• Aucun résultat trouvé

Boundary decay estimates for solutions of fourth-order elliptic equations

N/A
N/A
Protected

Academic year: 2022

Partager "Boundary decay estimates for solutions of fourth-order elliptic equations"

Copied!
23
0
0

Texte intégral

(1)

c 2007 by Institut Mittag-Leffler. All rights reserved

Boundary decay estimates for solutions of fourth-order elliptic equations

Gerassimos Barbatis

Abstract.We obtain integral boundary decay estimates for solutions of fourth-order elliptic equations on a bounded domain with regular boundary. We apply these estimates to obtain stability bounds for the corresponding eigenvalues under small perturbations of the boundary.

1. Introduction

Let Ω be a bounded region in RN and let H be a fourth-order, self-adjoint, uniformly elliptic operator onL2(Ω) subject to Dirichlet boundary conditions onΩ,

Hu(x) =

|η|≤2

|ζ|≤2

Dη(aηζ(x)Dζu(x)), x∈.

The scope of this paper is to obtain integral boundary decay estimates for solutions of the equation

Hu=f, f∈L2(Ω). (1)

More precisely, we want to establish ranges ofβ >0 for which the integrals

d−2−βu2dx and

d−β|∇u|2dx

are finite (where d(x)=dist(x, ∂Ω)). If Ω is regular in the sense that the Hardy–

Rellich inequality

(∆u)2dx≥c

|∇u|2 d2 +u2

d4

dx, u∈H02(Ω),

is valid, we then immediately have such an estimate sinceH02(Ω)=Dom(H1/2). Our aim is to establish better decay estimates that exploit the fact that the solution u of (1) belongs not only toH02(Ω) but also to Dom(H).

(2)

This problem is well studied in the case of second-order operators. In [D2]

Davies obtained boundary decay estimates of the form

|∇u|2 d + u2

d2+2α

dx≤c(Hu2H1/2u2+u22), u∈Dom(H), (2)

for α>0 in some interval (0, α0). Here α0 is an explicitly given constant which depends on the boundary regularity and the ellipticity constants of H. As an application of (2) stability estimates were obtained on the eigenvaluesn}n=1ofH under small perturbations of the boundaryΩ. More precisely, ifΩΩ is a domain such that⊂{x∈Ω:dist(x, ∂Ω)<ε}and if˜n}n=1are the corresponding Dirichlet eigenvalues (the operator H being defined by form restriction), then it was shown that (2) implies

0˜λn−λn≤cnε (3)

for all n∈N and all ε>0 small enough. This estimate has obvious applications in the numerical computation of eigenvalues; see [D2] for more on specific examples.

Inequality (3) was subsequently improved in [D3], where ε, α<α0, was re- placed by ε0, for the same α0; this was done by estimating the integrals

d(x)<ε|∇u|2dx and

d(x)<εu2dxfor small ε>0. For results analogous to those of [D3] for thep-Laplacian together with applications we refer to Fleckinger et al. [FHT].

See also [EHK] where estimates of this type were first obtained for eigenfunctions of second-order operators. For relevant results in the case of singular operators see [M].

In our main theorem we obtain integral decay estimates analogous to (2) for fourth-order operators. More precisely, for a fourth-order operator H with L coefficients we establish boundary decay estimates of the form

|∇2u|2

d +|∇u|2 d2+2α+ u2

d4+2α

dx≤c(Hu2Hα/2u2+u22), u∈Dom(H), (4)

forαin an interval (0, α0). Under additional assumptions we obtainα0=12, which is optimal. To prove (4) we first use some general inequalities, which lead to a prop- erty (Pα) being identified as sufficient for the validity of (4). We then study prop- erty (Pα), and find sufficient conditions under which it is valid; the distance function used here is taken to be the Finsler distance induced by the operator.

Technical reasons oblige us to make a regularity assumption that is not needed in the second-order case and requires d(x) to be C2 near Ω. This relates to a recurrent and largely unsolved issue in higher-order problems: a distance function is normally only once differentiable, but is required for technical reasons to be

(3)

differentiated more than once; see for example [B], where such an issue has arisen in the context of heat kernel estimates.

Finally, as an application of (4) we obtain stability bounds analogous to (3) on the eigenvalues ofH under small boundary perturbations.

The structure of the paper is as follows: in Section 2 we provide a sufficient condition (Pα) for the validity of (4); in Section 3 we establish a range of α for which (Pα) is valid for different classes of operators; and in Section 4 we present the application to eigenvalue stability.

Setting

We fix some notation. Given a multi-indexη=(η1, ..., ηN) we writeη!=η1!...ηN! and |η|=η1+...+ηN. We write γ≤η if γi≤ηi for all i, in which case we also set cηγ=η!!(η−γ)!. We use the standard notationDηu=(∂/∂x1)η1...(∂/∂xN)ηNuand (∇u)η=uηx11...uηxNN. By 2u we denote the vector (uxixj)Ni,j=1. The letter c will denote a constant whose value may change from line to line; the constants c1, c2

andc3however are the same throughout the paper.

We now describe our setting. We assume that Ω is a bounded domain inRN with boundaryΩ. We consider a distance function d(·,·) on Ω, and denote by d(·) the corresponding distance to the boundaryΩ. We say thatd(·) belongs to the classDif it satisfies:

(D1) There existc1, c2>0 such that for anyx, y∈Ω, c1≤ |∇zd(z, y)| ≤c2, z∈, (a)

c1dEuc(x, y)≤d(x, y)≤c2dEuc(x, y). (b)

(D2) There existθ, τ >0 such that

d(x) isC2 on{x∈Ω : 0< d(x)< θ}, (a)

|∇2d(x)| ≤cd(x)−1+τ on{x∈Ω : 0< d(x)< θ}.

(b)

(D3) The following Hardy–Rellich inequalities are valid for allv∈Cc2(Ω):

(∆v)2dx≥c3

v2 d4dx, (a)

(∆v)2dx≥c3

|∇v|2 d2 dx.

(b)

We note that a sufficient condition for (D3)(b) is the Hardy inequality

|∇v|2dx≥c3

v2 d2dx.

(4)

The distance d(·,·) will typically be a Finsler distance, in which case (a) and (b) of (D1) are equivalent. Condition (D2) is essentially a strong regularity assumption on Ω, as will be seen below. Its validity in examples will always involve the specific value τ=1; we choose however this more general and somewhat axiomatic setting because, we believe, it shows more clearly what the essential ingredients are. Finally, for more on Hardy–Rellich inequalities, optimal constants as well as improved versions of such inequalities we refer to [BFT] and [BT] and references therein.

In the sequel we shall often need to twice differentiated(x) near Ω. In order to avoid repeatedly splitting integrals in two, we redefine d(x) on{x∈Ω:d(x)>θ}

so that nowd(x) is a positiveC2function on Ω which equals inf{d(x, y):y∈∂}for x∈{x∈Ω:dist(x, ∂Ω)<θ}but not necessarily for allx∈Ω (of coursed(x,·) extends toΩ by uniform continuity). In relation to this we emphasize that throughout the paper what really matters is what happens near the boundary Ω. We also note that, while the validity of estimate (4) and assumption (D3) is independent of the specific distanced(·)∈Dchosen, we shall need to consider non-Euclidean distances since some of the intermediate calculations do depend on the specific choice of the distance.

We will consider operators of the form Hu(x) =

|η|=2

|ζ|=2

Dη(aηζ(x)Dζu(x)), x∈, (5)

subject to Dirichlet boundary conditions onΩ; lower-order terms can be easily ac- comodated. More precisely, we start with a matrix-valued function a(x)={aηζ(x)} which is assumed to be have entries in L(Ω) and to take its values in the set of all real, 12N(N+1)×12N(N+1) matrices (12N(N+1) is the number of multi-indices η of length|η|=2). We assume that{aηζ(x)} is symmetric for all x∈Ω and define a quadratic formQ(·) on the Sobolev spaceH02(Ω) by

Q(u) =

|η|=2

|ζ|=2

aηζ(x)Dηu(x)Dζu(x)dx, u∈H02(Ω).

We make the ellipticity assumption that there existλ,Λ>0 such that λQ0(u)≤Q(u)ΛQ0(u), u∈H02(Ω),

whereQ0(u)=

(∆u)2dxdenotes the quadratic form corresponding to the bilapla- cian ∆2. We then defineH to be the associated self-adjoint operator onL2(Ω), so that Hu, u=Q(u) for all u∈Dom(H).

(5)

2. Boundary decay Letd(·)∈D. Letα>0 be fixed and let us define

ω(x) =d(x)−α, x∈. We regularizeω defining

dn(x) =d(x)+1

n, ωn(x) =dn(x)−α, n= 1,2, ....

(6)

We note that u∈H02(Ω) impliesωnu∈H02(Ω), n∈N. It is crucial for the estimates which follow that, while they contain the functionsdnandωn, they involve constants that are independent ofn∈N.

Lemma 1. Let α>0. There exists a constantc which is independent ofn∈N such that

|∇2u|2

dn +|∇u|2 d2+2αn

+ u2 d4+2αn

dx≤cQ(ωnu), u∈H02(Ω). (7)

Proof. It suffices to prove (7) for all u∈Cc2(Ω). So let u∈Cc2(Ω) be given and letv=ωnu, a function also inCc2(Ω). Using (D3) we have

u2 d4+2αn

dx=

v2 d4n

dx≤

v2 d4dx≤c

(∆v)2dx.

Similarly,

|∇u|2 d2+2αn

dx=

1 d2+2αn

αdα−1n v∇dn+dαn∇v2dx

≤c

v2 d4ndx+c

|∇v|2 d2n dx≤c

(∆v)2dx, where we have used the fact that

|∇2v|2dx=

(∆v)2dx.

(8)

Finally, sincedanddn differ by a constant,

uxixj=dαnvxixj+αdα−1n (dxivxj+dxjvxi)+αdα−1n dxixjv+α(α−1)dα−2n dxidxjv, and therefore

|∇2u|2 dn dx≤c

|∇2v|2dx+

|∇v|2 d2n dx+

v2 d4ndx+

|∇2d|2 d2n v2dx

. Sincedn≥d, the second and third terms in the brackets are smaller thanc

(∆v)2dx by the Hardy–Rellich inequalities (D3). The same is true for the last term by (D2).

Thus, one more application of (8) concludes the proof.

(6)

Lemma 2. Let α∈(0,1) and ωn=d−αn . Then there exists a constant c>0, independent of n∈N, such that

Q(u, ωn2u)≤cHu2Hα/2u2, u∈Dom(H). Proof. For anyn∈Nandu∈Cc2(Ω) we have

ωn4/αu2dx≤

ω4/αu2dx=

u2

d4dx≤cQ(u).

Hence ω4/αn ≤cH in the quadratic form sense, which by [D1, Lemma 4.20] implies that ωn4≤cHα(sinceα∈(0,1)). Hence givenu∈Dom(H) we have

Q

u, ω2nu ≤ Hu2ωn2u

2≤cHu2Hα/2u2,

which is the stated inequality.

We can now establish a sufficient condition for the boundary decay estimates.

Theorem 3. Let α∈(0,1)be fixed and let ωn=d−αn . Assume that there exist k, k>0 independent ofn∈N such that

Q(ωnu)≤kQ

u, ωn2u +ku22, u∈Cc2(Ω), (9)

for all n∈N. Then there existsc>0 such that

|∇2u|2

d +|∇u|2 d2+2α+ u2

d4+2α

dx≤cHu2Hα/2u2, u∈Dom(H). (10)

Proof. The validity of (9) for allu∈Cc2(Ω) implies its validity for allu∈H02(Ω) and in particular for allu∈Dom(H). Hence givenu∈Dom(H) and applying Lem- mas 1 and 2 we conclude that there exists a constantcsuch that for anyn∈Nthere holds

|∇2u|2

dn +|∇u|2 d2+2αn

+ u2 d4+2αn

dx≤c

Hu2Hα/2u2+u22 .

Letting n!+, applying the dominated convergence theorem and using the fact that the spectrum of H is bounded away from zero we obtain (10).

(7)

3. The property (Pα)

The validity of assumption (9) of Theorem 3 will be our main interest in this section. For the sake of simplicity, for any α∈(0,1) we define the property (Pα) (relative to the distance functiond∈D) as

⎧⎪

⎪⎩

There exists constantsk, k>0 such that Q

d−αn u ≤kQ

u, d−2αn u +ku22 for alln∈N andu∈Cc2(Ω). (Pα)

This is precisely assumption (9) of Theorem 3. Our aim in this section is to obtain sufficient conditions under which property (Pα) is valid. In the following three sub- sections we present three theorems that provide such conditions. The first applies to all operators in the class under consideration; the second applies to operators of a specific type but gives a better range ofα>0; and the third applies to small perturbations of operators in the second class.

Remark. If Ω is smooth then the ground state φ of ∆2 decays as d(x)2 as x!∂Ω. Hence the integral in the left-hand side of (10) is not finite for α≥12. For this reason and throughout the rest of the paper we restrict our attention to α∈(0,12).

3.1. General operators

We always work in the context described at the beginning of Section 2. We recall that forα∈(0,12) we haveωn=d−αn =(d+1/n)−α; we also recall that λand Λ are the ellipticity constants of the operatorH.

Theorem 4. There exists a computable constantc>0such that property (Pα) relative to the Euclidean distance is valid forH for all α∈(0, c−1Λ−1λ).

Proof. Letu∈Cc2(Ω) be fixed. Settingv=d−αn uand using Leibniz’ rule we have Q(d−αn u)−Q(u, d−2αn u)

=Q(v)−Q(dαnv, d−αn v)

=

|η|=2

|ζ|=2

aηζ

DηvDζv−Dη

dαnv Dζ d−αn v

dx

(8)

=

|η|=2

|ζ|=2

γ≤ηδ≤ζ γ+δ>0

cηγcζδaηζ

Dγdαn Dδd−αn (Dη−γv)(Dζ−δv)dx

≤cΛ

0≤i,j≤2 i+j>0

idαnjd−αn |∇2−iv| |∇2−jv|dx.

But, by (D1) and (D2),

∇d±αn =αd±α−1n and 2d±αn ≤cαd±α−2n , and we thus obtain

Q(v)−Q

dαnv, d−αn v ≤cΛα

|∇2v|2+|∇v|2 d2n +v2

d4n

dx≤cΛλ−1αQ(v). Hence, if αis such thatcΛλ−1α<1, then property (Pα) is valid forH.

3.2. Regular coefficients

The weak point of Theorem 4 is the poor information it provides on the range ofαfor which (Pα) is valid. In this subsection we shall consider operators of a more specific type and for which we shall see that (Pα) is valid for allα∈(0,12).

It will be useful in this subsection to drop the multi-index notation and write the quadratic form as

Q(u) =

N i,j,k,l=1

aijkluxixjuxkxldx, u∈H02(Ω).

We may clearly assume that the functionsaijkl have the following symmetries:

aijkl=ajikl, aijlk, aklij. (11)

We make the following additional assumptions on the coefficients{aijkl}: (i) There existθ, τ >0 such that

eachaijkl is differentiable in{x∈Ω :d(x)< θ}

(12.a)

|∇aijkl| ≤cd−1+τ on{x∈Ω :d(x)< θ}

(12.b) (ii)

N i,j,k,l=1

aijkl(x)ξiξkηjηl N i,j,k,l=1

aijkl(x)ξiξjηkηl, ξ, η∈RN, x∈. (12.c)

(9)

Without any loss of generality we assume thatτ in (i) is the same as in (D2).

Condition (ii) is a technical one, whose necessity is not clear. We present two examples in which it is valid.

Example1. Suppose that aijkl=bijbkl for some non-negative N×N matrix {bij}i,j. Then (ii) is valid by the Cauchy-Schwarz inequality for the non-negative form (ξ, η)!bijξiηj. This for example includes operators of the form

a(x)∆, for which we haveaijkl=a(x)δijδkl.

Example2. Suppose that aijkl=δijδklaik, where aik=aki0 for i, k=1, ..., N. Then it is easily seen that (ii) is again valid.

We choose the distance functiond(·) to be the one naturally associated withH, that is the one induced by the Finsler metric p(x, η) whose dual metric (cf. (15) below) is

p(x, ξ) = N

i,j,k,l=1

aijkl(x)ξiξjξkξl

1/4 . (13)

This implies in particular that the functiond(·) satisfies N

i,j,k,l=1

aijkl(x)dxidxjdxkdxl= 1, a.e. x∈. (14)

Indeed, the inequalityN

i,j,k,l=1aijkl(x)dxidxjdxkdxl1 is shown in [A, Lemma 1.3].

To prove the converse inequality letydenote a point of differentiability ofd. Then y has a unique nearest point y0∈∂Ω; so d(y)=d(y, y0)=:s. Letyt,t∈[0, s], be the geodesic joiningy0 andy parametrised by arc length so thatys=y. Then for small ε>0 we have on the one hand

d(ys−ε)−d(y) =d(y, ys−ε) =p(y, y−ys−ε)+o(ε), and on the other hand, by differentiability,

d(ys−ε)−d(y) =∇d(y)·(ys−ε−y)+o(ε). Hence

p(y,∇d(y)) = sup

ξ∈RN

∇d(y)·ξ p(y, ξ) lim

ε&0

∇d(y)·(ys−ε−y) p(y, y−ys−ε) = 1. (15)

We note that the metric is Riemannian if the symbol of the operatorHis the square of a polynomial of degree two.

(10)

We assume that our basic hypotheses (D1)–(D3) of the introduction are valid;

Concerning in particular the validity of condition (D2), we note that it is satisfied if enough regularity is imposed on the boundary and the coefficients. If for example the boundary isC3 and the coefficientsaijkllie inC3({x∈ªΩ:0≤d(x)<θ}), thend∈

C2({x∈ªΩ:0≤d(x)<θ}); see [LN, Section 1.3]. On the other hand, for the Euclidean distance aC2boundary is enough [GT, p. 354].

It is useful to introduce at this point a classAof integrals that are in a sense negligible.

Definition. A family of quadratic integral forms Tn(v), v∈Cc2(Ω), n∈N, be- longs to the class A if for anyε>0 there existscε>0 (independent ofn∈N) such that

|Tn(v)| ≤εQ(v)+cε

v2dx, n∈N, v∈Cc2(Ω). (16)

Lemma 5. Let In(v)=

bn(Dγv)(Dδv)dx, |γ|,|δ|≤2, be a term that results after expanding Q(dαnv, d−αn v) and integrating by parts a number of times. If bn

contains as a factor either a derivative of aηζ or a second-order derivative of dn, then {In}n=1∈A.

Proof. After expandingQ(dαnv, d−αn v) (cf. (19) below) we obtain a linear com- bination of integrals, and direct observation shows that each one of them has one of the following three forms (we switch temporarily to multi-index notation):

aηζd−4+|γ+δ|n (∇dn)η+ζ−γ−δ(Dγv)(Dδv)dx, |η|=|ζ|= 2, γ≤η, δ≤ζ, (a)

aηζd−3+|γ|n (∇dn)η−γ(Dζdn)v(Dγv)dx, |η|=|ζ|= 2, γ≤η, (b)

aηζd−2n (Dηdn)(Dζdn)v2dx, |η|=|ζ|= 2. (c)

(These are distinguished by the number of second-order derivatives ofdn that they contain – none, one and two, respectively.) Hence all resulting integrals have the

form

bn(x)(Dγv)(Dδv)dx, 0≤ |γ|,|δ| ≤2,

wherebn(x) is a product ofaηζ with powers and/or derivatives ofdn and, since∇dn

is bounded,

|bn(x)| ≤cdn(x)−4+|γ+δ|, x∈. (17)

(11)

In cases (b) and (c) however, where bn(x) contains as a factor at least one second- order derivative ofdn, it follows from condition (D2) of the introduction that we have something more, namely

|bn(x)| ≤cdn(x)−4+|γ+δ|+τ, x∈. This easily implies that the integral lies inAin this case.

Suppose now that we integrate by parts in the integral (a) above, transferring one derivative from, say,Dγv, (|γ|≥1), to the remaining functions. If the derivative being transfered is∂/∂xi, we obtain – in an obvious notation – the integral

[aηζd−4+|γ+δ|n (∇dn)η+ζ−γ−δ(Dδv)]xi(Dγ−eiv)dx.

If the derivative ∂/∂xi “hits” either aηζ or one of the factors that make up (∇dn)η+ζ−γ−δ we obtain an integral of the form

bn(x)(Dγ−eiv)(Dδv)dx,

where|bn|≤cd−1+τn d−4+|γ+δ|n ; hence this integral belongs toA.

Example. We illustrate the last lemma with an example: in (21) below there appears the integral

In(v) =

aijkld−2n dxidxjvvxkxldx.

Letting {Tn}n=1,{Tn}n=1denote elements inAwe compute

aijkld−2n dxidxjvvxkxldx

=

(aijkl)xkd−2n dxidxjvvxldx+2

aijkld−3n dxidxjdxkvvxldx

aijkld−2n dxixkdxjvvxldx−

aijkld−2n dxidxjxkvvxkdx

aijkld−2n dxidxjvxkvxldx

=

aijkld−3n dxidxjdxk(v2)xldx−

aijkld−2n dxidxjvxkvxldx+Tn(v)

= 3

aijkld−4n dxidxjdxkdxlv2dx−

aijkld−2n dxidxjvxkvxldx+Tn(v).

(12)

Note. The summation convention over repeated indices will be used from now on.

Lemma 6. There exists {Tn}n=1∈Asuch that Q(v)−Q

dαnv, d−αn v = 2α2

aijkld−2n dxidxjvxkvxldx (18)

+4α2

aijkld−2n dxidxkvxjvxldx

(α4+11α2)

d−4n v2dx+Tn(v) for all v∈Cc2(Ω).

Proof. Forβ=αandβ=−αwe have

(dβnv)xixj=dβnvxixj+βdβ−1n dxivxj+βdβ−1n dxjvxi

(19)

+β(β−1)dβ−2n dxidxjv+βdβ−1n dxixjv.

We substitute inQ(dαnv, d−αn v) and expand. Now, by Lemma 5 all terms containing second-order derivatives of dn belong to A. Further, the symmetries (11) ofaijkl

give

aijkldxidxjdxkvxl=aijkldxidxkdxlvxj=..., aijkldxidxjvxkxl=aijkldxkdxlvxixj=..., (20)

aijkldxidxlvxjvxk=aijkldxidxkvxjvxl=....

Denoting by {Tn}n=1an element ofAwhich may change within the proof we thus arrive at

Q

dαnv, d−αn v =

aijkl[vxixjvxkxl+2α2d−2n dxidxjvvxkxl4α2d−2n dxidxkvxjvxl

(21)

+4α2d−3n dxidxjdxkvvxl

+α2(α21)d−4n dxidxjdxkdxlv2]dx+Tn(v).

We integrate by parts the second and fourth terms in the last integral. By Lemma 5, all terms that contain either derivatives of aijkl or second-order derivatives ofdn, belong toA. Hence, denoting always by{Tn}n=1 a generic element ofAwe obtain (cf. the example above)

aijkld−2n dxidxjvvxkxldx= 3

aijkld−4n dxidxjdxkdxlv2dx

aijkld−2n dxidxjvxkvxldx+Tn(v)

(13)

and, similarly,

aijkld−3n dxidxjdxkvvxldx=3 2

aijkld−4n dxidxjdxkdxlv2dx+Tn(v). Substituting in (21) yields

Q

dαnv, d−αn v =

aijkl[vxixjvxkxl4α2d−2n dxidxkvxjvxl

2α2d−2n dxidxjvxkvxl

+(α4+11α2)d−4n dxidxjdxkdxlv2]dx+Tn(v). Recalling that (14) holds, relation (18) follows.

Lemma 7. Let v∈Cc2(Ω) andw=d−3/2n v. Then Q(v)−Q(dαnv, d−αn v) = 2α2

aijkldndxidxjwxkwxldx +4α2

aijkldndxidxkwxjwxldx +

−α4+5α2

2 d−1n w2dx+Tn(v), where {Tn}n=1∈A.

Remark 1. When working with the functionw, all integrals have the form

bn(x)(Dγw)(Dδw)dx where the functionbn satisfies

|bn(x)| ≤cdn(x)−1+|γ+δ|, x∈. (22)

Such an integral lies inAif in addition

|bn(x)| ≤cdn(x)−1+|γ+δ|+τ, x∈,

for someτ >0; as before, these are precisely the integrals that contain either second- order derivatives ofdn or (first-order) derivatives ofaijkl.

Remark 2. Since dand dn differ by a constant, for the sake of simplicity we shall writedxi instead of (dn)xi, etc.

(14)

Proof of Lemma 7. We substitutevxi=d3/2n wxi+32d1/2n dxiwin (18). Recalling the symmetry relations (20) (with w in the place of v) and using the fact that aijkldxidxjdxkdxl=1 we obtain

Q(v)−Q

dαnv, d−αn v

= 4α2

aijkld−2n dxidxk

d3/2n wxj+3

2d1/2n dxjw

d3/2n wxl+3

2d1/2n dxlw

dx +2α2

aijkld−2n dxidxj

d3/2n wxk+3

2d1/2n dxkw

d3/2n wxl+3

2d1/2n dxlw

dx

(α4+11α2)

d−1n w2dx+Tn(v)

= 4α2

aijkld−2n dxidxk

d3nwxjwxl+3d2ndxjwxlw+9

4dndxjdxlw2

dx +2α2

aijkld−2n dxidxj

d3nwxkwxl+3d2ndxkwxlw+9

4dndxkdxlw2

dx

(α4+11α2)

d−1n w2dx+Tn(v)

= 4α2

aijkldndxidxkwxjwxldx+2α2

aijkldndxidxjwxkwxldx +

−α4+5α2

2 d−1n w2dx+18α2

aijkldxidxjdxkwxlw dx+Tn(v). But the last integral belongs to Aby an integration by parts; hence the proof is complete.

Lemma 8. Letv∈Cc2(Ω) andw=d−3/2n v. Then

Q(v) =

aijkld3nwxixjwxkxldx+9 2

aijkldndxidxjwxkwxldx

3

aijkldndxidxkwxjwxldx+ 9 16

d−1n w2dx+Tn(v), where {Tn}n=1 is an element ofA.

Proof. We have

vxixj=d3/2wxixj+32d1/2n dxiwxj+32d1/2n dxjwxi+34d−1/2n dxidxjw+32d1/2n dxixjw.

Références

Documents relatifs

This section will be concerned primarily with the construction of ideal boundaries from lattices of bounded 9e-harmonic functions, where 9€ is a harmonic class hyperbolic on

It is also very useful when we make a systematic analysis of the gradients of solutions of nonlinear elliptic and parabolic problems near the boundary of the irregular domain which

Keywords: Higher order elliptic equations; Polyharmonic operators on half spaces; Dirichlet problem; Navier boundary condition; Liouville-type theorem; Decay estimates for

Then we give at a formal level the extension of the system (1.8)–(1.9) to dimensions higher than 2. The estimate for the gradient is unchanged but the one on the curvature is

However, the Neumann problem with boundary data in the negative Sobolev space ˙ W −1 p (∂ R n+1 + ) was investigated in [66, Sections 4 and 22] in the case of harmonic and

Our results are based upon a priori estimates of solutions of (E) and existence, non-existence and uniqueness results for solutions of some nonlinear elliptic equations on

Boundary Harnack inequality and a priori estimates of singular solutions of quasilinear elliptic equations.. Marie-Françoise Bidaut-Veron, Rouba Borghol,

V´eron, The boundary trace and generalized boundary value problem for semilinear elliptic equations with coercive absorption, Comm. Stein, Singular integrals and