-minimizer of a free discontinuity problem (3.1), then for any compact set KJu∩Ω, we have

M?M(K)≤ Z

Ju

hMνu, νui1/2dHn−1. Proof. We may assume thatR

JuhMνu, νui1/2dHn−1 is finite, otherwise the result is ensured. Ac- cording to Proposition2.1, we have

λ1/2Hn−1(Ju)≤ Z

Ju∩K

hMνu, νui1/2dHn−1

(12)

so,Hn−1(Ju) is finite too.

According to [8], Section 5.9, the set Ju is rectifiable up to a Hn−1-negligible set N. More precisely, there exists a countable family (Ki)i∈Nof compactC1-hypersurfaces such that

Ju=N ∪ [

i∈N

Ki

! ,

whereHn−1(N) = 0. LetKJu∩Ω be compact, we have the decomposition

K=

KJu\Ju

"

K

q

[

i=1

Ki

#

K∩

[

i=q+1

Ki

∪[K∩ N].

Theorem 3.2 gives Hn−1(KJu\Ju) = 0. Let δ > 0 be fixed. As Hn−1(Ju) is finite, there existsq∈Nsuch that

Hn−1

K∩

[

i=q+1

Ki

≤δ. (3.1)

In the sequel, we omit the dependance withδfor the sake of simplicity. We set S=K∩Sq i=1Ki

and forA⊂Rn we adopt the following notation

Aρ:={x: distφ(x, A)< ρ}.

Letτ >0 andρ >0 be fixed, we decompose K = (K\Sτ ρ)∪Sτ ρ and setE =K\Sτ ρ. So, we have

KρEρS(1+τ)ρ, and then

Ln(Kρ)

2ρ ≤Ln(Eρ)

2ρ +Ln(S(1+τ)ρ)

2ρ . (3.2)

The rest of the proof consists in computing an upper bound, whenρ→0+, for the two terms in the right hand side of (3.2). AsS is a closed and rectifiable set, Theorem0.2gives

ρ→0lim+

Ln(S(1+τ)ρ)

2ρ = (1 +τ)Z

S

hMνu, νui1/2dHn−1 and then

ρ→0lim+

Ln(S(1+τ)ρ)

2ρ = (1 +τ)Z

Ju

hMνu, νui1/2dHn−1. (3.3) In the fourth following Claims, we prove that lim supLn(Eρ) converges to 0 whenδ converges to 0. The main tool is the regularity result given by Theorem3.3.

Claim 1: There exists p∈N andx1, . . . , xpEJu such thatE⊂Sp

i=1Bφ(xi, τ ρ) and (i, j)∈ {1, . . . , p}2, i6=jdistφ(xi, xj)≥τ ρ. (3.4)

We construct (xi)i by an iterative way and we show that the number of iterations is finite.

If E = ∅, then the result is obvious. Otherwise, there exists ˜x1E. As EJu, there exists (yk)kJuconverging to ˜x1. As distφ( ˜x1, S)> τ ρ, there existsk0∈Nsuch that distφ(yk0, S)> τ ρ and we setx1=yk0.

Let us assume that there exitsx1, . . . , xpEJuwhich satisfy (3.4). IfE⊂Sp

i=1Bφ(xi, τ ρ), then the iterative process stops. Otherwise, there exists ˜xp+1E\Sp

i=1Bφ(xi, τ ρ). As EJu, there exists (yk)kJu converging to ˜xp+1. As distφxp+1, S)> τ ρ and distφxp+1,x˜i)> τ ρfor any i∈ {1, . . . , p}, there exists kp ∈Nsuch that distφ(ykp, S)> τ ρand distφ(ykp,x˜i)> τ ρ. We setxp=ykp.

(13)

If the iterative process does not finish, then there exists a sequence (xi)i∈N⊂Ω which satisfies (3.4). According to Proposition2.1, we have

(i, j)∈ {1, . . . , p}2, i6=j ⇒ |xixj| ≥ τ ρ

√Λ.

As Ω is bounded there exists a converging subsequence which is a contradiction. So, the iterative process is finite.

Claim 2: There exists a constantc=c(n, λ,Λ)which only depends on the dimensionnand the ellipticity coefficients λ,Λ such that, for anyx∈Ω, we have

#{i∈ {1, . . . , p}:xBφ(xi, τ ρ)} ≤c(n, λ,Λ).

We set

c(n, λ,Λ) = sup

q∈N:∃y1, . . . , yq ∈Rn, ∀i,|yi| ≤ 1

λ, i6=j ⇒ |yiyj| ≥ 1

√Λ

.

We consider a finite partition ofB 0,1

λ

of parallelepipeds whose diameter is less than 1Λ. Then, c(n, λ,Λ) is finite and less than the cardinality of such partition. For x∈ Ω, we set yi = xiτ ρ−x. According to Proposition2.1, we have

xBφ(xi, τ ρ) ⇒ yiB

0,√1 λ

, i6=j ⇒ |yiyj| ≥ √1 Λ. and then #{i∈ {1, . . . , p}:xBφ(xi, τ ρ)} ≤c(n, λ,Λ).

Claim 3: We still denote by c a generic constant depending on(n, λ,Λ). We have pβρn−1c

τn−1Hn−1(JuEτ ρ). According toClaim 2, we have

p

X

i=1

1Ju∩Bφ(xi,τ ρ)c1Ju∩Eτ ρ.

Integrating with respect toHn−1 gives

p

X

i=1

Hn−1(JuBφ(xi, τ ρ))≤cHn−1(JuEτ ρ). Proposition2.1 condition givesB(xi, λ1/2τ ρ))⊂Bφ(xi, τ ρ)) and then

p

X

i=1

Hn−1(JuB(xi, λ1/2τ ρ))≤cHn−1(JuEτ ρ). (3.5) According to Theorem3.3, there existsβ >0 such that, for anyi∈ {1, . . . , p}, we have

Hn−1(JuB(xi, λ1/2τ ρ))≥βλn−12 τn−1ρn−1. (3.6) Inequalities (3.5) and (3.6) conclude the proof of Claim 3.

Claim 4: We have

lim sup

ρ→0+

Ln(Eρ)

2ρ ≤ (1 +τ)nc τn−1 δ

Références

Documents relatifs

The Brunn–Minkowski theory (or the theory of mixed volumes) of convex bodies, devel- oped by Minkowski, Aleksandrov, Fenchel, et al., centers around the study of geometric

This last step does NOT require that all indicator be converted - best if only a small percent need to be reacted to make the change visible.. Note the

of difficulties. Hence the unified field equation derived in this paper is based also on the generalization of Maxwell equations, the modifications of Ricci theorem and

In the second part of the paper, we show that every metric measure space satisfying the classical Brunn-Minkowski inequality can be approximated by discrete metric spaces with

By our assumption and Theorem 2, there is a skew-symmetric pairing on SstrA(Kn) which is nondegenerate modulo its maximal divisible subgroup.. Let ,&amp;#x3E; be

(2.2) As a consequence of Theorem 2.1 we obtain the following result which can be used in fluid mechanics to prove the existence of the pressure when the viscosity term is known to

We propose a simple, computationally feasible, nonparamet- ric estimator of the boundary Minkowski content L 0 (G) which is “nearly universal” in the sense that it provides a

The final step in obtaining the coarea formula for mappings defined on an H n -rectifiable metric space is the proof of the area formula on level sets:.. Theorem 7 (The Area Formula