• Aucun résultat trouvé

Analysis of boundary bubbling solutions for an anisotropic Emden-Fowler equation

N/A
N/A
Protected

Academic year: 2022

Partager "Analysis of boundary bubbling solutions for an anisotropic Emden-Fowler equation"

Copied!
23
0
0

Texte intégral

(1)

www.elsevier.com/locate/anihpc

Analysis of boundary bubbling solutions for an anisotropic Emden–Fowler equation

Juncheng Wei

a

, Dong Ye

b,

, Feng Zhou

c

aDepartment of Mathematics, The Chinese University of Hong Kong, Shatin, Hong Kong bDépartement de Mathématiques, UMR 8088, Université de Cergy-Pontoise, 95302 Cergy-Pontoise, France

cDepartment of Mathematics, East China Normal University, Shanghai 200062, China Received 31 October 2006; received in revised form 31 January 2007; accepted 1 February 2007

Available online 17 July 2007

Abstract

We consider the following anisotropic Emden–Fowler equation

∇(a(x)∇u)+ε2a(x)eu=0 inΩ, u=0 on∂Ω,

whereΩ⊂R2is a smooth bounded domain andais a positive smooth function. We study here the phenomenon of boundary bubbling solutions whichdo not existfor the isotropic caseaconstant. We determine the localization and asymptotic behavior of the boundary bubbles, and construct some boundary bubbling solutions. In particular, we prove that ifx¯∈∂Ω is a strict local minimum point ofa, there exists a family of solutions such thatε2a(x)eudxtends to 8π a(x)δ¯ x¯inD(R2)asε→0. This result will enable us to get a new family of solutions for the isotropic problemu+ε2eu=0 in rotational torus of dimensionN3.

©2007 Elsevier Masson SAS. All rights reserved.

Keywords:Boundary bubble; Blow-up analysis; Localized energy method

1. Introduction

The classical Emden–Fowler equation, or Gelfand equation

u+ε2eu=0 inΩ⊂RN, u=0 on∂Ω (1)

has motivated a lot of studies, because it has both geometrical and physical background. WhenN=2, (1) or more generally Eq. (2) below relates to the geometric problem of Riemannian surfaces with prescribed Gaussian curvature (see [7] and references therein). ForN3, it arises in the theory of thermionic emission, isothermal gas sphere, gas combustion. It is also considered in relation with Onsager’s formulation in statistical mechanics, the Keller–Segel system of chemotaxis, Chern–Simon–Higgs gauge theory and many other physical applications (see [4–6,11,13,19, 17,25] and the references therein).

It is well known that there exists a critical value ε>0 such that whenε > ε, no solution of (1) exists while forε(0, ε), we have a family of minimal solutions which tend uniformly to zero asε→0. WhenN=2, for any

* Corresponding author.

E-mail addresses:wei@math.cuhk.edu.hk (J. Wei), dong.ye@u-cergy.fr, ye@math.u-cergy.fr (D. Ye), fzhou@math.ecnu.edu.cn (F. Zhou).

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

doi:10.1016/j.anihpc.2007.02.001

(2)

ε(0, ε), we have also a second solution which is nonstable and blows up asε→0. The asymptotic behavior of nonstable solutions to (1), or to a more general equation

u+ε2k(x)eu=0 inΩ⊂R2, u=0 on∂Ω (2)

wherek(x)is a positive smooth function has been studied in [3,14,15,18,21,27]. LetGD denote the standard Green’s function of−with Dirichlet boundary condition andHDdenote the regular part ofGD, i.e.

HD(x, y)=GD(x, y)+ 1

2πlog|xy|. (3)

Ifuεis a family of solutions to (2) satisfying Tε=ε2

Ω

k(x)euεdx,

asε→0 and limε0 uε L(Ω)= ∞, then up to a subsequence, there holds either= ∞,uε→ ∞for allxΩ; or =8π m,m∈Nanduεmakesmpointssimpleblow-up onS= {x1, . . . , xm} ⊂Ω such that

ε2k(x)euεdx→8π m j=1

δxj, uε→8π m j=1

GD(·, xj) inClock \S),k∈N, where(x1, . . . , xm)is a critical point of defined by

(x)= m j=1

HD(xj, xj)+

i=j

GD(xi, xj)+2 m j=1

logk(xj).

Conversely, many authors have constructed blow-up solutions, see for example [2,9,10,20]. So the solutions of Eq. (1) or (2) are now well understood in dimension two.

Here we consider the following generalized Emden–Fowler equation

au+ε2eu=0 inΩ⊂R2, u=0 on∂Ω (4)

whereΩis a smooth bounded domain,ais the operator au= 1

a(x)

a(x)u

=u+ ∇logau anda(x)is a smooth function overΩsatisfying

0< a1a(x)a2<+∞. (5)

Our motivation is due to the fact that few is known for Eq. (1) in dimensionN3. As far as we know, the only explicit results in higher dimensions concern the radial solutions in spheres (see [11,13]) or in annuli (see [22]). It is worth to mention that Pacard proved in [23] (see also [16]) that for annuli, i.e.Ω=Ar0 = {x∈RN, r0< x <1}, there exists X(0,1)of measure equal to 1 such that for allr0X, there are infinitely many symmetry breaking points with bifurcation from the branch of radial solutions. Unfortunately, we do not have no more precise information about the behavior of these nonradial solutions. However, through these results, we observe already a quite different situation with the case in dimension two. Our idea here is to consider axially symmetric solutions of (1) in a torus, and try to give some precise descriptions of them. In fact, letTbe a standardN-dimensional torus (N3), i.e.

T=

x=(xi)∈RN;

x12+ · · · +x2N1−1 2+xN2 r02

(6) with 0< r0<1. If we look for solutions of (1) in the form ofu(x)=u(r, s)where

r=

x12+ · · · +xN21 and s=xN,

a direct calculus yields that the problem (1) is transformed to

(rN2u)+ε2rN2eu=0 inΩT, u=0 on∂ΩT, (7)

(3)

whereΩT= {(r, s)∈R2;(r−1)2+s2< r02}. This is just Eq. (4) with a(r, s)=rN2, so problem (4) represents a special case of (1) in higher dimension.

Eq. (4) seems to be similar to (2) or (1). We can show the existence of critical valueε>0 (depending onaandΩ), the existence of minimal solution and nonstable solution for allε(0, ε). But the structure of nonstable solutions is quite different. In [28], the authors studied the asymptotic behavior of bubbling solutions to (4), they proved that if Tε=O(1), then eitheruε→0 uniformly on any compact subset ofΩ, or there exists a finite setS= {xi} ⊂Ω and mi∈Nsuch thatuεuweakly inW1,p(Ω)for anyp(1,2), whereuverifies

au+8π

i

miδxi=0 inΩ, u=0 on∂Ω. (8)

Moreover,mi∈Nand eachximust be acritical pointofa. Recently, we have constructed in [26] bubbling solutions near anytopologically nontrivial critical pointofainΩ. In particular, near anyinterior strict local maximumofa, we have solutions with arbitrary given number of bubbles, which illustrates again the contrast with the isotropic situation.

Nevertheless, if we look at Eq. (7), the functiona(r, s)=rN2has no critical point inΩT. Consequently the blow- up cannot occur in the interior of the domain, it must appear near the boundary. A natural question is to understand these boundary bubbling solutions, which is just the aim of this paper.

First, we show the localization and asymptotic behavior of boundary blow-up whenahas no critical point inΩ. Theorem 1.1.Suppose that the anisotropic coefficientahas no critical point inΩ. Letuε be a family of solutions to problem(4)satisfying

Tε=ε2

Ω

euεdx <and max

Ω

uε→ ∞

asεtends to0. Then=8π mwithm∈Nand up to a subsequence, there exists a finite setS= {x1, . . . , xq} ⊂∂Ω andm1, . . . , mq∈Nsuch that

ε2euεχΩdx

1jq

8π mjδxj inD(R2).

Moreover, the tangential derivative∂τa(xj)=0for any1jpand uε→0 inClock \S),k∈N.

Remark 1.2.We should emphasize that the boundary blow-up phenomenon does not exist for the isotropic case, or more generally whenais constant in a neighborhood of the boundary. In that case, solutions of Eq. (4) are decreasing with respect tod(x, ∂Ω)in a fixed neighborhood of∂Ω, by moving plane argument as showed in [21] (see also [18]).

Remark 1.3.We can combine Theorem 1.1 with the result in [28] to get a more general conclusion forasatisfying just (5), see Proposition 3.4.

The second part of the paper concerns the existence of boundary bubbling solutions. Let us introduce some nota- tions. LetG(x, y)be the Green’s function associated to−a, that is, for anyyΩ,

aG(x, y)+8π δy=0 inΩ and G(x, y)=0 ifx∂Ω. (9)

DefineHto be the regular part ofG(x, y)as

H (x, y)=G(x, y)+4 log|xy|. (10)

Then our main results for this part can be stated as follows.

Theorem 1.4.Letx¯∈∂Ωbe a local minimum point ofaon∂Ω, i.e.

δ >0 such that a(x) < a(y),¯ ∀yBδ(x)¯ ∩Ω, y= ¯x.

(4)

We assume also νa(x) <¯ 0. Then forε >0 sufficiently small, problem (4) has a family of solutionsuε such that ε2euεχΩdx→8π δx¯inD(R2). More precisely, we have

uε(x)=log 1

2μ2ε+ |xξε|2)2+H (x, ξε)+o(1) inΩ (11) whereξεεsatisfy

ξε→ ¯x, (.ξε, ∂Ω)∼ 1

|logε| and με∼ 1

|logε|2 asε→0. (12)

Here we use the symbolfgto mean the existence ofC >0 such that 1

C lim inf

ε0

f

g lim sup

ε0

f g C.

Throughout the work, the symbol Cdenotes always a positive constant independent ofε, it could be changed from one line to another.

The following result shows the reason why we construct the bubbling solutions near a local minimum point ofa on the boundary, but not near a maximum point.

Theorem 1.5.Assume thatTε=O(1)andx¯∈∂Ωis a nondegenerate local maximum point ofa, thenx /¯∈S. If we return to the original equation (1) overT, we get then a family of solutions which blows up at a(N−2)- dimensional submanifold onT.

Theorem 1.6.LetTbe the torus defined by(6), then there exists a family of solutionsuεfor(1)such that asε→0,

εlim0

T

ε2euεdx=0 (13)

anduεblows up exactly on ST=

(xi)∈T;

x12+ · · · +xN21=1−r0, xN=0 .

The paper is organized as follows. First we consider the behavior of boundary bubbles, we prove Theorems 1.1 and 1.5 by potential analysis, Pohozaev identity, combined with blow-up arguments used in [15] and [28]. Then we prove Theorem 1.4 via thelocalized energy method, a combination of Liapunov–Schmidt reduction method and variational techniques similar to our previous paper [26]. The difficulties for proving all these results come from the fact that the distance between the bubbles and the boundary will tend to zero, therefore some refinements, in particular some precise informations for the Green’s function and the corresponding Robin function need to be developed to make our approach successful. Theorem 1.6 can be shown as a direct consequence of Theorems 1.5 and 1.4.

2. Behavior of the Green’s functionG(x, y)near∂Ω

In order to study the bubbles of (4) which tend to the boundary, we need to have a good understanding of the Green’s functionG(x, y)associated to−a, whenyis near∂Ω, and its regular partH, especially the corresponding Robin functionxHR(x)=H (x, x).

Lemma 2.1.There exist positive constantsd0andC such that for anyxΩ,d(x, ∂Ω)d0, there exists a unique pointxν∂Ωsatisfyingd(x, ∂Ω)= |xxν|and ifx=2xνxdenotes the reflection point ofxwith respect to∂Ω, then G(y, x)+4 log|xy| −4 log|xy|C,yΩ. (14) Moreover,

d(x,∂Ω)lim0

G(y, x)−4 log|xy|

|xy|

L(Ω)

=0. (15)

(5)

Proof. ConsiderL(y, x)=G(y, x)+4 log|xy| −4 log|xy|. We have

a(y)L(y, x)= −4∇loga(y)·

yx

|yx|2yx

|yx|2

inΩ.

It is clear that the right-hand side is uniformly bounded inLp(Ω)for anyp∈ [1,2). Furthermore, by the regularity of the domainΩ, we know that L(·, x) L(∂Ω)=O(d(x, ∂Ω))whend(x, ∂Ω)d0. The standard elliptic theory implies then L(·, x) L(Ω)=O(1).

More precisely, remark that for 1p <2, a(y)L(y, x) Lp(Ω)tends to 0 as|xx| →0. So L(·, x) L(Ω)→0 asd(x, ∂Ω)tends to 0. 2

Recall the following expansion ofxH (x, y)proved in [26].

Lemma 2.2. Let Hy(x)=H (x, y) for any yΩ. Then yHy is a continuous map from Ω into C0,γ(Ω),

γ(0,1). LetHD be the regular part of the standard Green’s function defined by(3), we have H (x, y)=8π HD(x, y)+ ∇loga(y)· ∇

|xy|2log|xy| +H1(x, y), (16) whereyH1(·, y)is a continuous map fromΩintoC1,γ(Ω)for allγ(0,1). Furthermore, the function(x, y)H1(x, y)C1×Ω), in particularxH (x, x)C1(Ω).

Using the equation satisfied byHy, we can get, for anyγ(0,1), Hy C0,γ(Ω)=O

1 d(y, ∂Ω)

uniformly inΩ. (17)

Now we show the behavior of the Robin functionxH (x, x)near the boundary.

Lemma 2.3.LetHR denote the Robin functionxH (x, x), then HR(x)=4 logd(x, ∂Ω)+O(1), ∇HR(x)=O

1 d(x, ∂Ω)

uniformly inΩ. (18)

Sketch of Proof. Using the equation ofxH (x, y), clearly H (x, y)=8π HD(x, y)+O(1)inΩ×Ω. By the behavior of HD (see for example [1]), we haveH (x, x)=4 logd(x, ∂Ω)+O(1)inΩ. For the estimate of∇HR, using the equation satisfied byHD(·, y), we obtain

HD(·, y)

C1(Ω)=O 1

d(y, ∂Ω)

foryΩ.

Consider the equation ofxH1(x, y), a(x)H1(x, y)=4

∇loga(x)− ∇loga(y)

· xy

|xy|2−8π∇loga(x)xHD(x, y)

− ∇x2

|xy|2log|xy| ·

∇loga(x),∇loga(y) inΩandH1(x, y)= −∇loga(y)· ∇x(|xy|2log|xy|)ifx∂Ω. We get then

H1(·, y)

C1(Ω)=O 1

d(y, ∂Ω)

uniformly inΩ. (19)

Moreover, for 1p <2, we can prove, by direct calculus, xy

|xy|2

W1p ,p1 (∂Ω)

=O 1

d(y, ∂Ω)

uniformly foryΩ,

hence ∇yHD(·, y) W1,p(Ω)=O(d(y, ∂Ω)1) inΩ. Checking carefully the equation satisfied byyH1(·, y), we obtain

yH1(·, y)

C0(Ω)=O 1

d(y, ∂Ω)

uniformly inΩ. (20)

(6)

AsHD(x, y)=HD(y, x)inΩ×ΩandHR(x)=8π HD(x, x)+H1(x, x), we are done. 2

Remark 2.4.Here we havea(x)G(x, y)=a(y)G(y, x)inΩ×Ω, but not the usual symmetry for the standard Green’s functionGDand thanks to the expansion (16), we see thatHyis not inC1(Ω)in general (except if∇a(y)=0). These facts make our estimate for the Robin functionHRmore involved.

3. Boundary blow-up analysis

Now we are in position to prove Theorems 1.1 and 1.5. IfTεis bounded, we know thatuεis bounded inW1,p(Ω) for anyp∈ [1,2). We get first a Brezis–Merle type result.

Lemma 3.1.LetΩ be a smooth bounded domain inR2. There existsα >0 (depending onΩ anda)such that if a solution of (4)uεsatisfies, forxΩandδ >0

Bδ(x)Ω

ε2euεdyα,

then uε L(B

δ/2(x)Ω)C.

By the results in [3], we need just to consider the situation near the boundary, i.e. whenx∂Ω. Indeed, we can use conformal transformation to changeΩ locally as a part ofR×R+(see the proof of Lemma 3.2 below), then we use a reflection argument in order to apply methods in [3] (see also [28]). We leave the detail for interested readers.

We define the blow up set foruεas follows:

Sdef=

xΩ; ∃εk→0 andxεkxsuch thatuεk(xεk)→ ∞ . Thanks to Lemma 3.1, we obtain

S=Σdef=

xΩ; ∀δ >0, lim inf

ε0

Bδ(x)Ω

ε2euεdy > α

.

Consequently, up to a subsequence, we have thatε2euεχΩdxtends to

xj∈Sηjδxj inD(R2)and #(S) <∞, since Tε=O(1). By the result in [28], if we assume thatahas no critical point inΩ, thenS∂Ω.

Lemma 3.2.For anyk∈N,uε→0inClock \S). Moreover, for anyx0S, we have∂τa(x0)=0andη0. Proof. First, by the definition of the Green’s functionG,

uε(x)= 1 8π a(x)

Ω

a(z)G(z, x)ε2euε(z)dz,xΩ. (21)

Fix any subsetKΩ such thatKS= ∅. Givenλ >0, sinceS∂Ω, applying (15) and Remark 2.4, there exists δ >0 such that G(y, x) L(K)λ, ifd(y,S)δ. Therefore, if we decomposeΩasΩ1= {yΩ;d(y,S)δ}and

Ω2 =

{yΩ;d(y,S) > δ}, for anyxK,

Ω1

a(y)G(y, x)ε2euε(y)dya2λ×

Ω1

ε2euε(y)dya2Tελ=O(λ).

On the other hand, the definition ofS implies thatuis uniformly bounded in any compact set away fromS. Hence for fixedδ >0, there existsC >0 such that

Ω2

a(y)G(y, x)ε2euε(y)dyCε2

Ω2

G(y, x) dy=O(ε2),

(7)

because G(·, x) L1(Ω)is uniformly bounded forxΩ. By (21), we get limε0 uε L(K)=0. The usual elliptic theory shows thenuε→0 inCk(K)for allk∈N.

Letx0S. Choosing a smooth open neighborhoodU ofx0such thatUS= {x0}, we can assume that there exists a conformal diffeomorphismΦfromBδintoUsatisfyingΦ(0)=x0,

Φ1(∂ΩU )= [−δ, δ] × {0} and Φ1U )=(R×R+)Bδdef=Ω0. It is not difficult to check thatvε=uεΦis a solution of

bvε=ε2|∇Φ|2evε inΩ0, withb=aΦ (22) andvε=0 onΓ0=∂Ω0(R× {0}). Takingb(x)∂1vεas a test function, sinceν1=0 andvε=0 onΓ0, we get

Ω0

∂b(x)

∂x1

|∇vε|2 2 dx=

Ω0

ε2∂(b(x)|∇Φ|2)

∂x1 (evε−1) dx−ε2

Γ

b(x)|∇Φ|2(evε−1)ν1

+

Γ

b(x)

2 |∇vε|2ν1

Γ

b(x)∂vε

∂x1

∂vε

∂ν dσ,

whereΓ =∂Ω0\Γ0. Thus all the right-hand side terms are uniformly bounded. Ifτa(x0)=0, by takingδ small enough, we can assume without loss of generality that1b(x) >0 inΩ0, because|x1b(0)| = |∇Φ(0)| × |τa(x0)| =0.

Therefore

Ω0

|∇vε|2dx=O(1),

this deduces thatuεis bounded inH1(UΩ). The Moser–Trudinger inequality shows then the boundedness ofeuε inLp(UΩ)for allp1, which yields uε L(UΩ)=O(1)by the equation, hence contradicts withx0S=Σ. Furthermore, letxεUrealize maxUΩuε(x). So limε0xε=x0. Letdε=d(xε, ∂Ω)and−2 logλε=uε(xε)+ 2 logε. We claim

εlim0

dε

λε = ∞. (23)

First, we have limε0λε =0, since otherwise ε2euε is uniformly bounded in UΩ and no blow-up will occur.

Suppose now (23) is not true, up to a subsequence, we may assume thatdεεC. Define wε(y)=uε(xε+λεy)+2 logε+2 logλε inΩε=

y∈R2;xε+λεyUΩ . Clearlywε(y)0 inΩε,

a(xε+λεy)wε=ewε inΩε and

Ωε

ewε(y)dy=ε2

UΩ

euε(x)dx=O(1).

Takingx=xεin (21) and applying Lemma 2.1 (asd(xε, ∂Ω)→0), we obtain

0= 1

8π a(xε)

Ω

a(z)G(z, xε2euε(z)dzuε(xε)

= 1

2π a(xε)

Ω

log|xεz|

|xεz|+O(1)

a(z)ε2euε(z)dzuε(xε)

= 1

2π a(xε)

UΩ

a(z)ε2euε(z)log|xεz|

|xεz|dzuε(xε)+O(1). (24)

Here, we have used the fact that |xεz|/|xεz| →1 uniformly in Ω\U, as xεx0. Let z=xε+λεy and yε=(xεxε)/λε, the equality (24) leads to

0= 1

2π a(xε)

Ωε

a(xε+λεy)ewε(y)log|yεy|

|y| dyuε(xε)+O(1). (25)

(8)

Notice that|yε|2C and ewε L1Lε)=O(1). By decomposing the last integral over domains{|y|C} ∩Ωε and{|y|C} ∩Ωε, we obtain clearly

Ωε

a(xε+λεy)ewε(y)log|yεy|

|y| dya2×

Ωε

ewε(y)log

1+2C

|y|

dy=O(1).

But limε0uε(xε)= ∞, we reach a contradiction with (25), so the claim (23) holds.

Thus,Ωεtends to the whole spaceR2and standard blow-up analysis implies that up to a subsequence,wεconverges towinCloc2 (R2)where

w=ew inR2 and

R2

ewdy <.

It is well known from [8] that

R2

ewdy=8π.

Fatou’s lemma yields thenη08π. 2 More precisely, we will quantifyη0.

Proposition 3.3.For anyx0S, we haveη0∈8πN.

Proof. Let x0S, under conformal transformation, we can assume that x0=0 and there existsδ >0 such that BδS= {0},

Ω0=Bδ(0)Ω=Bδ(R×R+) and Bδ(0)∂Ω= [−δ, δ] × {0} =Γ0. It suffices to prove that (recall thatvε=uεΦ verifies (22)),

εlim0

Ω0

ε2|∇Φ|2evεdx∈8πN. (26)

Defineζεby

ζε= ∇logb· ∇vε inΩ0, ζε=vε on∂Ω0. Thengε=vεζεsatisfies

gε=ε2Vεegε inΩ0, gε=0 on∂Ω0,

whereVε= |∇Φ|2eζε. Since vε C(∂Ω0)→0 andvεconverges to 0 weakly inW1,p0)for 1< p <2 by Lemma 3.2, soζε is a family of continuous function satisfying ζε C(Ω

0)→0, thusVε is a family of continuous functions which converges uniformly to the positive functionV = |∇Φ|2.

We will prove a Li–Shafrir type result as in [15]. Indeed, we can get a first bubble by considering maxΩ

0gε= gε(xε)as in the proof of Lemma 3.2. We claim:

If max

Ω0

gε(x)+2 logε+2 log|xxε|

=O(1),thenη0=lim

ε0

Ω0

ε2Vεegεdx=8π. (27) Recall thatdε=d(xε, ∂Ω), let

wε(y)=vε(xε+dεy)+2 logε+2 logdε

be defined inB1. Then−wε=Vε(xε+dεy)ewε. Since all the assumptions of Proposition 2 in [15] are verified by wε, we conclude then

εlim0

B(xε)

ε2Vεegεdx=lim

ε0

B1

Vε(xε+dεy)ewεdy=8π. (28)

(9)

LetDε=Ω0\Bdε(xε), we will prove gε L(Dε)=O(1). Unfortunately we do not have the sup+inf inequality as in [15], so we need some new arguments. In fact, we will use the potential analysis. Write

gε(x)=

Ω0

G0(y, x)ε2Vε(y)egε(y)dy,

where G0 is the standard Green’s function associate to − over the domain Ω0. Let xε =(xε1, xε2), thanks to Lemma 3.2, we can assume that xε1=0 by translation. By Lemma 2.1, we can also fixδ0(0, δ) small enough such that for any|x|< δ0,

2π G0(y, x)+log|xy| −log|xy|C,yΩ0.

LetxBδ0Dε. If|x|/4|xy|3|x|, as|xx|2|x|,|yx||yx| + |xx|5|x|and 2π G0(y, x) log 20+O(1); if|xy|3|x|, then|y||xy| − |x|2|x|hence|xy||y| − |x||y|/2. Now|yx|

|y|+|x|3|y|/2, we get 2π G0(y, x)log 3+O(1). SoG0(·, x)is uniformly bounded in the domainΩ0\B|x|/4(x).

It remains to consider|yx||x|/4. In this case, since|x||xxε| +dε andxDε,

|yxε||xxε| − |yx||xxε| −|x|

4 3|xxε| −dε

4 dε

2,

so|y||yxε|+|xε|3|yxε|. By the hypothesis in (27), we getg(y)+2 logε+2 log|y|C, if|yx||x|/4, yΩ0andxDε. Hence

I=

B|x|/4(x)Ω0

G0(y, x)ε2Vεegε(y)dy

B|x|/4(x)

log|yx|

|yx|

+O(1)

ε2× C ε2|y|2dy.

Using polar coordinatesy=x+ |x|re, as|yx|9|x|/4 and|y|3|x|/4 inB|x|/4(x),

IC 1/4 0

log

9 4r

+1

r dr <.

Finally, forxBδ0Dε, gε(x)=

Ω0

G0(y, x)ε2Vεegεdy

=

Ω0\B|x|/4(x)

G0(y, x)ε2Vεegεdy+

Ω0B|x|/4(x)

G0(y, x)ε2Vεegεdy

O(1)×

Ω0\B|x|/4(x)

ε2Vεegεdy+I

=O(1).

Applying Lemma 3.2, we conclude then gε L(Dε)=O(1), which implies ε2Vεegε L1(Dε)=O(ε2). Combining with (28), the claim (27) is proved.

If now maxΩ0

gε(x)+2 logε+2 log|xxε|

→ ∞, (29)

we prove similar results as Lemmas 4 and 5 in [15] by induction, that is, up to a subsequence, we can get a family {xε,k}1kmsuch that

maxΩ0

gε(x)

2 +logε+ min

1kmlog|xxε,k|

=O(1) and η0=8π m. (30)

The main idea is to compare the distance between the different bubbles and their distances to the boundary. By suitable gauge transformation, we can either transform some of them into interior bubbles and then use the result in [15]; or

(10)

split them into boundary bubbles which concentrate in different places of∂Ω and then use the induction hypothesis.

We show here just the situation withm=2 and leave the complete proof for interested readers.

Assume that (29) holds. Then by considering the point yε which realizes the maximum of gε(x)+2 logε+ 2 log|xxε| over Ω0, we get a second bubble by repeating the argument in [15] and the proof of Lemma 3.2.

Suppose that the first estimate in (30) holds withxε,1=xεandxε,2=yε. Define d˜ε=min

|xεyε|, d(xε, ∂Ω), d(yε, ∂Ω) .

Up to a subsequence and a permutation of index, we have the following possibilities:

(i) d˜ε=d(xε, ∂Ω)and|yεxε|C0d˜ε. Under translation, we can assumed˜ε= |xε|. TakeMε=(C0+1)d˜εand considerwε(x)=gε(Mεx)+2 logε+2 logMεinB4(R×R+). Up to a subsequence, we have then two interior bubbles where we may use the result in [15] to claim that

εlim0

B4MεΩ0

ε2Vεegεdx=16π. (31)

(ii) d˜ε=d(xε, ∂Ω)and|yεxε|/d˜ε→ ∞. So|yε|/|xε| tends to infinity. TakeMε= |yε| and consider wε(x)= gε(Mεx)+2 logε+2 logMεinB4(R×R+). The two bubbles are now split away, either we have one interior bubble and one boundary bubble, or we have two separated boundary bubbles. So we turn back to the simple boundary bubble case and again (31) holds.

(iii) d˜ε= |yεxε|, it suffices to take Mε = |xε| andwε as above, we transform then the bubbles to two interior bubbles as in case (i), so the result in [15] yields also (31).

Meanwhile, for all three cases, since|xxε||xyε|/2 inΩ0\B4Mε (for correspondingMε), the first estimate in (30) impliesgε(x)+2 logε+2 log|xyε| =O(1)inΩ0\B4Mε. We can show always gε L0\B4Mε)=O(1)by using the Green’s function as for (27). Now it is easy to conclude thatη0=16π. 2

Combining with the result in [28], we get

Proposition 3.4.Letuεbe a family of solutions to problem(4)satisfying Tε=ε2

Ω

euεdx=O(1) and lim

ε0max

Ω

uε= ∞.

Then up to a subsequence, uε tends tou inD(Ω), where u is the solution given by(8) with a finite set S= {xi} ⊂Ω, composed by critical points ofa. On the other hand, there exists a finite setS= {zj} ⊂∂Ω such that in D(R2),

ε2euεχΩdx

xi∈S

8π miδxi+

zj∈S

8π λjδzj. Moreovermi, λj∈N,∂τa(zj)=0for anyzjSand

uεu inClock

Ω\(SS),k∈N.

Remark 3.5.As we have mentioned earlier, whenais constant in a neighborhood of∂Ω,S= ∅.

Proof of Theorem 1.5. If it is not true, we havex¯∈S which is a nondegenerate local maximum point ofa. As in the proof of Lemma 3.2, we will take a conformal transformation to change a small neighborhood ofx¯ inΩ into Ω0=Bδ(R×R+)witha replaced byb=aΦ anduε replaced byvε=uεΦ. Moreover, 0=Φ(x)¯ becomes a nondegenerate local maximum point ofb. Choosingδ >0 small enough, we may assume thatΩ0S= {0}and x· ∇b(x)0 inΩ0. Using the Pohozaev identity (see [24]) obtained by multiplyingb(x)x· ∇vεto Eq. (22), we get (Γ =∂Ω0\(R× {0}))

(11)

Ω0

div

b(x)|∇Φ|2x

ε2evεdx=

Γ

ε2b(x)|∇Φ|2(x·ν)evε−1 2

Γ

b(x)(x·ν)|∇vε|2 +

Γ

b(x)∂vε

∂ν(x· ∇vε) dx+1 2

Ω0

(x· ∇b)|∇vε|2dx,

sincex·ν=x· ∇vε=0 over[−δ, δ] × {0}. Thanks to Lemma 3.2, we see that the first three terms in the right-hand side tend to zero asε→0 while the last term is nonpositive. However, the left-hand side tends to 16a(x)π m¯ with m∈N, this is just impossible. 2

Remark 3.6.The nondegeneracy condition can be erased if we know thatx· ∇b(x)0 near 0 inΩ0. It is worth to mention that Theorem 1.5 cannot be proved by the usual moving plane method, however we state here a special case which is a direct consequence of this method.

Proposition 3.7.Assume thata(x)a(r)forx=(r, s)Ω anda is nondecreasing with respect tor. Assume also Tε=O(1). ThenS∩ {r=rmax}is empty, wherermax=maxxΩr.

Proof. In this case, we see that Eq. (4) is in the formu+c1(x)∂1u+ε2eu=0 withc10 inΩ, thus the moving plane argument (see Theorem 2.1 in [12]) implies that for anyx0=(rmax, s)∂Ω, there exists a neighborhoodUof x0, independent ofε, such thatuε is decreasing with respect tor inU. Hence no blow up can occur atx0because Tε=O(1). 2

4. Existence of boundary bubbling solution

From now on, we will construct a family of blow up solutions stated in Theorem 1.4. We should define some approximate solutions, and choose an appropriate configuration space for the parameters. Then we need to understand the behavior of linearized operator around these approximate solutions, to get suitable functional settings and the inverse of these linearized operators. Finally, we solve the nonlinear problem and conclude by variational techniques.

This procedure has been used successfully in constructing interior bubbles for Eq. (2) in [9,10] and for Eq. (4) in [26].

Here we construct boundary bubbles. Our difficulty is to estimate precisely the distance between the bubble and the boundary.

4.1. Approximate solution

GivenξΩ andμ >0, we define u(x)=log 8μ2

2μ2+ |xξ|2)2.

The configuration space forξ is chosen as the following Λdef=

ξBδ(x)Ω; C1

|logε| (.ξ, ∂Ω) C2

|logε|

(32) whereC1, C2>0 will be determined later on andμis chosen as

log(8μ2)=HR(ξ )=H (ξ, ξ ). (33)

Using the choice ofΛand the estimate (18), there existsC >0 such that forε >0 small, 1

C|logε|2μ C

|logε|2. (34)

We get alsou(x)=log(8μ2)−4 log|xξ| +O(ε2μ)on∂Ω. The ansatz is thenU (x)=u(x)+Hε(x)whereHεis a correction term defined as the solution of

aHε+ ∇logau=0 inΩ, Hε= −u on∂Ω. (35)

Références

Documents relatifs

L’accès aux archives de la revue « Rendiconti del Seminario Matematico della Università di Padova » ( http://rendiconti.math.unipd.it/ ) implique l’accord avec les

The uniqueness result, as well as the local uniform estimates are crucial ingredients for obtaining a priori estimate, degree counting formulas and existence results for

The first ingredient of our proof is the power deterioration of all derivatives of solutions to Monge–Ampère equations with constant right-hand side (Lemma 1.1), which we prove

Vol. - In this paper we study the structure of certain level set of the Ginzburg-Landau functional which has similar topology with the configuration space. As an

MARCUS M., Large solutions of semilinear elliptic equations: existence, uniqueness and asymptotic behaviour, J. DYNKIN, A probabilistic approach to one class

Analogous boundary regularity questions for solutions of the nonpara- metric least area problem were studied by Simon [14]. 2) arises in a similar fashion as for

Global regularity for solutions of the minimal surface equation with continuous boundary values.. Annales

The proof of it is by adopting the method used in our previous paper [16] and is given in Appendix A..