• Aucun résultat trouvé

Higher integrability of the gradient for the Thermal Insulation problem

N/A
N/A
Protected

Academic year: 2021

Partager "Higher integrability of the gradient for the Thermal Insulation problem"

Copied!
35
0
0

Texte intégral

(1)

HAL Id: hal-03207514

https://hal.archives-ouvertes.fr/hal-03207514v3

Preprint submitted on 11 Oct 2021

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

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

Higher integrability of the gradient for the Thermal Insulation problem

Camille Labourie, Emmanouil Milakis

To cite this version:

Camille Labourie, Emmanouil Milakis. Higher integrability of the gradient for the Thermal Insulation problem. 2021. �hal-03207514v3�

(2)

Higher integrability of the gradient for the Thermal Insulation problem

C. Labourie, E. Milakis

Abstract

We prove the higher integrability of the gradient for local mini- mizers of the thermal insulation problem, an analogue of De Giorgi’s conjecture for the Mumford-Shah functional. We deduce that the sin- gular part of the free boundary has Hausdorff dimension strictly less than n − 1.

AMS Subject Classifications: 35R35, 35J20, 49N60, 49Q20.

Keywords: Thermal Insulation, Higher Integrability, Free Boundary Problems.

Contents

1 Introduction 2

2 Generalities about minimizers 3

2.1 Definition . . . 3 2.2 Properties . . . 4

3 Porosity of the singular part 10

4 Higher integrability of the gradient 12

5 Dimension of the singular part 18

Appendices 20

A Generalities about BV functions 20

B A Robin problem 23

B.1 Statement . . . 23 B.2 Hölder continuity up to the boundary . . . 24 B.3 Gradient estimates . . . 29

C Extracts from David’s book 30

(3)

1 Introduction

We fix a bounded connected set Ω ⊂ Rn. The thermal insulation problem consists in minimizing the functional

I(A, u) :=

ˆ

A

|∇u|2dLn+ ˆ

∂A

|u|2dHn−1+ Ln(A) (1) among all pairs (A, u) where A ⊂ Rn is an admissible domain and u ∈ W1,2(A) is a function such that u = 1 for Ln-a.e. on Ω. Here, u is the trace of u on ∂A.

The problem has been studied by Caffarelli–Kriventsov in [4], [12] and Bucur–Giacomini–Luckhaus in [2], [3]. The authors transpose the problem to a slightly different setting in order to apply the direct method of the calculus of variation. The authors represent a pair (A, u) by the function u1A and relax the functional on SBV. The new problem consists in minimizing the functional

F (u) :=

ˆ

Rn

|∇u|2dLn+ ˆ

Ju

(u2+ u2) dHn−1+ Ln({ u > 0 }) (2) among all functions u ∈ SBV(Rn) such that u = 1 Ln-a.e. on Ω. The definition of Ju and u, u are given in Appendix A. This new setting is more suited to a direct minimization since it enjoys the compactness and closure properties of SBV. In parenthesis, there always exist functions u ∈ SBV(Rn) such that u = 1 Ln-a.e. on Ω and F (u) < ∞. For example, u := 1Bwhere B is an open ball containing Ω. In [4, Theorem 4.2], Caffarelli and Kriventsov prove that the SBV problem has a solution u. A key point property of solutions is that there exists 0 < δ < 1 (depending on n, Ω) such that spt(u) ⊂ B(0, δ−1) and

u ∈ { 0 } ∪ [δ, 1] Ln-a.e. on Rn. (3) This property has also been proved in [2]. On another note, some minimality criteria have been proved by calibrations in [5].

The main goal of the present article is to prove that there exists p > 1 such that |∇u|2 ∈ Lploc(Rn\ Ω) (Theorem 4.1). A parallel property was conjectured by De Giorgi for minimizers of the Mumford-Shah functional and solved by De Lellis and Focardi in the planar case ([6]) and then De Philippis and Figalli ([9]) in the general case. Our proof is inspired by the technique of [9] and it relies on three key properties: the Ahlfors-regularity of the free boundary, the uniform rectifiability of the free boundary and the ε-regularity theorem. In particular, this implies a porosity property which means that the singular part Σ of the free boundary has many holes in a quantified way. In contrast to the Mumford-Shah situation, the ε-regularity theorem describes a regular part of the boundary as a pair of graphs rather

(4)

than just one graph. The minimizers satisfies an elliptic equation with a Robin boundary condition at the boundary rather than a Neumann boundary condition. We present the technique of [9] in a different way by singling out a higher integrability lemma and a covering lemma and by removing the need of [9, Lemma 3.2] (the existence of good radii). Once we establish the higher integrability of the gradient, we are also able to conclude that the dimension of Σ is strictly less than n − 1 (Theorem 5.1). The link between the higher integrability of the gradient and the dimension of the singular part has been first observed for the Mumford-Shah functional by Ambrosio, Fusco, Hutchinson in [8]. An open question of Caffarelli–Kriventsov hints that for all minimizers in the planar case, Σ is empty and the optimal exponent is p = ∞ (see also Remark 5.2).

Acknowledgement. We would like to thank Guido De Philippis for his helpful correspondence concerning estimates for elliptic equations.

2 Generalities about minimizers

2.1 Definition

Notations. Our ambient space is an open set X of Rn. One can think of X as Rn\ Ω. For x ∈ Rn and r > 0, B(x, r) is the open ball centered in x and of radius r. If there is no ambiguity, it is simply denoted by Br. Given an open ball B := B(x, r) and a scalar t > 0, the notation tB means B(x, tr). Given a set A ⊂ Rn, the indicator function of A is denoted by 1A. Given two sets A, B ⊂ Rn, the notation A ⊂⊂ B means that there exists a compact set K ⊂ Rn such that A ⊂ K ⊂ B.

Given u ∈ SBVloc(X), we denote by K the support of the singular part of Du:

K := spt(|u − u|Hn−1 Ju) (4a)

:= spt(Hn−1 Ju). (4b)

For x ∈ K and r > 0 such that B(x, r) ⊂ X, we define ω2(x, r) := r−(n−1)

ˆ

B(x,r)

|∇u|2dLn, (5a)

β2(x, r) := r−(n+1)inf

V

ˆ

K∩B(x,r)

d(y, V )2dHn−1(y)

!12

, (5b)

where V runs among (n − 1) planes V ⊂ Rnpassing through x. When there is an ambiguity, we will write βK,2 instead of β2. We have gathered some definitions and results from the theory of BV functions in the introduction of Appendix A.

(5)

For any open ball B such that B ⊂ X, we define a competitor of u in B as a function v ∈ SBVloc(X) such that v = u Ln-a.e. on X \ B. We fix a constant δ ∈ ]0, 1[ for all the paper.

Definition 2.1. We say that u ∈ SBVloc(X) is a local minimizer if 1. for Ln-a.e. x ∈ X, we have u ∈ { 0 } ∪ [δ, δ−1];

2. for all open balls B such thatB ⊂ X, for all competitors v of u in B, ˆ

B

|∇u|2dLn+ ˆ

Ju∩B

(u2+ u2) dHn−1+ Ln({ u > 0 } ∩ B)

≤ ˆ

B

|∇v|2dLn+ ˆ

Jv∩B

(v2+ v2) dHn−1+ Ln({ v > 0 } ∩ B). (6)

As a first consequence, we have that u, u ∈ { 0 } ∪ [δ, δ−1] everywhere in X. In particular, u ≥ δ everywhere on Su. For all open balls B such that B ⊂ X, we have

ˆ

B

|∇u|2dLn+ ˆ

Ju∩B

(u2+ u2) dHn−1< ∞. (7)

This shows that |∇u|2 ∈ L1loc(X) and that Su is Hn−1-locally finite in X.

In X \ Su, the function u is Wloc1,2 and locally minimizes its Dirichlet energy.

Therefore, u is harmonic (and thus continuous) in X \Su. We conclude that in each connected component of X \ Su, we have either u > δ everywhere or u = 0 everywhere.

2.2 Properties

The next results (Ahlfors-regularity, uniform rectifiability and ε-regularity theorem) also hold true for the almost-minimizers of [12, Definition 2.1]. We are going to cite [4, Corollary 3.3 and Theorem 5.1].

Proposition 2.2 (Ahlfors-regularity). Let u ∈ SBVloc(X) be a local mini- mizer. There exists 0 < r0 ≤ 1 and C ≥ 1 (both depending on n, δ) such that the following holds true.

1. For all x ∈ X, for all 0 ≤ r ≤ r0 such that B(x, r) ⊂ X, ˆ

B(x,r)

|∇u|2dLn+ Hn−1(K ∩ B(x, r)) ≤ Crn−1. (8)

2. For all x ∈ Su, for all 0 ≤ r ≤ r0 such that B(x, r) ⊂ X,

Hn−1(K ∩ B(x, r)) ≥ C−1rn−1. (9)

(6)

Corollary 2.3. Let u ∈ SBVloc(X) be a local minimizer.

(i) We have K = Su = Ju and Hn−1(K \ Ju) = 0.

(ii) The set Au := { u > 0 } \ K is open and ∂Au = K.

Proof. It is straighforward by definition that K ⊂ Ju ⊂ Su. On the other hand, property (9) shows that Su ⊂ K. We justify that Hn−1(K \ Ju) = 0.

The jump set Ju is Borel and Hn−1 locally finite in X, so for Hn−1-a.e.

x ∈ X \ Ju,

r→0lim

Hn−1(Ju∩ B(x, r))

rn−1 = 0 (10)

(see [14, Theorem 6.2]). We draw our claim from the observation that this limit contradicts (9).

We study the set Au. We recall that the function u is continuous in X \ K (since it coincides with u outside Su) and u ∈ { 0 } ∪ [δ, 1] everywhere in X \ K. As a consequence, the sets

Au := { u > 0 } \ K, (11)

Bu := { u = 0 } \ K (12)

are open subsets of X \ K and thus of X. The space X is the disjoint union

X = K ∪ Au∪ Bu, (13)

where Au and Bu are open and K is relatively closed, so Au⊂ Au∪ K.

We show that Su ⊂ Au. Let us suppose that there exists x ∈ Suand r > 0 such that Au∩ B(x, r) = ∅. Then B(x, r) \ K ⊂ { u = 0 } so we have u = 0 Ln-a.e. on B(x, r) and thus x is a Lebesgue point of u (a contradiction). We conclude that Su ⊂ Au and in turn K ⊂ Au so Au = Au∪ K.

We are going to apply [7] to justify that K is locally contained in a uniformly rectifiable set. We underline that our local minimizers are not quasiminimizers as in [7, Definition 7.21]. In Appendix C, we have sum- marised the relevant results of [7] and how their proofs adapt to our case (Remark C.5).

Proposition 2.4 (Uniform Rectifiability). Let u ∈ SBVloc(X) be a local minimizer. There exists 0 < r0 ≤ 1 (depending on n, δ) such that the following holds true. For all x ∈ K and 0 < r ≤ r0 such that B(x, r) ⊂ X, there is a closed, Ahlfors-regular, uniformly rectifiable set E of dimension (n − 1) such that K ∩12B(x, r) ⊂ E. The constants for the Ahfors-regularity and uniform rectifiability depends on n, δ.

Proof. We want to show that (u, K) satisfy Definition C.1, or rather the alternative Definition given in Remark C.5. Then the Proposition will follow

(7)

from Theorem C.4. First, it is clear that (u, K) is an admissible pair. Let B be an open ball of radius r > 0 such that B ⊂ X. Let an admissible pair (v, L) be a competitor of (u, K) in B. As explained in Remark C.2, we can assume without loss of generality that L is Hn−1 locally finite. Therefore, v ∈ SBVloc(X) and Hn−1(Jv\ L) = 0. We have included more details about the construction of SBV functions in Appendix A. We can now apply the minimality inequality. We have

ˆ

B

|∇u|2dLn+ ˆ

Ju∩B

(u2+ u2) dHn−1+ Ln({ u > 0 } ∩ B)

≤ ˆ

B

|∇v|2dLn+ ˆ

Jv∩B

(v2+ v2) dHn−1+ Ln({ v > 0 } ∩ B) (14) so

ˆ

B

|∇u|2dLn+ δ2Hn−1(Ju∩ B) + Ln({ u > 0 } ∩ B)

≤ ˆ

B

|∇v|2dLn+ δ−2Hn−1(Jv∩ B) + Ln({ v > 0 } ∩ B). (15) We omit the term Ln({ u > 0 } ∩ B) at the left and we bound the term Ln({ v > 0 } ∩ B) at the right by ωnrn where ωn is the Lebesgue volume of the unit ball. We can replace Ju by K at the left since Hn−1(K \ Ju) = 0.

We can replace Jv by L at the right since Hn−1(Jv\ L) = 0. It follows that Hn−1(K ∩ B) ≤ δ−4Hn−1(L ∩ B) + δ−2∆E + δ−2ωnrn (16) where

∆E :=

ˆ

B

|∇v|2− ˆ

B

|∇u|2dLn. (17)

We are going to cite the ε-regularity theorem for our problem [12, The- orem 14.1]. Contrary to the ε-regularity theorem for the Mumford-Shah problem, it does not require ω2(x, r) to be small. It says that when K is very close to a plane, K is given by a pair of smooth graphs. We describe this situation in the next definition.

Given a point x ∈ Rn, a vector en∈ Sn−1, we can decompose each point y ∈ Rn under the form y = x + (y0 + ynen) where y0 ∈ en and yn ∈ R.

Then for all function f : en → R, we define the graph of f in the coordinate system (x, en) as

Γ(x,en)(f ) := { y ∈ Rn| yn= f (y0) } . (18) Definition 2.5. Let u ∈ SBVloc(X) be a local minimizer. Let x ∈ K and R > 0 be such that B(x, R) ⊂ X. Let 0 < α ≤ 1. We say that K is C1,α- regular in B := B(x, R) ⊂ X if it satisfies the three following conditions.

(8)

(i) There exists a vector en ∈ Sn−1 and two functions fi: en → R (i = 1, 2) such that f1 ≤ f2 and

K ∩ B =

 [

i=1,2

Γ(x,en)(fi)

∩ B. (19)

The functions f1, f2 are C1,α and

R−1|fi|+ |∇fi|+ Rα[∇fi]α≤ 1

4. (20)

(ii) There are two possible cases. The first case is

(u > 0 in { y ∈ B | yn< f1(y0) or yn> f2(y0) }

u = 0 in { y ∈ B | f1(y0) < yn< f2(y0) } (21) The second case is f1 = f2 and

(u > 0 in { y ∈ B | yn> f1(y0) }

u = 0 in { y ∈ B | yn< f1(y0) } (22) or inversely.

Theorem 2.6 (ε-regularity theorem). Let u ∈ SBVloc(X) be a local mini- mizer and let x ∈ K.

(i) For all ε > 0, there exists ε1> 0 (depending on n, δ, β) such that the following holds true. For r > 0 such that B(x, r) ⊂ X and β2(x, r) + r ≤ ε1, we have ω2(x,r2) ≤ ε.

(ii) There exists ε > 0, C ≥ 1 and 0 < α < 1 (both depending on n, δ) such that the following holds true. For r > 0 such that B(x, r) ⊂ X and β2(x, r) + r ≤ ε, the set K is C1,α-regular in B(x, C−1R).

This last result is specific to local minimizers and does not hold true for the general almost-minimizers of [12, Definition 2.1].

Proposition 2.7. Let u ∈ SBVloc(X) be a local minimizer. Let x ∈ K and R > 0 be such that B(x, R) ⊂ X. We assume that K is regular in B := B(x, R) (Definition 2.5) and we denote by Γi the graph of fi in B (i = 1, 2). Then for each i = 1, 2, u|Ai solves the Robin problem

 ∆u = 0 in Ai

νiu − ui = 0 in Γi, (23) where

A1 := { y ∈ B | yn< f1(y0) } (24) A2 := { y ∈ B | yn> f2(y0) } (25) and νi is the inner normal vector to Ai.

(9)

Proof. We only detail the case (21) of Definition 2.5 and we prove the Propo- sition for i = 2. We partition B in three sets (modulo Ln)

A1:= { z ∈ B | yn< f1(y0) } (26) A2:= { z ∈ B | yn> f2(y0) } (27) A3:= { z ∈ B | f1(y0) < xn< f2(y0) } . (28) The first paragraph is devoted to detail a few generalities about traces and upper/lower limits. We consider a general v ∈ L(B) ∩ Wloc1,2(B \ K) such that v = 0 in A3. For each i = 1, 2, there exists vi ∈ L1i) such that for Hn−1-a.e. x ∈ Γi,

r→0limr−n ˆ

Ai∩B(x,r)

|v(y) − vi(x)| dLn(y) = 0. (29) The boundary Γi is C1 so for all x ∈ Γi, there a vector νi(x) ∈ Sn−1 such that

r→0limr−nLn((Ai∆Hi+(x)) ∩ B(x, r)) = 0 (30) where

Hi+(x) := { y ∈ Rn| (y − x) · νi(x) > 0 } . (31) Therefore, (29) is equivalent to

r→0limr−n ˆ

Hi+(x)∩B(x,r)

|v(y) − vi(x)| dLn(y) = 0. (32)

For Hn−1-a.e. x ∈ Γ2, we detail the relationship between v(x)2+ v(x)2 and vi(x). We fix x ∈ Γ2\ Γ1 such that (32) is satisfied for i = 2. We have v = 0 on A3 and B(x, r) is disjoint from A1 for small r > 0 so

r→0limr−n ˆ

(X\A2)∩B(x,r)

|v| dLn= 0 (33)

which is equivalent to

r→0limr−n ˆ

H2(x)∩B(x,r)

|v| dLn= 0. (34)

Combining (32) for i = 2 and (34), we deduce

v(x)2+ v(x)2= v2(x)2. (35) Next, we fix x ∈ Γ1 ∩ Γ2 such that (32) holds true for i = 1 and i = 2.

The surfaces Γ1 and Γ2 have necessary the same tangent plane at x and the vectors νi are opposed. Combining (32) for i = 1 and i = 2, we deduce

v2+ v2= (v1)2+ (v2)2. (36)

(10)

We come back to our local minimizer u ∈ SBVloc(X). We fix ϕ ∈ Cc1(B).

For ε ∈ R, we define v : X → R by v :=

(u + εϕ in A2

u in X \ A2 (37)

It is clear that { v 6= u } ⊂⊂ B and that v is C1 in X \ K. As K is Hn−1 locally finite in X, we have v ∈ SBVloc(X) and Sv ⊂ K. Remember that u ≥ δ in A1 ∪ A2, and u = 0 in A3. We take ε small enough so that ε|ϕ| < δ. As a consequence v > 0 in A1 ∪ A2 and v = 0 in A3. Let us check the multiplicities on the discontinuity set. As we have seen before, Jv ∩ B ⊂ Γ1 ∪ Γ2. We observe that for x ∈ Γ2 such that the trace u2(x) exists, we have

v2(x) = u2(x) + εϕ(x) (38) and for x ∈ Γ1 such the trace u1(x) exists, we have

v1(x) = u1(x). (39)

Using the previous discussion, we deduce that for Hn−1-a.e. on Γ2\ Γ1,

v2+ v2 = (u2+ εϕ)2 (40)

= (u2+ u2) + 2εϕu2+ ε2|ϕ|2 (41) that for Hn−1-a.e. on Γ2∩ Γ1,

v2+ v2 = (u2+ εϕ)2+ (u1)2 (42)

= (u2+ u2) + 2εϕu2+ ε2|ϕ|2 (43) and that for Hn−1-a.e. on Γ1\ Γ2,

v2+ v2= (u1)2 (44)

= u2+ u2. (45)

Finally, it is clear that ˆ

B

|∇v|2dLn= ˆ

B

|∇u|2dLn+ 2ε ˆ

A2

h∇u, ∇ϕi dLn + ε2

ˆ

A2

|∇ϕ|2dLn. (46) We plug all these informations in the minimality inequality and we obtain that

0 ≤ 2ε ˆ

A2

h∇u, ∇ϕi dLn+ 2ε ˆ

Γ2

ϕu2dHn−1+ C(ϕ)ε2. (47) As this holds true for all small ε (positive or negative), we conclude that

ˆ

A2

h∇u, ∇ϕi dLn+ ˆ

Γ2

ϕu2dHn−1= 0. (48)

(11)

3 Porosity of the singular part

The following result says that the part where K is not regular has many holes in a quantified way. It also holds true for the almost-minimizers of [12, Definition 2.1]. It is simpler to obtain than its Mumford-Shah counterpart (see [15]) because the ε-regularity theorem of [12] only requires to control the flatness.

Proposition 3.1 (Porosity). Let u ∈ SBVloc(X) be a local minimizer. There exists 0 < r0 ≤ 1, C ≥ 2 and 0 < α < 1 (all depending on n,δ) for which the following holds true. For all x ∈ K and all 0 < r ≤ r0such that B(x, r) ⊂ X, there exists a smaller ball B(y, C−1r) ⊂ B(x, r) in which K is C1,α-regular.

Proof. The letter C is a constant ≥ 1 that depends on n, δ. The letter α is the constant of Theorem 2.6. For y ∈ K and t > 0 such that B(y, t) ⊂ X, we define the L flatness

βK(y, t) := inf

V sup

z∈K∩B(y,t)

t−1d(z, V ), (49)

where the infimum is taken over the affine hyperplanes V of Rn passing through y. Note that in [7, (41.2)], the infimum is taken over all affine hyperplanes V of Rn(not necessarily passing through y); this would decrease our β number but no more than a factor 12. Indeed, if V is any hyperplane of Rn and y0 is the orthogonal projection of y onto V , then the hyperplane V − (y0− y) is passing through y so we have we have

βK(y, t) ≤ sup

z∈K∩B(y,t)

d(z, V − (y0− y)) (50)

≤ y0− y

+ sup

z∈K∩B(y,t)

d(z, V ) (51)

≤ 2 sup

z∈K∩B(y,t)

d(z, V ). (52)

We also observe that

βK,2(y, t)2 ≤ t−(n−1)Hn−1(K ∩ B(y, t))βK(y, t)2 (53) so as soon as t is small enough for the Ahlfors-regularity to hold, we have βK,2(y, t) ≤ CβK(y, t).

Let r0 be the minimum between the radius of Proposition 2.2 (Ahfors- regularity) and the radius of Proposition 2.4 (uniform rectifiability). We fix x ∈ K and 0 < r ≤ r0 such that B(x, r) ⊂ X. According to Proposition 2.4, there exists an Ahlfors-regular and uniformly rectifiable set E such that K ∩ 12B(x, r) ⊂ E. Moreover, the constants for the Ahfors-regularity and uniform rectifiability depends on n, δ. For y ∈ E and t > 0, we define as before

βE(y, t) := inf

V sup

z∈E∩B(y,t)

t−1d(z, V ), (54)

(12)

where the infimum is taken on the set of all affine hyperplanes V of Rn passing through x. As E is Ahlfors-regular and uniformly rectifiable, the Weak Geometric Lemma [7, (73.13)] states that for all ε > 0, the set

{ (y, t) | y ∈ E, 0 < t < diam(E), βE(y, t) > ε } (55) is a Carleson set. This means that for all ε > 0, there exists C0(ε) ≥ 1 (depending on n, δ, ε) such that for all y ∈ E and all 0 < t < diam(E),

ˆ t

0

ˆ

E∩B(y,t)

1{ βE(z,s)>ε }(z) dHn−1(z)ds

s ≤ C0(ε)tn−1. (56) We only apply this property with y := x. We observe that for all z ∈ K ∩ B(x,14r) and for all 0 < s ≤ 14r, we have K ∩ B(z, s) ⊂ E ∩ B(z, s) so βK(z, s) ≤ βE(z, s). Thus for all 0 < t < diam(K ∩ 12B(x, r)) such that t ≤ 14r, we have

ˆ t

0

ˆ

K∩B(x,t)

1{ βK(z,s)>ε }(z) dHn−1(z)ds

s ≤ C0(ε)tn−1. (57) We only apply this property with t := 14diam(K ∩ B(x, r)).

ˆ t

0

ˆ

K∩B(x,t)

1{ βK(z,s)>ε }(z) dHn−1(z)ds

s ≤ C0(ε)tn−1. (58) Note that C−1r ≤ t ≤ 14r, where the first inequality comes from the Ahlfors- regularity of K. We are going to deduce from (58) that for all ε > 0, there exists C(ε) ≥ 1, a point z ∈ K ∩ B(x, t) and a radius s such that C(ε)−1t ≤ s ≤ t and βK(z, s) ≤ ε. We proceed by contradiction for some C(ε) to be precised. We have therefore

ˆ t

0

ˆ

K∩B(x,t)

1{ βK(z,s)>ε }(z) dHn−1(z)ds s

≥ Hn−1(K ∩ B(x, t)) ˆ t

C(ε)−1t

ds s

(59)

≥ Hn−1(K ∩ B(x, t)) ln(C(ε)) (60)

≥ C−1tn−1ln(C(ε)) (61)

This contradicts (58) if C(ε) is too big compared to C0(ε).

We fix ε > 0 (to be precised soon) and we assume that we have a cor- responding pair (z, s) as above. In particular, βK,2(z, s) ≤ CβK(z, s) ≤ Cε.

According to the second statement of Theorem 2.6, we can fix ε (depending on n, δ) so that if r0 ≤ ε, then K is C1,α-regular in B(z, C−1s).

(13)

4 Higher integrability of the gradient

Theorem 4.1. Let u ∈ SBVloc(X) be minimal. There exists 0 < r0 ≤ 1, C ≥ 1 and p > 1 (depending on n, δ) such that the following holds true. For all x ∈ X, for all 0 ≤ r ≤ r0 such that B(x, r) ⊂ X,

ˆ

1 2B(x,r)

|∇u|2pdLn≤ Crn−p. (62) The higher integrability is well known for weak solutions of elliptic sys- tems ([10, Theorem 2.1]). In this case, the proof consists in combining the Caccioppoli-Leray inequality and the Sobolev-Poincaré inequality to deduce that |∇u|n+22n satisfies a reverse Hölder inequality. The higher integrability is then an immediate consequence of Gehring Lemma. In our case, u is still a weak solution of an elliptic system but we lack information about the regularity of K to carry out this method.

We draw inspiration from [9] but we simplify the proof by singling out an higher integrability lemma (Lemma 4.2 below) and a covering lemma (Lemma 4.3 below) and by removing the need of [9, Lemma 3.2] (the exis- tence of good radii).

Proof of Theorem 4.1. There exists 0 < r0 ≤ 1, such that for all x ∈ X and all 0 ≤ R ≤ r0 such that B(x, R) ⊂ X, one can applies Lemma 4.2 below in the ball B(x, R) to the function v := R|∇u|2. The assumption (i) follows from the Ahlfors-regularity of K (Proposition 2.2). The assumption (ii) follows from the porosity (Proposition 3.1). The assumption (iii) follows from interior/boundary gradient estimates for the Robin problem and from the Ahlfors-regularity. In particular, the interior estimate can be derived from the subharmonicity of |∇u|2 in X \ K and the boundary estimate is detailed in Lemma B.1 in Appendix B.

Lemma 4.2. We fix a radius R > 0 and an open ball BRof radius R. Let K be a closed subset of BRand v : BR→ R+ be a non-negative Borel function.

We assume that there exists C0 ≥ 1 and 0 < α ≤ 1 such that the following holds true.

(i) For all ball B(x, r) ⊂ BR,

C0rn−1≤ Hn−1(K ∩ B(x, r)) ≤ C0rn−1. (63) (ii) For all ball B(x, r) ⊂ BR centered in K, there exists a smaller ball

B(y, C0−1r) ⊂ B(x, r) in which K is C1,α-regular (Definition 2.5).

(iii) For all ball B(x, r) ⊂ BR such that K is disjoint from B(x, r) or K is C1,α-regular in B(x, r) (Definition 2.5), we have

sup

1 2B(x,r)

v(x) ≤ C0 R r



. (64)

(14)

Then there exists p > 1 and C ≥ 1 (depending on n, C0) such that

1 2BR

vp ≤ C. (65)

The proof of Lemma 4.2 takes advantage of the following covering lemma.

We use the notation Γ(x,en) defined at line (18). The assumption (ii) says that in each double ball 2Bk, the set E is an union of Lipschitz graphs which are close to an hyperplane.

Lemma 4.3 (Covering Lemma). Let E ⊂ Rn be a bounded set. Let (Bk) be a family of open balls of center xk∈ Rnand radius Rk> 0. We assume that

(i) for all k 6= l, 2Bk∩ Bl= ∅;

(ii) for all k, for all x ∈ E ∩ 2Bk, there exists a vector en ∈ Sn−1 and a

1

2-Lipschitz function f : en → R such that |f | ≤ 12Rk and

x ∈ Γ(xk,en)(f ) ∩ 2Bk⊂ E. (66) Let 0 < r ≤ infkRk. There exists a sequence of open balls (Di)i∈I of radius r and centered in E \S

kBk such that E \[

k

Bk⊂[

i∈I

Di (67)

and the balls (20−1Di)i∈I are pairwise disjoint and disjoint from S

kBk. Proof. Let 0 < r0 ≤ infkRk. We introduce the set

F := E \[

k

Bk. (68)

The goal is to cover F with a controlled number of balls of radius r0. Let r be a radius 0 < r ≤ r0 which will be precised during the proof. As F is bounded, there exists a maximal sequence of points (xi) ∈ F such that B(xi, r) ⊂ Rn\S

kBk and |xi− xj| ≥ r. For i 6= j, we have |xi− xj| ≥ r so the balls (B(xi,12r))i are disjoint. Next, we show that

F ⊂[

i

B(xi, 10r). (69)

Let x ∈ F . If B(x, r) ⊂ Rn\S

kBk, then by maximality of (xi), there exists i such that x ∈ B(xi, r) ⊂ B(xi, 10r). Now we focus on the case where there exists an index k0 such that B(x, r) ∩ Bk0 6= ∅. The radius of Bk0 is denoted by R and we assume without loss of generality that its center is 0.

As x ∈ F = E \S

kBk and B(x, r) ∩ B(0, R) 6= ∅, we have R < |x| < R + r.

(15)

We are going to build a point y ∈ E such that R + r < |y| < R + 7r and

|x − y| < 9r.

Since r ≤ R, we observe that x ∈ B(0, 2R). According to the assumptions of the lemma, there exists two scalars 0 < ε, L ≤ 12, a vector en∈ Sn−1 and a L-Lipschitz f : en → R such that |f | ≤ εR and

x ∈ { y ∈ B(0, 2R) | yn= f (y0) } ⊂ E. (70) Here, we have decomposed each point y ∈ Rn under the form y = y0+ ynen where y0 ∈ n and yn∈ R. The estimate R < |x| < R + r can be rewritten

R <

x0+ f (x0)en

< R + r. (71)

We consider t ≥ 1 such that |tx0+ f (x0)en| = R + 4r and we estimate how close tx0 is to x0. We have

x0

≥ q

R2− |f (x0)|2 (72)

tx0

≤ q

(R + 4r)2− |f (x0)|2 (73) so

tx0− x0

q

(R + 4r)2− |f (x0)|2− q

R2− |f (x0)|2 (74)

≤ 4Rr + 8r2 q

R2− |f (x0)|2

. (75)

We assume r ≤ 18R and we recall that |f (x0)| ≤ εR with ε ≤ 12 so this simplifies to

tx0− x0

≤ 5r

1 − ε2 < 6r. (76)

Next, we define y := tx0+ f (tx0)en and we recall that f is L-Lipschitz with L ≤ 12 to estimate

y − [tx0+ f (x0)en] =

f (tx0) − f (x0)

(77)

< 3r. (78)

Since |tx0+ f (x0)en| = R + 4r, this yields

R + r < |y| < R + 7r. (79) We also estimate

|y − x| ≤

tx0− x0 +

f (tx0) − f (x0)

(80)

< 9r. (81)

(16)

As r ≤ 18R, the inequalities (79) imply y ∈ B(0, 2R) and thus y ∈ E. We are going to justify that B(y, r) ⊂ Rn\S

kBk. We recall that B(0, 2R) is disjoint from all the other balls of the family (Bk). By (79), we observe that B(y, r) ⊂ B(0, R + 8r) \ B(0, R) (82)

⊂ B(0, 2R) \ B(0, R) (83)

and our claim follows. By maximality of the family (xi), there exists i such that |y − xi| < r and in turn by (81), |x − xi| < 10r. We finally choose r := 101r0. The balls (Di) are given by Di:= B(xi, 10r) = B(xi, r0).

Proof of Lemma 4.2. We observe that any ball B ⊂ BRand for p ≥ 1, ˆ

B

vpdLn= ˆ

0

Ln(B ∩ { vp > t }) dt (84)

= p ˆ

0

sp−1Ln(B ∩ { v > s }) ds (85) and for M ≥ 1,

p ˆ

1

sp−1Ln(B ∩ { v > s }) ds

≤ p

X

h=0

ˆ Mh+1

Mh

sp−1Ln(B ∩ { v > s }) ds

(86)

≤ p

X

h=0

ˆ Mh+1

Mh

sp−1ds

!

Ln(B ∩ { v > Mh}) (87)

≤ (Mp− 1)

X

h=h

MhpLn(B ∩ { v > Mh}). (88)

Thus, it suffices to prove that there exists N > M ≥ 1, C ≥ 1 (depending on n, C0) such that for all h ≥ 0,

Ln(12BR∩ { v > Mh}) ≤ CRnN−h (89) and then take p > 1 such that MpN−1 < 1.

To simplify the notations, we change the constant C0 so that (ii) yields that 40B(y, C0−1r) ⊂ B(x, r) and that K is C1,α-regular in 4B(y, C0−1r).

Let M := max { 4C0,14C02} ≥ 4. We define for h ≥ 1,

Ah := { x ∈12BR\ K | v > Mh} . (90) The proof is based on the fact that Ah is at distance ∼ M−hR from K and has many holes of size ∼ M−hR near K. We justify more precisely these

(17)

observations. For the first one, let h ≥ 1, let x ∈ Ah and assume that B(x, C0M−hR) is disjoint from K. Then we use property (ii) to estimate

v(x) ≤ Mh. (91)

This contradicts the definition of Ah. We deduce that there exists y ∈ K such that |x − y| < C0M−hR. For the second observation, let h ≥ 2, let x ∈ K ∩1516BR and apply the porosity property to the ball B(x, M−hR).

We obtain an open ball B ⊂ X centered in K, of radius C0−1M−hR and such that K is C1,α-regular in 4B. Then by point (iii) and since M ≥ 14C02,

sup

2B

v ≤ 14C02Mh≤ Mh+1 (92) In particular, 2B is disjoint from Ah+1.

We start the proof by defining for h ≥ 1,

r(h) := M−hR (93)

R(h) := 3

4 + M−h+1



R. (94)

The sequence (R(h)) is decreasing, limh→∞R(h) = 34R and R(h+1)+r(h) ≤ R(h). For each h ≥ 1, we build an index set I(h) and a family of balls (Bi)i∈I(h) as follow. First we define I(1) := ∅ and (Bi)i∈I(1) := ∅. Let h ≥ 2 be such that (Bi)i∈I(1), . . . , (Bi)i∈I(h−1)have been built. We assume that the index sets I(g), where g = 1, . . . , h−1, are pairwise disjoint. We assume that for all i ∈ Ig, the balls Bi have radius C0−1r(g) = C0−1M−gR. We assume that for all indices i, j ∈Sh−1

g=1I(g) with i 6= j, we have that 2Bi ∩ Bj = ∅ and that K is C1,α-regular in 2Bi. Then, we introduce the sets

Kh:= K ∩ BR(h)\

h−1

[

g=1

[

i∈I(g)

Bi (95)

Kh:= K ∩ BR(h+1)\

h−1

[

g=1

[

i∈I(g)

Bi. (96)

According to Lemma 4.3, there exists a sequence of open balls (Di)i∈I(h) centered in Kh of radius r(h) = M−hR such that

Kh⊂ [

i∈I(h)

Di, (97)

and such that the balls (20−1Di) are pairwise disjoint and disjoint from Sh−1

g=1

S

i∈I(g)Bi. We can assume that index set I(h) is disjoint from the sets

(18)

I(g), g = 1, . . . , h − 1. Since R(h + 1) + r(h) ≤ R(h), we observe that the balls (20−1Di) are included in

BR(h)\

h−1

[

g=1

[

i∈I(g)

Bi. (98)

Next, we apply the porosity to the balls (Di). For each i ∈ I(h), there exists an open ball Bi centered in K, of radius C0−1M−hR such that Bi ⊂ 40−1Di, K is C1,α-regular in 4 Bi and by (iii),

sup

2Bi

v ≤ 14C02Mh≤ Mh+1 (99) We should not forget to mention that for all i ∈ I(h), we have 2Bi⊂ 20−1Di so 2Bi is disjoint from all the other balls we have built so far.

Now, we estimate Ln(Ah) for h ≥ 1. We show first that the points of Ah cannot be too far from Kh. Let x ∈ Ah. We have seen earlier that there exists y ∈ K such that |x − y| < C0M−hR. We are going to show that y ∈ Kh. Since |x| ≤ 12R and M ≥ 4C0, we have

|y| ≤ 12R + C0M−hR ≤ 34R. (100) Let us assume that there exists g = 1, . . . , h − 1 and i ∈ I(g) such that y ∈ Bi. The radius of Bi is C0−1M−(h−1)R and since |x − y| < C0M−hR, we have x ∈ 2Bi. However 2Bi is disjoint from Ah by construction. We have shown that y ∈ Kh. As a consequence, there exists i ∈ I(h) such that y ∈ Di. The radius of Di is r(h) = M−hR and |x − y| < C0M−hR so

Ah ⊂ [

i∈I(h)

(1 + C0)Di. (101)

This allows to estimate

Ln(Ah) ≤ ωn(1 + C0)n|I(h)|r(h)n (102) where ωn is the Lebesgue measure of the unit ball.

Next, we want to control |I(h)|. The balls (20−1Di)i∈I(h) are disjoint and included in the set B(R(h)) \Sh−1

g=2

S

i∈I(g)Bi so by Ahlfors-regularity, C0−120−(n−1)r(h)(n−1)|I(h)| ≤ X

i∈I(h)

Hn−1(K ∩ 12−1Di) (103)

≤ Hn−1(Kh). (104)

(19)

We are going to see that Hn−1(Kh) is bounded from above by a decreasing geometric sequence. We have

Hn−1(Kh) ≤ X

i∈I(h)

Hn−1(K ∩ Di) (105)

≤ C0 X

i∈I(h)

r(h)n−1 (106)

≤ C0n+1 X

i∈I(h)

(C0−1r(h))n−1 (107)

≤ C0n+1 X

i∈I(h)

Hn−1(K ∩ Bi) (108)

≤ C0n+1Hn−1(Kh\ Kh+1). (109) We deduce

Hn−1(Kh) ≤ C0n+1Hn−1(Kh\ Kh+1) + Hn−1(K ∩ BR(h)\ BR(h+1)). (110) We rewrite this inequality as

Hn−1(Kh+1) ≤ λ−1Hn−1(Kh) + C0−(n+1)Hn−1(K ∩ BR(h)\ BR(h+1)) (111) where λ := C0n+1(C0n+1− 1)−1 > 1. Then, we multiply both sides of the inequality by λh+1:

λh+1Hn−1(Kh+1)

≤ λhHn−1(Kh) + C0−(n+1)λ−hHn−1(K ∩ BR(h)\ BR(h+1))

(112)

≤ λhHn−1(Kh) + Hn−1(K ∩ BR(h)\ BR(h+1)). (113) Summing this telescopic inequality, we obtain that for all h ≥ 1,

λhHn−1(Kh) ≤ 2Hn−1(K ∩ BR) (114)

≤ 2C0Rn−1. (115)

In summary, we have proved that for some constant C ≥ 1, λ > 1 (depending on n, C0) and for h ≥ 1

Ln(Ah) ≤ CRn(λM )−h. (116)

5 Dimension of the singular part

Notation. The Hausdorff dimension of a set A ⊂ Rn is defined by

dimH(A) := inf { s ≥ 0 | Hs(A) = 0 } . (117)

(20)

We take the convention that for s < 0, the term Hs-a-e. means everywhere and the inequality dimH(A) < 0 means A = ∅.

The goal of this section is to explain the link between the integrability exponent of the gradient and the dimension of the singular part. It has been first observed for the Mumford-Shah functional by Ambrosio, Fusco, Hutchinson in [8].

Theorem 5.1. Let u ∈ SBVloc(X) be a local minimizer. We define

Σ := { x ∈ K | K is not regular at x } . (118) For p > 1 such that |∇u|2 ∈ Lploc(X), we have

dimH(Σ) ≤ max { n − p, n − 8 } < n − 1. (119) Remark 5.2. In dimension n ≤ 7, Caffarelli–Kriventsov have shown that if a point x ∈ K is at the boundary of two local connected components where u > 0 or if it is a 0-density point of { u = 0 }, then x is a regular point ([4, Theorem 8.2]). In dimension n = 2, they show furthermore that if x is at the boundary of a connected component of { u = 0 }, then it is a regular point ([4, Corollary 9.2]). Thus in the planar case, a point of Σ must be an acculumation point of connected components of { u = 0 }. There is however no known example of such a situation.

Theorem 5.1 will be proved very easily with the help of [4, Theorem 8.2]

and the following well-known result.

Lemma 5.3. Let v ∈ Lploc(X) for some p ≥ 1 and let s < n. Then, for Hn−p(n−s)-a.e. x ∈ X,

r→0limr−s ˆ

B(x,r)

v dLn= 0. (120)

Proof. Without loss of generality, we assume v ≥ 0. We start with the case p = 1. We define µ as the measure vLnand we want to show that for Hs-a.e.

x ∈ X, we have

r→0limr−sµ(B(x, r)) = 0. (121) If s < 0, the limit is indeed 0 for every x ∈ X. In the case 0 ≤ s < n, we fix a closed ballB ⊂ X, a scalar λ > 0 and a set

A := { x ∈ B | lim sup

r→0

r−sµ(B(x, r)) > λ } . (122) According to [1, Theorem 2.56],

µ(A) ≥ λHs(A). (123)

As A ⊂ B and µ is a Radon measure, we have µ(A) < ∞. Then (123) gives Hs(A) < ∞ and since s < n, Ln(A) = 0. The measure µ is dominated by

(21)

Ln so µ(A) = 0 and now (123) gives Hs(A) = 0. We can take a sequence of scalars λk→ 0 to deduce

Hs({ x ∈ B | lim sup

r→0

r−sµ(B(x, r)) > 0 }) = 0. (124) We can then conclude that

Hs({ x ∈ X | lim sup

r→0

r−sµ(B(x, r)) > 0 }) = 0. (125) by covering X with a a sequence of closed balls Bk⊂ X.

Now we come to the general case p ≥ 1. Let us fix t < n. For x ∈ X and for r > 0, the Hölder inequality shows that

r

 n−np

ˆ

B(x,r)

v dLn≤ ˆ

B(x,r)

vpdLn

!1p

(126) so

r

 n+ptn

p

ˆ

B(x,r)

v dLn≤ r−t ˆ

B(x,r)

vpdLn

!1

p

. (127)

We apply the first part to see that for Ht-a.e. x ∈ X,

r→0limr



n+ptnpˆ

B(x,r)

v dLn= 0. (128)

The scalar t such that s = n +ptnp is t := n − p(n − s) < n.

Proof of Theorem 5.1. According to Lemma 5.3, we have for Hn−p-a.e. x ∈ X,

r→0limω2(x, r) = 0 (129)

and according to [4, Theorem 8.2], the set n

x ∈ X ∩ Σ lim

r→0ω2(x, r) = 0o

(130) has a Hausdorff dimension ≤ n − 8.

Appendices

A Generalities about BV functions

We recall a few definitions and results from the theory of BV functions ([1]).

We work in an open set X of the Euclidean space Rn(n > 1). When a point x ∈ X is given, we abbreviate the open ball B(x, r) as Br.

(22)

Let u ∈ L1loc(X). The upper and lower approximate limit of u at at a point x ∈ X are defined by

u(x) := inf { t ∈ R | lim

r→0r−n ˆ

{ u>t }∩Br

(u − t) dLn= 0 } , (131) u(x) := sup { t ∈ R | lim

r→0r−n ˆ

{ u<t }∩Br

(t − u) dLn= 0 } . (132) The functions u, u : X → R are Borel and satisfies u ≤ u. We have two examples in mind. We say that x is a Lebesgue point if there exists t ∈ R such that

r→0lim Br

|u − t| dLn= 0. (133)

We then have u(x) = u(x) = t and we denote t by ˜u(x). The set of non- Lebesgue points x ∈ X is called the singular set Su. Both the set Su and the function X \ Su → R, x 7→ ˜u(x) are Borel ([1, Proposition 3.64]). The Lebesgue differentiation theorem states that for Ln-a.e. x ∈ X, we have x ∈ X \ Su and u(x) = ˜u(x). We say that x is a jump point if there exist two real numbers s < t and a (unique) vector νu(x) ∈ Sn−1 such that

r→0lim H+∩Br

|u(y) − s| dLn(y) = 0 (134a)

r→0lim H∩Br

|u(y) − t| dLn(y) = 0, (134b) where

H+ := { y ∈ Rn| (y − x) · νu(x) > 0 } (135a) H := { y ∈ Rn| (y − x) · νu(x) < 0 } . (135b) We then have u(x) = t and u(x) = s. The set of jump points x ∈ X is called the jump set Ju. Both the set Ju and the function Ju → Sn−1, x 7→ νu(x) are Borel ([1, Proposition 3.69]).

Here we summarize [1, Proposition 3.76, 3.78]. Let u ∈ BV(X). The singular set Su is Hn−1 rectifiable and Hn−1(Su\ Ju) = 0. According to the Besicovitch derivation theorem, we can write

Du = Dau + Dsu (136)

where Dau is the absolutely continuous part of Du with respect to Ln and Dsu is the singular part of Du with respect to Ln. As a consequence, there exists a unique vector-valued map ∇u ∈ L1(X; Rn), the approximate gra- dient, such that Dau = ∇uLn. The measures Ln and kDsuk are mutually singular which means that there exists a Borel set S ⊂ X such that

Ln(S) = kDsuk(Rn\ S) = 0. (137)

Références

Documents relatifs

One could also use &#34;scissors and glue&#34; techniques to produce a concrete Riemann surface leading to an example of the desired kind.. However, implicit in such

The first algorithmic proof of the asymmetric version, where the condition requires the existence of a number in (0, 1) for each event, such that the probability of each event to

In effect it says that if we randomly give values (0 or 1) to a large proportion of the variables which are also randomly selected (this procedure is called a random restriction),

In this second article, we give tractable conditions under which the limit flow is Lipschitz continuous and its links with unique- ness of solutions of rough differential equations..

Using a different ortho- normal basis would not affect the absolute value Jdet (T)[. Similar remarks apply when the range is a Hilbert space.. Extremal mappings

The last inequality is due to the fact that, in order to create a stack of white books of length F s, we need first to form a stack of white books with length s 2 s, removing toward

It is my great pleasure to welcome you all to the thirteenth meeting of national tuberculosis programme managers in the WHO Eastern Mediterranean Region, as well as the second

On the other hand, a careful analysis of localisation the numerical range of quasi-sectorial contractions [6, 1], allows to lift the estimate in Theorem 3.3 and in Corollary 3.6 to