• Aucun résultat trouvé

Regularity of the Optimal Shape for the First Eigenvalue of the Laplacian With Volume and Inclusion Constraints

N/A
N/A
Protected

Academic year: 2022

Partager "Regularity of the Optimal Shape for the First Eigenvalue of the Laplacian With Volume and Inclusion Constraints"

Copied!
15
0
0

Texte intégral

(1)

www.elsevier.com/locate/anihpc

Regularity of the optimal shape for the first eigenvalue of the Laplacian with volume and inclusion constraints

Tanguy Briançon

a

, Jimmy Lamboley

b,

aLycée Agora, 92800 Puteaux, France

bENS Cachan Bretagne, IRMAR, UEB, av. Robert Schuman, 35170 Bruz, France

Received 26 February 2008; accepted 12 July 2008 Available online 26 July 2008

Abstract

We consider the well-known following shape optimization problem:

λ1)= min

|Ω|=a ΩD

λ1(Ω),

whereλ1denotes the first eigenvalue of the Laplace operator with homogeneous Dirichlet boundary condition, andDis an open bounded set (a box). It is well-known that the solution of this problem is the ball of volumeaif such a ball exists in the boxD (Faber–Krahn’s theorem).

In this paper, we prove regularity properties of the boundary of the optimal shapesΩin any case and in any dimension. Full regularity is obtained in dimension 2.

©2008 Elsevier Masson SAS. All rights reserved.

Keywords:Shape optimization; Eigenvalues of the Laplace operator; Regularity of free boundaries

1. Introduction and main results

LetDbe a bounded open subset ofRd. For all open subsetΩofD, we denote byλ1(Ω)the first eigenvalue of the Laplace operator inΩ, with homogeneous boundary conditions, and byuΩ a normalized eigenfunction, that is

⎧⎪

⎪⎪

⎪⎪

⎪⎩

uΩ=λ1(Ω)uΩ inΩ,

uΩ=0 on∂Ω,

Ω

u2Ω=1.

* Corresponding author.

E-mail addresses:briancon_tanguy@yahoo.fr (T. Briançon), jimmy.lamboley@bretagne.ens-cachan.fr (J. Lamboley).

0294-1449/$ – see front matter ©2008 Elsevier Masson SAS. All rights reserved.

doi:10.1016/j.anihpc.2008.07.003

(2)

We are interested here in the regularity of the optimal shapes of the following shape optimization problem, where a(0,|D|)(|D|denotes the Lebesgue measure ofD):

Ωopen, ΩD, |Ω| =a, λ1)=min

λ1(Ω);Ωopen, Ω⊂D,|Ω| =a . (1)

By a well-known theorem of Faber and Krahn, if there is a ballBDwith|B| =a, then this ball is an optimal shape and it is unique, up to translations (and up to sets of zero capacity).

Here we address the question of existence of aregularoptimal set in all cases.

Existence of aquasi-open optimal setΩ may be deduced from a general existence result by G. Buttazzo and G. Dal Maso (see [5]) for an extended version of (1), where the variable setsΩ are not necessarily open. An optimal shapeΩmay not be more than a quasi-open set ifDis not connected (we reproduce in the appendix the example mentioned in [4]). On the other hand, it is proved in [4] or [12] that such an open optimal setΩalways exists for (1) and, if moreoverDis connected, then all optimal shapesΩare open. More precisely, it is proved in [4] that, for any D,uΩ is locally Lipschitz continuous inD. If moreoverDis connected, thenΩ coincides with the support ofuΩ(and is therefore open). Let us summarize this as follows (see also [13]):

Proposition 1.1.AssumeDis open and bounded. The problem(1)has a solutionΩ, anduΩis non-negative and locally Lipschitz continuous inD. IfDis connected,Ω= {xD, uΩ>0}.

Moreover, we have

uΩ+λ1)uΩ0 inD, (2)

which means thatuΩ+λ1)uΩ is a positive Radon measure.

Here, we are interested in the regularity of∂Ωitself, and we prove the following theorem:

Theorem 1.2.AssumeDis open, bounded and connected. Then any solution of (1)satisfies:

1. Ωhas locally finite perimeter inDand Hd1

(∂Ω\Ω)D

=0, (3)

whereHd1is the Hausdorff measure of dimensiond−1, andΩ is the reduced boundary(in the sense of sets with finite perimeter, see[9]or[10]).

2. There existsΛ >0such that uΩ+λ1)uΩ=√

ΛHd1∂Ω,

in the sense of distribution inD, whereHd1∂Ωis the restriction of the(d−1)-Hausdorff measure to∂Ω. 3. Ωis an analytic hypersurface inD.

4. Ifd=2, then the whole boundary∂ΩDis analytic.

We use the same strategy as in [3] (where the regularity is studied for another shape optimization problem). The- orem 1.2 essentially relies on the proof of the equivalence of (1) with a penalized version for the constraint|Ω| =a, as stated in Theorem 1.5 below. Once we have this penalized version, we can use techniques and results from [1] (see also [11] and [3]).

Remark 1.3.According to the results in [1], the third point in Theorem 1.2 is a direct consequence of the second one which says thatuΩ is a “weak solution” in the sense of [1]. To obtain the full regularity of the boundary ford=2, the fact thatuΩ is a weak solution is not sufficient, and more information has to be deduced from the variational problem. The approach is essentially the same as in Theorem 6.6 and Corollary 6.7 in [1]. The necessary adjustments are given at the end of this paper.

Remark 1.4.According to the result of [14,15,6,8], it is likely that full regularity of the boundary may be extended to higher dimension (d6?), and therefore that the estimate (3) can be improved.

But this needs quite more work and is under study.

(3)

By a classical variational principle, we know that, for allΩDopen, λ1(Ω)=

Ω

|∇uΩ|2=min

Ω

|∇u|2, uH01(Ω),

Ω

u2=1

. (4)

Here,λ11(Ω)for all open set ΩD with|Ω| =a. Since[ΩΩλ1(Ω)λ1(Ω) ], it follows that λ11(Ω)for all open setΩDwith|Ω|a. Coupled with (4), this leads to the following variation property ofΩanduΩ(see [4] for more details), where we denoteu=uΩ, λa=λ1), andΩv= {xD;v(x)=0}:

λa=

D

|∇u|2=min

D

|∇v|2;vH01(D),

D

v2=1, |Ωv|a

. (5)

Let us rewrite this as follows. ForwH01(D), we denoteJ (w)=

D|∇w|2λa

Dw2. Then applying (5) with v=w/(

Dw2)1/2, we obtain thatuis a solution of the following optimization problem:

J (u)J (w), for allwH01(D), with|Ωw|a. (6)

One of the main ingredient in the proof of Theorem 1.2 is to improve the variational property (6) in two directions, as stated in Theorem 1.5 below. The approach is local.

LetBRbe a ball included inDand centered on∂ΩuD. We define F=

vH01(D), uvH01(BR) .

Forh >0, we denote byμ(h)the biggestμ0 such that,

vFsuch thatah|Ωv|a, J (u)+μ|Ωu|J (v)+μ|Ωv|. (7) We also defineμ+(h)as the smallestμ+0 such that,

vFsuch thata|Ωv|a+h, J (u)+μ+|Ωu|J (v)+μ+|Ωv|. (8) The following theorem is a main step in the proof of Theorem 1.2:

Theorem 1.5.Letu, BRandFas above. Then forRsmall enough(depending only onu, aandD), there existsΛ >0 andh0>0such that,

h(0, h0), 0< μ(h)Λμ+(h) <+∞, and, moreover,

hlim0μ+(h)=lim

h0μ(h)=Λ. (9)

Remark 1.6.We can compare the existence ofμ+(h)with Theorem 2.9 in [4]. This theorem shows that there exists μ+such that

D

|∇u|2

D

|∇v|2+λa

1−

D

v2 +

+μ+

|Ωv| −a ,

forvH01(D)and|Ωv|a. The difference with [4] is that, in (8), we have the termλa[1−

Dv2](not only the positive part), but we allowed only perturbations inBR. We cannot expect to have something like (8) for perturbations in allD(because we may findvwith|Ωv|> aandJ (v) <0, so limt→+∞J (t v)= −∞).

In the next section, we will prove Theorem 1.5. In the third section, we will prove Theorem 1.2. In Appendix A, we discuss the caseDnon-connected.

(4)

2. Proof of Theorem 1.5

In the next lemma, we give an Euler–Lagrange equation for our problem. The proof follows the steps of the Euler–

Lagrange equation in [7].

Lemma 2.1 (Euler–Lagrange equation). Let u be a solution of (6). Then there exists Λ0 such that, for all ΦC0(D,Rd),

D

2(DΦ∇u,u)

D

|∇u|2∇ ·Φ+λa

D

u2∇ ·Φ=Λ

Ωu

∇ ·Φ.

Proof. We start by a general remark that will be useful in the rest of the paper. IfvH01(D)and ifΦC0(D,Rd), we definevt(x)=v(x+t Φ(x)); therefore, fort small enough,vtH01(D). A simple calculus gives (whent goes to 0),

|Ωvt| = |Ωv| −t

Ωv

∇ ·Φ+o(t ),

J (vt)=J (v)+t

D

2(DΦ∇v· ∇v)

D

|∇v|2∇ ·Φ+λa

D

v2∇ ·Φ

+o(t ).

Now we apply this withv=uandΦsuch that

Ωu∇ ·Φ >0. Such aΦ exists, otherwise we would get, using thatD is connected,Ωu=Dor∅a.e. We have|Ωut|<|Ωu|fort >0 small enough and, by minimality,

J (u)J (ut)

=J (u)+t

D

2(DΦ∇u,u)

D

|∇u|2∇ ·Φ+λa

D

u2∇ ·Φ

+o(t ), and so,

D

2(DΦ∇u,u)

D

|∇u|2∇ ·Φ+λa

D

u2∇ ·Φ0. (10)

Now, we take Φ with

Ωu∇ ·Φ =0. LetΦ1be such that

Ωu∇ ·Φ1=1. Writing (10) withΦ+ηΦ1and letting η goes to 0, we get (10) with thisΦ and, using−Φ, we get (10) with an equality instead of the inequality. For a generalΦ, we use this equality withΦΦ1(

Ωu∇ ·Φ)(we have

Ωu∇ ·Φ1(

Ωu∇ ·Φ))=0), and we get the result with

Λ=

D

2(DΦ1u,u)

D

|∇u|2∇ ·Φ1+λa

D

u2∇ ·Φ10, using (10). 2

Remark 2.2.We will have to prove that, in fact,Λ >0.

Let us remind our notations: letube a solution of (6), and letBRbe a ball included inDand centered on∂ΩuD.

We define F=

vH01(D), uvH01(BR) .

Before proving Theorem 1.5, we give the following useful lemma:

Lemma 2.3.Letu, BRandFas above. Then there exists a constantCsuch that, forRsmall enough,

vF, J (v)1 2

BR

|∇v|2C.

(5)

Proof. We know thatλ1(BR)=λ1(B1)/(R2)(we just use the change of variablexx/R). IfRis small enough we have:

λ1(BR)1, 4λa

λ1(BR)1/2. (11)

LetvF; souvH01(BR), and using the variational formulation ofλ1(BR), we get uv2L2(BR)(uv)2L2(BR)

λ1(BR) . We deduce that,

v2L2(BR)2∇(uv)2L2(BR)

λ1(BR) +2u2L2(BR)

4∇v2L2(BR)

λ1(BR) + C λa

,

(we use (11)) whereCdepends only on theL2norms ofuand his gradient. Now we have J (v)

D

|∇v|2λa

4∇v2L2(BR)

λ1(BR) + C λa

,

and we get the result using (11). 2

Remark 2.4.This lemma is interesting for two reasons. The first one is thatJ is bounded from below on F. The second one is that, ifvnF is a sequence such thatJ (vn)is bounded, then ∇vnL2(BR) is also bounded. Since vn=uoutsideBR we deduce thatvnis bounded inH01(D)(and so weakly converges up to a sub-sequence. . . ).

Proof of Theorem 1.5. We divide our proof into four parts. LetΛ0 be as in Lemma 2.1.

First part:ΛΛΛμμμ+++(h) <(h) <(h) <+∞+∞+∞.

We start the proof by showing thatμ+(h)is finite. SinceBRis centered on the boundary on∂Ωu, we first show:

0<|ΩuBR|<|BR|.

The first inequality comes from the fact thatΩuis open. The second one comes from the following lemma:

Lemma 2.5.Letωbe an open subset ofD, and letube a solution of (6). If|Ωuω| = |ω|, then

u=λau inω, and thereforeωΩu.

Proof of Lemma 2.5. Sinceu >0 a.e. onω, we definevH01(D)byv=uoutsideωand−v=λauinω. From the strict maximum principle, we getv >0 onωand|Ωv| = |Ωu|. By minimality (J (u)J (v)) we have,

ω

(u− ∇v)·(u− ∇v+2∇v)λa

ω

(uv)(u+v)0,

ω

|∇u− ∇v|2+λa

ω

(uv)(2uuv)0,

(we use thatuvH01(ω)and−v=λauinω). We get thatu=va.e. inωand by continuityu=v >0 everywhere inω. 2

(6)

If|ΩuBR| = |BR|, applying this lemma toω=BR, we would getΩuBR=BR, which is impossible sinceBR is centered on∂Ωu. IfRis small enough we can also suppose,

0<|Ωu\BR|<|D\BR|.

For the first inequality, we need that|BR|< a, and for the second one we needa <|D| − |BR|.

Leth >0 be such thath <|BR| − |ΩuBR|(and so, ifvF with|Ωv|a+h, then|ΩvBR|<|BR|). Let n)an increasing sequence to+∞. There existsvnFsuch that|Ωvn|a+hand,

J (vn)+μn

|Ωvn| −a+

= min

v∈F,|Ωv|a+h

J (v)+μn

|Ωv| −a+

. (12)

For this we use Remark 2.4, and so the functionalJ (v)+μn(|Ωv| −a)+is bounded by below forvF. Moreover, a minimizing sequence for this functional is bounded inH01(D)and so weakly converges inH01(D), strongly inL2(D) and almost everywhere (up to a sub-sequence) to somevn. Using the lower semi-continuity ofv

D|∇v|2for the weak convergence, the strong convergence inL2(D)and the lower semi-continuity ofv→ |Ωv|for the convergence almost everywhere we see thatvnis such that (12) is true.

If|Ωvn|athen (8) is true withμn, so we will suppose to the contrary that|Ωvn|> afor alln.

Step 1. Euler–Lagrange equation forvn.If we setbn= |Ωvn|, thenvnis also solution of J (vn)= min

v∈F,|Ωv|bn

J (v).

With the same proof as in Lemma 2.1, we can write an Euler–Lagrange equation forvn inBR. That is, there exists Λn0 such that, forΦC0(BR,Rd),

D

2(DΦ∇vn· ∇vn)

D

|∇vn|2∇ ·Φ+λa

D

vn2∇ ·Φ=Λn

Ωvn

∇ ·Φ. (13)

Step 2.ΛΛΛnnnμμμnnn.There existsΦC0 (BR)such that

Ωvn∇ ·Φ=1. Letvnt(x)=vn(x+t Φ(x)). We havevtnF fort0 small enough, and using derivation results recalled in the proof of Lemma 2.1 and|Ωvn|> a, we get

a <|Ωvt

n| = |Ωvn| −t+o(t )a+h, J (vnt)=J (vn)+t Λn+o(t ).

Now we use (12) withv=vtnin order to get, J (vn)+μn

|Ωvn| −a

J (vn)+t Λn+o(t )+μn

|Ωvn| −ta , and dividing byt >0 and lettingtgoes to 0, we finally getΛnμn. Step 3.vnstrongly converges to somev.Using (12) withv=u, we get

J (vn)+μn

|Ωvn| −a

J (u) (14)

and so, using Remark 2.4, we can deduce thatvnweakly converge inH01(up to a sub-sequence) to somevFwith

|Ωv|a+h. We also have the strong convergence inL2(D)and the convergence almost everywhere. SinceJ is bounded from below onF, we see from (14) thatμn(|Ωvn| −a)is bounded and we get limn→∞|Ωvn| =a, and so

|Ωv|a. FromJ (vn)J (u), we getJ (v)lim infJ (vn)J (u)and sovis a solution of (6). Finally we can write, using (12), thatJ (vn)J (v)and we get, using the strong convergence ofvninL2,

lim sup

n→∞

D

|∇vn|2

D

|∇v|2.

We also have, with weak convergence inH01(D)that

D

|∇v|2lim inf

n→∞

D

|∇vn|2.

(7)

We deduce that limn→∞vnL2(D)= ∇vL2(D). With the weak-convergence, this gives the strong convergence of vntovinH01(D).

Step 4.limlimlimΛΛΛnnn===ΛΛΛ.We see thatvis a solution of (6), so we can apply Lemma 2.1 to get that there exists aΛvsuch that

ΦC0 D,Rd

,

D

2(DΦ∇v· ∇v)

D

|∇v|2∇ ·Φ+λa

D

v2∇ ·Φ=Λv

Ωv

∇ ·Φ.

We haveu=voutsideBR so, using this equation and the Euler–Lagrange equation foruwe see thatΛv=Λ. Now, we write the Euler–Lagrange forvnandΦC0(D,Rd)such that

Ωv∇ ·Φ=0,

D

2(DΦ∇vn· ∇vn)

D

|∇vn|2∇ ·Φ+λa

D

vn2∇ ·Φ=Λn

Ωvn

∇ ·Φ,

and, using the strong convergence ofvntov, we get that

nlim→∞Λn= lim

n→∞

D2(DΦ∇vn· ∇vn)

D|∇vn|2∇ ·Φ+λa

Dv2n∇ ·Φ

Ωvn∇ ·Φ

=

D2(DΦ∇v· ∇v)

D|∇v|2∇ ·Φ+λa

Dv2∇ ·Φ

Ωv∇ ·Φ

=Λ.

Since limμn= +∞we get the contradiction from Steps 2 and 4, and soμ+(h)is finite.

To conclude this first part, we now have to see thatΛμ+(h). LetΦC0be such that

Ωu∇ ·Φ= −1, and let ut(x)=u(x+t Φ(x)). Using the calculus in the proof of Lemma 2.1 we have, fort0 small enough,

a|Ωut| =a+t+o(t )a+h, J (ut)=J (u)t Λ+o(t ).

Now, using (8), we have

J (u)+μ+(h)aJ (u)t Λ+μ+(h)(a+t )+o(t ), and we getΛμ+(h).

Second part:limlimlimμμμ+++(h)(h)(h)===ΛΛΛ.

We first see thatμ+(h) >0 forh >0. Indeed, ifμ+(h)=0 we write for everyϕC0(BR)with{ϕ=0}< h, J (u)J (u+t ϕ), so

u=λau inBR,

which contradicts 0<|ΩuBR|<|BR|.

Letε >0 andhn>0 a decreasing sequence tending to 0. Becausehμ+(h)is non-increasing, we just have to see that limμ+(hn+εfor a sub-sequence ofhn. IfΛ >0, letε∈ ]0, Λ[and 0< αn:=μ+(hn)ε < μ+(hn);

ifΛ=0, let 0< αn=μ+(hn)/2< μ+(hn). There existsvnsuch that J (vn)+αn

|Ωvn| −a+

= min

v∈F,|Ωv|a+hn

J (v)+αn

|Ωv| −a+ .

Sinceαn< μ+(hn)we see that|Ωvn|> a(otherwise we writeJ (u)J (vn)+αn(|Ωvn| −a)+). We now have 4 steps that are very similar to the 4 steps used in the previous part to show thatμ+(hn)is finite.

Step 1. Euler–Lagrange equation forvn.IfvF is such that|Ωv||Ωvn|, we have J (vn)J (v). Then, as in Lemma 2.1 we can write the Euler–Lagrange equation (13) forvninBRfor someΛn.

(8)

Step 2.ΛΛΛnnnαααnnn.Since|Ωvn|> athe proof is the same as Step 2 in the first part, withαninstead ofμn. Step 3.vnstrongly converge to somev.As in Step 3 above, we just write,

J (vn)+αn

|Ωvn| −a+

J (u),

to get (up to a sub-sequence) thatvnweakly converges inH01(D), strongly inL2(D)and almost-everywhere tovF.

We havea <|Ωvn|a+hnand so limn→∞|Ωvn| =a. As in Step 3 above, we deduce thatvis a solution of (6), and using

J (vn)+αn

|Ωvn| −a

J (v), we get the strong convergence inH01(D).

Step 4.limlimlimΛΛΛnnn===ΛΛΛ.The proof is the same as in Step 4 of the first part of the proof. We write the Euler–Lagrange equation forvinDand useu=voutsideBR. We get that limΛn=Λby lettingngo to+∞in the Euler–Lagrange equation forvninBR(using the strong convergence ofvn).

We can now conclude this second part: ifΛ >0, we have, fornlarge enough, μ+(hn)ε=αnΛnΛ+ε,

and soμ+(hn+2ε.

IfΛ=0 we have

μ+(hn)/2=αnΛnε, and so 0μ+(hn)2ε.

In both cases, we haveΛμ+(hn+2ε.

Third part:limlimlimμμμ(h)(h)(h)===ΛΛΛ.

Lethnbe a sequence decreasing to 0, and letε >0. Becausehμ(h)is increasing, we just have to show that limn→∞μ(hnεfor a sub-sequence ofhn.

We first see thatμ(h)Λ. LetΦC0(BR,Rd)be such that

BR∇ ·Φ=1 and letut=u(x+t Φ(x))fort0.

We have (using the proof of Lemma 2.1), ah|Ωut| =at+o(t )a, J (ut)=J (u)+t Λ+o(t ).

Now, using (7), we have

J (u)+μ(h)aJ (u)+t Λ+μ(h)(at )+o(t ), and we getμ(h)Λ.

Letvnbe a solution of the following minimization problem, J (vn)+

μ(hn)+ε

|Ωvn| −(ahn)+

= min

w∈F,|Ωw|a

J (w)+

μ(h)+ε

|Ωw| −(ahn)+

. (15) We will first see that,

ahn|Ωvn|< a.

If|Ωvn| =awe have, J (u)+

μ(hn)+ε

|Ωu|J (vn)+

μ(hn)+ε

|Ωvn|J (w)+

μ(hn)+ε

|Ωw|, forwFwithahn|Ωw|awhich contradicts the definition ofμ(hn).

(9)

Now, if|Ωvn|< ahn, we haveJ (vn)J (vn+t ϕ)for everyϕC0(BR)with|{ϕ=0}|< ahn− |Ωvn|.

And we get that−vn=λavn inBR and so, we havevn≡0 onBR orvn>0 onBR, but this last case contradicts

|Ωvn|< a. Ifvn≡0 onBR, becausevn=uoutsideBR, we getuH01(BR), and usingJ (u)J (vn),

BR

|∇u|2λa

BR

u20.

We now deduce (u≡0 onBR) thatλaλ1(BR), which is a contradiction, at least forRsmall enough.

We now study the sequencevnin a very similar way than above.

Step 1. Euler–Lagrange equation forvn.J (vn)J (v)forvFwith|Ωv||Ωvn|, so we have an Euler–Lagrange equation (13) forvninBRfor someΛn.

Step 2.ΛΛΛnnn(μ(μ(μ(h(h(hnnn)))+++ε)ε)ε).Since|Ωvn|< a, we takeΦC0(BR,Rd)with

BR∇ ·Φ= −1 andvnt(x)=vn(x+ t Φ(x))fort0 small. We have|Ωvt

n| = |Ωvn| +t+o(t )aandJ (vnt)=J (vn)Λnt+o(t )and writing (15) with w=vtnwe get the result.

Step 3.vnstrongly converge to somev.As in Step 3 above we just write that J (vn)+

μ(hn)+ε

|Ωvn| −(ahn)

J (u)+

μ(hn)+ε hn,

to get (up to a sub-sequence) thatvnweakly converge inH01(D), strongly inL2(D)and almost-everywhere tovF. We haveahn<|Ωvn|aand so limn→∞|Ωvn| =a. As in Step 3 above, we deduce thatvis a solution of (6), and using

J (vn)+

μ(hn)+ε

|Ωvn| −(ahn)

J (v)+

μ(hn)+ε

|Ωv| −(ahn)+ , we get the strong convergence inH01(D).

Step 4.limlimlimΛΛΛnnn===ΛΛΛ.The proof is exactly the same as in Step 4 above in the study of the limit ofμ+(hn).

Now we have, using steps 2 and 4, fornlarge enough, ΛεΛnμ(hn)+εΛ+ε,

and so limn→∞μ(hn)=Λ.

Fourth part:Λ >Λ >Λ >000.

We would like to show thatΛ >0 (which impliesμ(h) >0 forhsmall enough). We argue by contradiction and we suppose thatΛ=0. The proof is very close to the proof of Proposition 6.1 in [3]. We start with the following proposition:

Proposition 2.6. Assume Λ=0. Then, there exists η a decreasing function with limr0η(r)=0 such that, if x0BR/2andB(x0, r)BR/2with|{u=0} ∩B(x0, r)|>0, then

1

r

∂B(x0,r)

uη(r). (16)

Proof of Proposition 2.6. Letx0, rbe as above, and we setBr=B(x0, r). Letvbe defined by, −v=λau inBr,

v=u on∂Br,

andv=uoutsideBr. We havev >0 onBr. We get, using (8),

Br

|∇u|2− |∇v|2

λa

Br

u2v2

μ+

ωdrd{u=0} ∩Br, (17)

(10)

we also get (using−v=λauinBr),

Br

|∇u|2− |∇v|2

λa

Br

u2v2=

Br

(uv)· ∇(uv+2v)−λa

Br

u2v2

=

Br

(uv)2+λa

Br

(uv)2. (18)

Now, with the same computations as in [1,11] (withλauinstead off) we get, {u=0} ∩Br

1 r

∂Br

u 2

C

Br

(uv)2. (19)

Now, using (17), (18) and (19) we get the result. 2

End of proof of Theorem 1.5. Now, the rest of the proof is the same as Proposition 6.2 in [3] with λau instead off χΩu. The idea is that, from the estimate (16) of Proposition 2.6,∇utends to 0 at the boundary, and consequently the measureu does not charge the boundary∂Ωu. It follows that −u=λauinBR, which, by strict maximum principle, contradicts thatuis zero on some part ofBR. 2

3. Proof of Theorem 1.2

LetΩbe a solution of (1). Thenu=uΩis a solution of (6), and thus satisfies Proposition 1.1 and Theorem 1.5;

moreover,Ω=Ωu. Like in the previous section, we work inB, a small ball centered in∂Ωu. Since the approach is local, we will show regularity for the part of∂Ωuincluded inB; butB can be centered on every point of∂ΩuD, so this is of course enough to lead to the announced results in Theorem 1.2.

Coupled with Remark 1.3, we conclude that it is sufficient to prove:

(a) Ωhas finite perimeter inBandHd1

(∂Ω\Ω)B

=0, (b) uΩ+λ1)uΩ=√

ΛHd1∂ΩinB, (c) ifd=2, ∂ΩB=ΩB.

⎫⎪

⎪⎭ (20) We use the same arguments as in [1] and [11], but we have to deal with the term in

u2instead of

f u(in [11]). So we first start with the following technical lemma.

Lemma 3.1.There existC1, C2, r0>0such that, forB(x0, r)Bwithrr0, if 1

r

∂B(x0,r)

uC1 then u >0onB(x0, r),

if 1

r

∂B(x0,r)

uC2 then u≡0onB(x0, r/2). (21)

Proof. The first point comes directly from the proof of Proposition 2.6. We take the samevand, using equation (19), we see that there existsC1such that if 1r

∂B(x0,r)uC1, then|{u=0} ∩B(x0, r)| =0.

For the second part we argue as in Theorem 3.1 in [2]. We will denoteBr forB(x0, r). In this proof,C denotes (different) constants which depend only ona, d, D, uandB, but not onx0orr.

Letε >0 small and such that{u=ε}is smooth (true for almost everyε), letDε=(Br\Br/2)∩ {u > ε}andvεbe defined by

⎧⎪

⎪⎩

vε=λau inDε, vε=u inD\Br, vε=u inBr∩ {}, vε=ε inBr/2∩ {u > ε}.

(11)

We see thatuvε is harmonic inDε.

We now show that(vεu)εis bounded inH1(D), for smallε >0. Letϕbe inC0(Br)with 0ϕ1 andϕ≡1 onBr/2. LetΨ =(1ϕ)u+εϕ=u+ϕ(εu). We have:

Ψu=0=vεuon∂Br

∂Dε(Br\Br/2) , and

Ψu=εu=vεuuon∂Dε∂Br/2,

so using thatvεuis harmonic, we get−uvεu0 onDεand,

Dε

(vεu)2

Dε

u)2.

Now, using that∇Ψ = ∇u(1ϕ)(ϕ)u+εϕand theLbounds foruand∇u, we see thatvεuis bounded inH1(D).

Now, up to a subsequence,vεweakly converges inH01(D)tovsuch that:

⎧⎨

v=λau in(Br\Br/2)Ωu, v=u inD\Br,

v=0 inBr/2

Br∩ {u=0}

.

Using (7) withh= |Br/2|, andu=vinD\Br, we have:

Br

|∇u|2λa

Br

u2+μ(h)|ΩuBr|

Br

|∇v|2λa

Br

v2+μ(h)|ΩvBr|,

and so,

Br/2

|∇u|2μ(h)|ΩuBr/2|

Br\Br/2

(vu)· ∇(uv+2v)−λa

Br\Br/2

v2u2 +λa

Br/2

u2

lim inf

ε0 2

Dε

(vεu)· ∇vελa

Dε

vε2u2 +λa

Br/2

u2

=lim inf

ε0 2

∂Br/2∩{u>ε}

u)∂vε

∂n +2λa

Dε

(vεu)uλa

Dε

vε2u2 +λa

Br/2

u2

=lim inf

ε0 2

∂Br/2∩{u>ε}

u)∂vε

∂n +λa

Dε

2uvεu2vε2 +λa

Br/2

u2

lim inf

ε0 2

∂Br/2∩{u>ε}

u)vε·n+λa

Br/2

u2, (22)

wheren is the outward normal ofDε and so the inward normal ofBr/2. Letwεbe such that,

⎧⎨

wε=λau onBr\Br/2, wε=u on∂Br∩ {u > ε}, wε=ε on

∂Br∩ {}

∂Br/2.

Becausewεεon ∂(Br \Br/2)and super-harmonic inBr \Br/2, we get that wεε inBr \Br/2. In particular wεvε=εin∂Dε(Br\Br/2). Moreover, we also havewεvε on∂Dε(∂Br∂Br/2), and sincewεvεis harmonic inDε, we getwεvεinDε. Usingwε=vε=εon∂Br/2∩ {u > ε}, we can now compare the gradients of wεandvεon this set,

0−∇vε·n−∇wε·n on∂Br/2∩ {u > ε}. (23)

Références

Documents relatifs

If d = 2, in order to have the regularity of the whole boundary, we have to show that Theorem 6.6 and Corollary 6.7 in [1] (which are for solutions and not weak-solutions) are

Chipalkatti [4] introduced a notion of regularity for coherent sheaves on a Grassmannian G and proved that his definition gives a nice resolution of any such sheaf ([4], Theorem

Dual parametrization: Since our results are stated for the analytic functionals (4), we may apply them to the dual parametrization of convex sets instead of the parametrization with

In this paper, we prove regularity properties of the boundary of the optimal shapes Ω ∗ in any case and in any dimension.. Full regularity is obtained in

First, we mention that the numerical results satisfy the theoretical properties proved in [9] and in Theorem 1.3: the lack of triple points, the lack of triple points on the

Our main objective is to obtain qualitative properties of optimal shapes by exploiting first and second order optimality conditions on (1) where the convexity constraint is

In this section we will prove that any isolated connected component of K is made of a finite number of Jordan arcs.. THEOREM

PIN010DUK, Proper holomorphic mappings of strictly pseudoconvex domains, (Russian), Dokl. 1; English translation