• Aucun résultat trouvé

An asymmetric Neumann problem with weights

N/A
N/A
Protected

Academic year: 2022

Partager "An asymmetric Neumann problem with weights"

Copied!
14
0
0

Texte intégral

(1)

www.elsevier.com/locate/anihpc

An asymmetric Neumann problem with weights

M. Arias

a

, J. Campos

a

, M. Cuesta

b

, J.-P. Gossez

c,

aDepartamento de Matemática Aplicada, Universidad de Granada, 18071 Granada, Spain bLMPA, Université du Littoral, 50, rue F. Buisson, BP 699, 62228 Calais, France cDépartement de Mathématique, C.P. 214, Université Libre de Bruxelles, 1050 Bruxelles, Belgium

Received 17 March 2006; accepted 21 July 2006 Available online 8 January 2007

Abstract

We prove the existence of a first nonprincipal eigenvalue for an asymmetric Neumann problem with weights involving the p-Laplacian (cf. (1.2) below). As an application we obtain a first nontrivial curve in the corresponding Fuˇcik spectrum (cf. (1.4) below). The case where one of the weights has meanvalue zero requires some special attention in connexion with the (PS) condition and with the mountain pass geometry.

©2007 Elsevier Masson SAS. All rights reserved.

Résumé

Nous démontrons l’existence d’une première valeur propre non principale pour un problème de Neumann asymétrique avec poids faisant intervenir lep-laplacien (cf. (1.2) ci-dessous). Comme application nous obtenons une première courbe non triviale dans le spectre de Fuˇcik correspondant (cf. (1.4) ci-dessous). Le cas où l’un des poids est de moyenne nulle demande une attention particulière en liaison avec la condition de Palais–Smale et avec la géométrie du col.

©2007 Elsevier Masson SAS. All rights reserved.

MSC:35J60; 35P30

Keywords:Neumann problem; Asymmetry;p-Laplacian; Weights; Fuˇcik spectrum; Cerami (PS) condition

1. Introduction

In a previous work [2], we investigated the eigenvalues of the following asymmetric Dirichlet problem with weights:

pu=λ

m(x)(u+)p1n(x)(u)p1

inΩ, u=0 on∂Ω, (1.1)

wherepis thep-Laplacian,Ωis a bounded domain inRNandm, nsatisfy some summability conditions together withm+≡0, n+≡0. We proved the existence of a first nonprincipal positive eigenvalue for (1.1). Various appli- cations were given to the study of the Fuˇcik spectrum and to the study of nonresonance. The construction of this

* Corresponding author.

E-mail addresses:marias@ugr.es (M. Arias), campos@ugr.es (J. Campos), cuesta@lmpa.univ-littoral.fr (M. Cuesta), gossez@ulb.ac.be (J.-P. Gossez).

0294-1449/$ – see front matter ©2007 Elsevier Masson SAS. All rights reserved.

doi:10.1016/j.anihpc.2006.07.006

(2)

distinguished eigenvalue was obtained by applying a version of the mountain pass theorem to the functional

Ω|∇u|p restricted to the manifold{uW01,p(Ω):

Ω[m(u+)p+n(u)p] =1}. In this process the (PS) condition was shown to hold at all levels and the geometry of the mountain pass was derived from the observation thatϕ1(m)and−ϕ1(n) were strict local minima (whereϕ1(m)denotes the normalized positive first eigenfunction of the Dirichletp-Laplacian with weightm).

Our purpose in the present paper is to investigate the corresponding Neumann problem:

pu=λ

m(x)(u+)p1n(x)(u)p1

inΩ, ∂u

∂ν =0 on∂Ω, (1.2)

whereνdenotes the unit exterior normal. When trying to adapt the preceding approach to the present situation, the relevant functional is still

Ω|∇u|pbut now restricted to the manifold Mm,n:=

uW1,p(Ω):Bm,n(u):=

Ω

m(u+)p+n(u)p

=1

. (1.3)

A first difficulty arises in connexion with the (PS) condition. It turns out that the (PS) condition remains satisfied at all levels when

Ωm=0 and

Ωn=0, but it is not satisfied anymore at level 0 when

Ωm=0 or

Ωn=0. In this latter case, which we will call the singular case, we do not know whether the (PS) condition still holds at all positive levels (see Remark 3.4). However one can show that the Palais–Smale condition of Cerami (abbreviated into (PSC)) holds at all positive levels. Another difficulty arises when dealing with problem (1.2), which is now connected with the geometry of the functional. It turns out that in the singular case, at least one of the two natural candidates for local minimum fails to belong to the manifoldMm,n. To bypass this difficulty we will consider a minimax procedure defined from a family of paths having free endpoints (cf. (3.1)).

The existence of a first nonprincipal positive eigenvalue for (1.2) is derived in Section 3. The argument uses a version of the mountain pass theorem for aC1functional restricted to aC1manifold and which satisfies the (PSC) condition at certain levels. Section 4 is devoted to such a theorem. In Section 5 we briefly indicate some properties of the eigenvalue constructed in Section 3 as a function of the weightsm, nand in Section 6 we apply our results to the study of the Fuˇcik spectrum. Recall that the latter is defined as the setΣ of those(α, β)∈R2such that

pu=αm(x)(u+)p1βn(x)(u)p1 inΩ, ∂u

∂ν =0 on∂Ω, (1.4)

has a nontrivial solution. As in the Dirichlet case we obtain for (1.4) the existence inΣof hyperbolic-like first curves.

Note however that contrary to what was happening in the Dirichlet case, the asymptotic behaviour of these first curves does not depend on the supports of the weights (at least when the weights are bounded, cf. Proposition 6.4 and Remark 6.5).

In the preliminary Section 2 we collect some results relative to the usual eigenvalue problem

pu=λm(x)|u|p2u inΩ, ∂u

∂ν =0 on∂Ω. (1.5)

We also recall there some general definitions relative to (PS) and (PSC) conditions.

2. Preliminaries

Throughout this paperΩ will be a bounded domain inRN with Lipschitz boundary and the weightsm, nwill be assumed to belong toLr(Ω)withr >Np ifpN andr=1 ifp > N. We also assume unless otherwise stated

m+andn+≡0 inΩ. (2.1)

Solutions of (1.2) or of related equations are always understood in the weak sense, i.e.uW1,p(Ω)with

Ω

|∇u|p2uϕ=λ

Ω

m(u+)p1n(u)p1

ϕ,ϕW1,p(Ω).

Regularity results from [13] on general quasilinear equations imply that such a solutionuis locally Hölder continuous inΩ; moreover the derivation of theLestimates in [1] can be adapted to the present situation to show thatu

(3)

L(Ω). Note that if in additionm, nL(Ω)and Ω is of class C1,1, then uC1,α(Ω)for some 0< α <1 (cf. [12]).

Our main purpose in this preliminary section is to collect some results relative to the eigenvalue problem (1.5).

Clearly 0 is a principal eigenvalue of (1.5), with the constants as eigenfunctions. The search for another principal eigenvalue involves the following quantity:

λ(m)=inf

Ω

|∇u|p: uW1,p(Ω)and

Ω

m|u|p=1

. (2.2)

By (2.1),λ(m) <∞.

Proposition 2.1.

(i) Suppose

Ωm <0. Then λ(m) >0 andλ(m)is the unique nonzero principal eigenvalue;this eigenvalue is simple and admits an eigenfunction which can be chosen>0inΩ;moreover the interval]0, λ(m)[does not contain any other eigenvalue.

(ii) Suppose

Ωm >0. Thenλ(m)=0and 0 is the unique nonnegative principal eigenvalue.

(iii) Suppose

Ωm=0. Thenλ(m)=0and 0 is the unique principal eigenvalue.

Proposition 2.1 is proved in [10] (see also [6,11]) whenmL(Ω), but the arguments can easily be adapted to the present situation. We observe in this respect that in the case of an unbounded weight, Harnack’s inequality as given in [13,9] should be used instead of Vazquez maximum principle [14] to derive in case (i) that the eigenfunction can be chosen>0 inΩ. See [4] for similar considerations in the Dirichlet case. In case (i) or (ii) of Proposition 2.1, the positive eigenfunction associated toλ(m)and normalized so as to satisfy the constraint in (2.2) will be denoted byϕm. The infimum (2.2) is then achieved atϕm. In case (iii) the fact thatλ(m)=0 is easily verified by considering the sequence

vk= (1+ψ/k)1/p [

Ωm(1+ψ/k)]1/p, (2.3)

whereψis any fixed smooth function withψ0 and

Ωmψ >0. Note that in that case (iii), the infimum (2.2) is not achieved (since no constant satisfies the constraint in that case).

Let us conclude this section with some general definitions relative to the (PS) condition. LetEbe a real Banach space and let M:= {uE: g(u)=1}wheregC1(E,R)and 1 is a regular value ofg. Let fC1(E,R)and consider the restrictionf˜off toM. The differentialf˜ atuM, has a norm which will be denoted by ˜f (u)

and which is given by the norm of the restriction off (u)Eto the tangent space ofMatu Tu(M):=

vE: g(u), v =0 ,

where,denotes the pairing betweenEandE. A critical point off˜is a pointuMsuch that ˜f (u)=0;f (u)˜ is then called a critical value off˜.

We recall thatf˜is said to satisfy the (PS)ccondition (resp. (PSC)ccondition) at levelc∈Rif for any sequence ukMsuch thatf (u˜ k)cand ˜f (uk)→0 (resp.f (u˜ k)cand(1+ ukE) ˜f (uk)→0), one has that uk admits a convergent subsequence. We will also say thatf˜satisfies the (PS)c condition along bounded sequences if for any bounded sequenceukMsuch thatf (u˜ k)cand ˜f (uk)→0, one has thatukadmits a convergent subsequence. Condition (PSC)cwas introduced in [3] as a weakening of the classical (PS)ccondition.

Going back to case (iii) of Proposition 2.1, one can see that the functional

Ω|∇u|prestricted to the manifoldMm,n

(cf. (1.3)) does not satisfy the (PS)0 condition. Indeed the sequence vk from (2.3) provides an unbounded (PS)0

sequence. That the (PSC)0condition does not hold neither will follow from Proposition 4.3.

3. A first nontrivial eigenvalue

The assumptions onm, nin this section are those indicated at the beginning of Section 2. We look for nonnegative eigenvaluesλof (1.2).

(4)

Clearly the only nonnegative principal eigenvalue of (1.2) are 0, λ(m)andλ(n). Moreover multiplying byu+ oru, one easily sees that if (1.2) withλ0 has a solution which changes sign, thenλ >max{λ(m), λ(n)}. Proving the existence of such a solution which changes sign, and which in addition corresponds to a minimum value ofλ, is our purpose in this section.

As indicated in the introduction we will use a variational approach and consider the functionalA(u):=

Ω|∇u|p onW1,p(Ω), the manifoldMm,ndefined in (1.3) and the restrictionA˜ofAtoMm,n. In this context one easily verifies that λ >0 is an eigenvalue of (1.2) if and only if λ is a critical value of A. The case of the eigenvalue˜ λ=0 is particular: it is a critical value ofA˜iffMm,ncontains a constant function, i.e. iff

Ωm >0 or

Ωn >0. It follows in particular from these considerations that if

Ωm=0, thenλ(m)is a critical value ofA˜corresponding to the critical pointϕm, and similarly forλ(n)and−ϕnif

Ωn=0.

To state our main result let us introduce the following family of paths inMm,n: Γ :=

γC

[0,1], Mm,n

:γ (0)0 andγ (1)0 . (3.1)

Lemma 3.1.Γ is nonempty.

Proof. ChooseuW1,p(Ω)such that

Ωm(u+)p>0 and

Ωn(u)p>0, which is possible by (2.1), and define γ1(t ):=t1/pu+(1t )1/pufort∈ [0,1]. Using the fact thatu+anduhave disjoint supports, one obtains

Bm,n

γ1(t )

=t

Ω

m(u+)p+(1t )

Ω

n(u)pmin

Ω

m(u+)p,

Ω

n(u)p

>0.

The pathγ2(t ):=γ1(t )/(Bm,n1(t )))1/p is thus well defined and clearly belongs toΓ. 2 Define now the minimax value

c(m, n):= inf

γΓ max

uγ[0,1]A(u),˜ (3.2)

which is finite by Lemma 3.1.

Theorem 3.2.c(m, n)is an eigenvalue of (1.2)which satisfies max

λ(m), λ(n) < c(m, n). (3.3)

Moreover there is no eigenvalue of (1.2)betweenmax{λ(m), λ(n)}andc(m, n).

The rest of this section is devoted to the proof of Theorem 3.2. As indicated in the introduction, some difficulty arises in connexion with the (PS) condition.

Proposition 3.3.

(i) A˜satisfies(PS)calong bounded sequences for allc0.

(ii) A˜satisfies(PSC)cfor allc >0.

(iii) If

Ωm=0and

Ωn=0, thenA˜satisfies(PS)cfor allc0.

Remark 3.4.One can show that ifp=2, thenA˜satisfies (PS)cfor allc >0, but the casep=2 remains undecided.

On the other hand, if

Ωm=0 or

Ωn=0, thenA˜does not satisfy (PSC)0. This latter fact can be seen as in Section 2:

assuming

Ωm=0, one first observes thatvk from (2.3) provides an unbounded (PS)0sequence forA, and then one˜ applies Proposition 4.3 below; similar argument when

Ωn=0.

Proof of Proposition 3.3. (i) LetukMm,nbe a bounded (PS)csequence forA. So˜

Ω|∇uk|pcand

Ω

|∇uk|p2ukξ

εkξξTukMm,n, (3.4)

(5)

whereεk→0 and · denotes theW1,p(Ω)norm. For a subsequence and someu0W1,p(Ω), one has thatuk u0 inW1,p(Ω). Let us write forwW1,p(Ω)

ak(w):=w

Ω

m(u+k)p1n(uk)p1 w

ukTukMm,n.

Takingξ=ak(w)in (3.4), one deduces

Ω

|∇uk|p2ukw

Ω

m(u+k)p1n(uk)p1 w

Ω

|∇uk|p

εkak(w)Dεk

ukp+1 w

for some constantD; taking noww=uku0in the above, one obtains

Ω

|∇uk|p2uk(uku0)→0.

It then follows from the(S+)property thatuku0inW1,p(Ω), which yields the conclusion of part (i).

(ii) Let nowukMm,nbe a (PSC)csequence forA, with˜ c >0. So

Ω|∇uk|pcand (3.4) is replaced by

Ω

|∇uk|p2ukξ

εk

1+ ukξξTukMm,n (3.5)

whereεk→0. We will show thatukremains bounded so that part (i) applies and yields the conclusion of part (ii). Let us assume by contradiction that, for a subsequence,uk → ∞. Writevk=uk/uk. For a further subsequence and somev0W1,p(Ω), one has thatvk v0inW1,p(Ω). Since

Ω|∇uk|premains bounded, one has

Ω|∇vk|p→0 and it follows easily thatv0cst=0 and thatvkv0inW1,p(Ω). On the other hand, takingξ=ak(w)in (3.5) and dividing byukp1, one gets

Ω

|∇vk|p2vkw

Ω

m(vk+)p1n(vk)p1 w

Ω

|∇uk|p εk uk

1+ uk w

ukp

Ω

m(vk+)p1n(vk)p1 w

vk

.

This implies thatv0is a solution of

pv0=c

m(v+0)p1n(v0)p1

inΩ, ∂v0

∂ν =0 on∂Ω, (3.6)

wherecis the level appearing in the (PSC)csequence. Sincev0cst, the right-hand side of (3.6) is≡0, and since c >0, one getsm(v0+)p1n(v0)p1≡0. This relation with a nonzero constantv0impliesm≡0 orn≡0, which contradicts (2.1).

(iii) Let us finally consider the case where

Ωm=0,

Ωn=0, and letukMm,n be a (PS)c sequence forA˜ withc0. We will show thatuk remains bounded so that part (i) applies and yields the conclusion. Assume that for a subsequenceuk → +∞. For a further subsequence one obtains as above thatvkv0 inW1,p(Ω)withv0

a nonzero constant. ButBm,n(uk)=1 and so, dividing byukpand going to the limit, one obtains

Ω

m(v0+)p+n(v0)p

=0.

This is a contradiction sincev0is a nonzero constant and

Ωm=0,

Ωn=0. 2

We now turn to the geometry ofA. The situation here is again simpler in the nonsingular case where the following˜ proposition applies.

(6)

Proposition 3.5.If

Ωm=0, thenϕmMm,n is a strict local minimum ofA, with in addition for some˜ ε0>0and all0< ε < ε0,

A(ϕ˜ m)=λ(m) <infA(u):˜ uMm,n∂B(ϕm, ε) , (3.7) whereB(ϕm, ε)denotes the ball inW1,p(Ω)of centerϕmand radiusε. Similar conclusion forϕnif

Ωn=0.

Proof. We only sketch it since it is adapted from [2]. One first shows that for someε0>0,

A(ϕ˜ m) <A(u)˜ ∀uMm,nB(ϕm, ε0), u=ϕm. (3.8) To prove (3.8) one distinguishes two cases: (i)λ(m)=0 or (ii)λ(m) >0. In case (i) one choosesε0 such that Mm,nB(ϕm, ε0) only containsϕm as constant function. This clearly implies (3.8). In case (ii) one assumes by contradiction the existence of a sequenceukMm,nwithuk=ϕm, ukϕminW1,p(Ω)andA(u˜ k(m). One then deduces, as on p. 585 of [2], thatuk changes sign forksufficiently large. One also has

λ(m)

Ω

m(u+k)p+n(uk)p

=λ(m)A(u˜ k(m)

Ω

m(u+k)p+

Ω

|∇uk|p

and consequently λ(m)

Ω

n+(uk)pλ(m)

Ω

n(uk)p

Ω

|∇uk|p.

Sinceukϕm,|uk >0| →0 where|uk >0|denotes the measure of the set whereuk is>0. The desired contra- diction then follows from Lemma 3.6 below. Thus (3.8) is proved.

The fact that (3.8) implies (3.7) follows from Lemma 6 in [2], after observing that it suffices in this lemma that the functional satisfies (PS) along bounded sequences, a property which holds here by Proposition 3.3. This concludes the proof of Proposition 3.5 when

Ωm=0. Similar arguments when

Ωn=0. 2

Lemma 3.6.LetvkW1,p(Ω)withvk0,vk≡0and|vk>0| →0. Letnkbe bounded inLr(Ω). Then

Ω

nkvkp

Ω

|∇vk|p→0.

Proof. Without loss of generality, one can assumevk =1. So for a subsequence,vk vinW1,p(Ω)andvkv inLp(Ω). The assumption on|vk>0|impliesv≡0 and consequently

Ω|∇vk|p→1. The conclusion then follows since, by Hölder inequality,

Ωnkvpk →0. 2

In the singular case, one at least of the two local minima provided by Proposition 3.5 is missing. The search for suitable endpoints of paths which allow the application of a mountain pass argument will be based on the following lemmas (see in particular Lemma 3.10).

Lemma 3.7.Inequality(3.3)holds.

Proof. The inequalityeasily follows from the definition ofλ(m)andλ(n). Indeed for anyγΓ,γ (1)belongs toMm,n, is0 and so satisfies the constraint in the definition (2.2) ofλ(m). Consequentlyc(m, n)λ(m), and a similar argument applies toλ(n). To prove the strict inequality assume by contradiction that for instanceλ(m)= c(m, n). So, there exists a sequenceγkΓ such that

tmax∈[0,1]A˜ γk(t )

λ(m). (3.9)

Putuk:=γk(1). Sinceuk0, one has λ(m)

Ω

|∇uk|p max

t∈[0,1]A˜ γk(t )

λ(m), (3.10)

(7)

and consequently

Ω|∇uk|pλ(m). Let us now distinguish two cases: either (i)ukremains bounded or (ii) for a subsequenceuk → ∞.

In case (i), for a subsequence and for someu0W1,p(Ω), one has thatuk u0inW1,p(Ω). Sinceuk0, one

has

Ω

m|u0|p=1, (3.11)

and so λ(m)

Ω

|∇u0|plim inf

Ω

|∇uk|p=λ(m),

which implies that

Ω|∇u0|p=λ(m). Consequently uku0inW1,p(Ω). If

Ωm=0, then λ(m)=0 and so u0cst, which leads to a contradiction with (3.11). So

Ωm=0 and we conclude thatu0=ϕm. Let us now choose ε >0 such that (3.7) holds andB(ϕm, ε)does not contain any functionvwithv0, which is clearly possible. Fork sufficiently largeuk=γk(1)B(ϕm, ε), whileγk(0) /B(ϕm, ε)sinceγk(0)0. It follows that the pathγkintersects

∂B(ϕm, ε)and consequently max

t∈[0,1]A˜ γk(t )

infA(u):˜ uMm,n∂B(ϕm, ε) > λ(m).

This contradicts (3.9).

In case (ii) we put vk =uk/uk. For a subsequence and some v0W1,p(Ω), vk v0 in W1,p(Ω). Since

Ω|∇uk|premains bounded, we obtain

Ω|∇vk|p→0 and sov0cst; alsov0≡0 sincevk =1 impliesv0 =1.

Moreover

Ωm|v0|p=0 since

Ωm|uk|p=1. We have reached a contradiction if

Ωm=0. So let us assume from now on that

Ωm=0. We first observe that for anyγΓ there existst0=t0(γ )∈ [0,1]such that

Ω

m

γ (t0)+p

=

Ω

n

γ (t0)p

=1

2. (3.12)

Consider nowwk:=γk(t0k)). We have now instead of (3.10) 0

Ω

|∇wk|p max

t∈[0,1]A˜ γk(t )

λ(m)=0. (3.13)

We again distinguish two cases: eitherwk remains bounded, or for a subsequencewk → ∞. In the first case, for a subsequence and somew0W1,p(Ω),wk w0inW1,p(Ω). It follows from (3.13) thatw0cst and that wkw0inW1,p(Ω). A contradiction then follows from

Ω

m(w+0)p=

Ω

n(w0)p=1 2.

In the second case we put zk :=wk/wk. For a subsequence and some z0W1,p(Ω), zk z0 in W1,p(Ω). It follows from (3.13) thatz0cst and thatzkz0inW1,p(Ω); consequentlyz0 =1. Ifz0>0 then|wk<0| =

|zk<0| →0; moreoverwkchanges sign and by (3.12)

Ωn+|wk|p

Ω|∇wk|p 1/2

Ω|∇wk|p → +∞.

This yields a contradiction with Lemma 3.6. A similar argument applies ifz0<0. 2 Lemma 3.8.For anyd >0, the set

O:=

uMm,n: u0andA(u) < d˜

is arcwise connected. Similar conclusion ifu0is replaced byu0.

(8)

Note that by the definition ofc(m, n),{uMm,n:A(u) < d˜ }is not arcwise connected when max{λ(m), λ(n)}<

d < c(m, n).

Proof of Lemma 3.8. SinceOis empty if(m), we can assume from now ond > λ(m). We first consider the case where

Ωm=0. Using Lemma 3.9 below, one constructs a weightnˆ∈Lr(Ω)such thatnˆ+≡0,nˆm,

Ωn <ˆ 0 andλ(n) > d. Whenˆ m≡0, it suffices in this construction to take nˆ=εm+mwithε >0 sufficiently small;

whenm=0 i.e.m0, it suffices to takenˆ=εmB withεsufficiently small andksufficiently large, whereχB

is the characteristic function of a ballsuch thatm+≡0 onΩ\B. We then consider the manifoldMm,nˆ and the sublevel set

O:=

uMm,nˆ: A(u) < d .

By part (iii) of Proposition 3.3, the restrictionAˆofAtoMm,nˆsatisfies (PS)c for allc0. Lemma 14 from [2] then implies that any (nonempty) component ofOcontains a critical point ofA. But the first two critical levelsˆ λ(m), λ(n)ˆ ofAˆverifyλ(m) < d < λ(n), and consequentlyAˆadmits only one critical point inO. We can conclude in this way thatOis arcwise connected.

Let nowu1, u2O. Since they are0, they also belong toO. Letγ be a path inOfromu1tou2and consider the path

γ1(t ):= |γ (t )| (

Ωm|γ (t )|p)1/p. By the choice ofn,ˆ

Ω

mγ (t )p

Ω

m γ (t )+p

+ ˆn

γ (t )p

=1, (3.14)

and consequently γ1is a well defined path inMm,n, which clearly goes from u1tou2and is made of nonnegative functions. Moreover, by (3.14),

A γ1(t )

= A(γ (t ))

Ωm|γ (t )|p A γ (t )

< d for allt, and we conclude that the pathγ1lies inO.

Consider now the case where

Ωm=0. Let u1, u2O. One starts by decreasing a little bit the weightminto a weightmˆ ∈Lr(Ω)such thatmˆ m,

Ωm <ˆ 0,

Ωmuˆ p1>0,

Ωmuˆ p2>0 and

Ω|∇u1|p

Ωmuˆ p1 < d,

Ω|∇u2|p

Ωmuˆ p2 < d, which is clearly possible sinceλ(m) < d. Put

v1:= u1 (

Ωmuˆ p1)1/p and v2:= u2 (

Ωmuˆ p2)1/p.

By the first part of this proof, there exists a pathγinMm,ˆmˆ which goes fromv1tov2, is made of nonnegative functions and is such thatA(γ (t )) < dfor allt. Consider now the path

γ1(t ):= γ (t ) (

Ωm|γ (t )|p)1/p. By the choice ofm,ˆ

Ω

mγ (t )p

Ω

ˆ

mγ (t )p=1, (3.15)

and consequently γ1is a well defined path inMm,n, which clearly goes from u1tou2and is made of nonnegative functions. Moreover, by (3.15),

A γ1(t )

= A(γ (t ))

Ωm|γ (t )|p A γ (t )

< d

(9)

for allt. This concludes the proof of Lemma 3.8 forOwithu0. Similar argument in the caseu0. 2 Lemma 3.9.LetmkLr(Ω)withm+k ≡0andmkminLr(Ω)wherem0,m≡0. Thenλ(mk)→ +∞.

Proof. Suppose by contradiction that for a subsequence,λ(mk)λ <+∞. Letϕk be the positive eigenfunction associated toλ(mk)and normalized byϕkpr =1, where · qdenotes theLq(Ω)norm. One has

Ω

|∇ϕk|p=λ(mk)

Ω

mkϕkpλ(mk)m+kr.

It follows that for a subsequence, ϕk ϕ in W1,p(Ω), with ϕpr =1. Moreover, by the above inequality,

Ω|∇ϕk|p→0, which implies ϕcst=0 (call itA) and ϕkϕ inW1,p(Ω). Consequently, for k sufficiently large so that

Ωmk<0, one has 0< 1

λ(m)

Ω

|∇ϕk|p=

Ω

mkϕpkAp

Ω

m <0,

a contradiction. 2

Lemma 3.10.There existsu10andu20inMm,nsuch thatA(u˜ 1) < c(m, n)andA(u˜ 2) < c(m, n). Moreover, for any such choice ofu1, u2, one has

c(m, n)= inf

γΓ

umaxγ[0,1]A(u)˜ (3.16)

where Γ :=

γC

[0,1], Mm,n

: γ (0)=u2andγ (1)=u1 . Proof. If

Ωm=0, one takesu1=ϕmand the inequalityA(u˜ 1) < c(m, n)follows from Lemma 3.7. Similarly with u2= −ϕnin case

Ωn=0. If now

Ωm=0, one takesu1=vk forksufficiently large, wherevk is defined in (2.3).

IndeedA(v˜ k)→0 and by Lemma 3.7, 0< c(m, n), so thatA(v˜ k) < c(m, n)forksufficiently large. Similar argument for the choice ofu2in case

Ωn=0.

It remains to prove (3.16). Callc¯the right-hand side of (3.16). One clearly hasc(m, n)c. To prove the converse¯ inequality, letε >0 and takeγεΓ such that

umaxγε[0,1]A(u) < c(m, n)˜ +ε.

By Lemma 3.8 there exits a pathη1inMm,njoiningγε(1)andu1, made of nonnegative functions, and such that

umaxη1[0,1]A(u) < c(m, n)˜ +ε.

Similarly there exists a pathη2inMm,njoiningγε(0)andu2, made of nonpositive functions, and such that

umaxη2[0,1]A(u) < c(m, n)˜ +ε.

Gluing togetherη2,γε andη1, one gets a path inMm,n joiningu2 andu1, and such thatA˜ remains< c(m, n)+ε along this path. This impliesc < c(m, n)¯ +ε. Sinceε >0 is arbitrary, the conclusion follows. 2

We are now ready to give the

Proof of Theorem 3.2. Inequality (3.3) was established in Lemma 3.7. To prove thatc(m, n)is an eigenvalue, we picku1, u2as in Lemma 3.10 and we will show thatc, the right-hand side of (3.16), is a critical value of¯ A. If˜

Ωm=0 and

Ωn=0, thenA˜satisfies (PS)cfor allc0 and the classical mountain pass theorem for aC1functional on aC1 manifold (cf. e.g. Proposition 4 from [2]) yields the conclusion. If either

Ωm=0 or

Ωn=0, then we only know thatA˜satisfies (PSC)cfor allc >0. It is then Theorem 4.1 from the following section which yields the conclusion.

(10)

It remains to show that there is no eigenvalue between max{λ(m), λ(n)}andc(m, n). Assume by contradiction the existence of such an eigenvalue λand letu be the corresponding nontrivial solution of (1.2). We know thatu changes sign (sinceλ >max{λ(m), λ(n)}); moreover

0<

Ω

|∇u+|p=λ

Ω

m(u+)p, 0<

Ω

|∇u|p=λ

Ω

n up

,

and we can normalizeuso thatuMm,n. The functions u1:= u+

(

Ωm(u+)p)1/p, u2:= −u (

Ωn(u)p)1/p

belongs toMm,n, withu10,u20. We will construct a pathγinMm,njoiningu1andu2, and such thatA˜remains equal toλ along that path. This will give a contradiction with the definition ofc(m, n). To constructγ we first go fromu1touby the path

γ1(t ):= u+t u (Bm,n(u+t u))1/p and then fromutou2by the path

γ2(t ):= t u+u (Bm,n(t u+u))1/p.

It is easily verified that the path constructed in this way is well defined and satisfies all the required conditions. 2 Remark 3.11.Reproducing the end of the above proof withλreplaced byc(m, n), we conclude that the infimum in (3.2) is achieved.

4. A mountain pass theorem

Our purpose in this section is to derive a mountain pass theorem for aC1functional on aC1manifold and which satisfies the (PSC) condition.

We put ourselves in the general setting of the end of Section 2: Eis a real Banach space,gC1(E,R),M:=

{uE: g(u)=1}with 1 a regular value ofg,fC1(E,R),f˜the restriction off toM. The spaceEin this section is assumed to be uniformly convex.

Theorem 4.1. LetK be a compact metric space,K0K, andh0C(K0, M). Consider the family of extensions ofh0:

H:=

hC(K, M): h|K

0 =h0 .

AssumeHnonempty as well as the following condition:

maxtK0f h0(t )

<max

tKf h(t ) for anyhH. Define

c:= inf

hHmax

tKf h(t )

. (4.1)

Assume thatf˜satisfies(PSC)cforcgiven in(4.1). Thencis a critical value off˜. Typically, as in the application in Section 3,K= [0,1]andK0= {0,1}.

Proof of Theorem 4.1. Arguing as in the proof of Theorem 2.1 in [5] but using the strong form of Ekeland variational principle (cf. [8,7]) instead of the usual one, one obtains that ifhHandε >0 are such that

maxtKf h(t )

< c+ε

2, (4.2)

(11)

then, for eachμ >0, there existsuμMwith cf (uμ)c+ε

2, dist

uμ, h(K)

μ,

f˜(uμ)

ε μ.

We let ε= 1k and pickh=hk such that (4.2) holds, which is possible by the definition (4.1) of c. We also take μ=μk=1+ hk, where · denotes theC(K, E)norm. So there existsukMsuch that

cf (uk)c+ 1 2k, dist

uk, hk(K)

1+ hk, (4.3)

f˜(uk)

k

1+ hk1

. (4.4)

It follows from (4.3) that ukEdist

uk, hk(K)

+ hk1+2hk and so 1+ ukE2(1+ hk). Replacing in (4.4) gives

f˜(uk)

2k

1+ ukE

1

.

Thusukis a (PSC)csequence, and the conclusion follows. 2

The following additional information will be used later (cf. Proposition 2.3 in [5]).

Proposition 4.2.LetK, K0, h0,Handcbe as in Theorem4.1and lethHsatisfy maxtKf

h(t )

=c.

Thenh(K)contains a critical point off˜at levelc.

The strong form of Ekeland variational principle can also be used in a way rather similar to the above to derive Proposition 4.3.Assumef˜bounded from below and letc:=inf{f (u): uM}. Thenf˜satisfies(PSC)cif and only if f˜satisfies(PS)c.

Proof. It clearly suffices to prove that (PSC)c implies (PS)c. Let ukM satisfy cf (uk)c+1/k with ˜f (uk)→0 anduk → ∞ (ifuk remains bounded thenuk is a (PSC)c sequence and the conclusion fol- lows immediately). Using the strong form of Ekeland variational principle, we obtain for anyk and anyμ >0 the existence ofvμMwith

cf (vμ)c+ 1 2k, vμukμ, f˜(vμ)

1 kμ.

We takeμ= uk/2 and we writevk instead ofvμ. We have 1

2ukvk3 2uk, 1+ vkf˜(vk)

1 k

2 uk

1+3

2uk

cst

k , (4.5)

and so, by (PSC)c,vkhas a subsequencevnkwhich converges. Combining with (4.5) leads to a contradiction with the fact thatuk → ∞. 2

(12)

5. Some properties ofc(m, n)

We briefly study here the dependence ofc(m, n)with respect tom, n. All the weights in this section are assume to belong toLr(Ω)and to satisfy (2.1). The results and proofs below are similar to those of Section 4 in [2].

Proposition 5.1.If(mk, nk)(m, n)inLr(Ω)×Lr(Ω), thenc(mk, nk)c(m, n).

Proof. It is easily adapted from that of Proposition 22 in [2]. One successively proves upper and lower semicontinuity.

In the latter part it is convenient here to normalizeukso thatukpr =1. 2 Proposition 5.2.Ifmm˜ andnn˜a.e., then

c(m,˜ n)˜ c(m, n).

If in addition

Ω

(m˜−m)(u+)p+

Ω

(n˜−n)(u)p>0

for at least one eigenfunctionuassociated toc(m, n), thenc(m,˜ n) < c(m, n).˜

Proof. It is easily adapted from that of Propositions 23 and 25 of [2]. Proposition 4.2 is used to derive the strict monotonicity. 2

Finally let us observe that c(m, n)is homogeneous of degree−1. Some sort of separate sub-homogeneity also holds, which will be used later:

Proposition 5.3.If0< s <s, thenˆ

c(sm, n) < c(sm, n)ˆ and c(m,sn) < c(m, sn).ˆ

Proof. It is easily adapted from that of Proposition 31 of [2]. Again Proposition 4.2 is used here. 2 6. Fuˇcik spectrum

Let m, nLr(Ω) withr as before. Unless otherwise stated, we also assume (2.1). The Fuˇcik spectrum Σ= Σ (m, n) clearly contains the lines {0} ×R,R× {0}, R× {λ(n)}, {λ(m)} ×R and also possibly the linesR× {−λ(n)}and{−λ(m)} ×R. It will be convenient to denote byΣ=Σ(m, n)the setΣ (m, n)without these 2, 3 or 4 lines.

We start by looking at the part ofΣwhich lies inR+×R+. From the properties ofλ(m), λ(n)follows that if (α, β)Σ(R+×R+), thenα > λ(m)andβ > λ(n).

Theorem 6.1.For anys >0, the lineβ=sαin the(α, β)plane intersectsΣ(R+×R+). Moreover the first point in this intersection is given byα(s)=c(m, sn),β(s)=sα(s), wherec(·,·)is defined in(3.2).

Proof. An easy consequence of Theorem 3.2. 2

Lettings >0 varying, we thus get a first curveC:= {(α(s), β(s)): s >0}inΣ(R+×R+).

Proposition 6.2.The functionsα(s)andβ(s)in Theorem6.1are continuous. Moreoverα(s)is strictly decreasing andβ(s)is strictly increasing. One also has thatα(s)→ +∞ass→0andβ(s)→ +∞ass→ +∞.

Proof. The first two statements are direct consequences of the results of Section 5. The last one easily follows from Lemma 6.3 below. 2

Références

Documents relatifs

Because of significant differences between urban and rural women in such characteristics as level of education, number of children ever born and number of living children,

Wheeden, Some weighted norm inequalities for the Fourier transform of functions with vanishing moments, Trans.. Sagher, Integrability conditions for the Fourier

We would like to analyze the relations between the nodal domains of the eigenfunctions of the Dirichlet Laplacians and the partitions by k open sets D i which are minimal in the

Black level clamping, often referred to as DC restoration is accomplished by applying a back porch clamp signal to the clamp gate input pin (pin 14).. The clamp comparator is

They had come to Tanna for a 6-month period of service as part of the Victoria-Vanuatu Physicians Project (Viva), which is made up of a group of family physicians and their

Indeed, we prove that on any given Riemannian manifold, there exists families of densities such that the associated Cheeger constants are as small as desired while the

In this article, we continue this tradition of tabulating automorphic forms on Shimura varieties by computing extensive tables of Hilbert modular cusp forms of parallel weight 2

T is the address of the current subterm (which is not closed, in general) : it is, therefore, an instruction pointer which runs along the term to be performed ; E is the