• Aucun résultat trouvé

Linear elasticity obtained from finite elasticity by Γ-convergence under weak coerciveness conditions

N/A
N/A
Protected

Academic year: 2022

Partager "Linear elasticity obtained from finite elasticity by Γ-convergence under weak coerciveness conditions"

Copied!
21
0
0

Texte intégral

(1)

Ann. I. H. Poincaré – AN 29 (2012) 715–735

www.elsevier.com/locate/anihpc

Linear elasticity obtained from finite elasticity by Γ -convergence under weak coerciveness conditions

Virginia Agostiniani

, Gianni Dal Maso, Antonio DeSimone

SISSA, via Bonomea 265, 34136 Trieste, Italy

Received 23 June 2011; received in revised form 15 March 2012; accepted 17 April 2012 Available online 23 April 2012

Abstract

The energy functional of linear elasticity is obtained asΓ-limit of suitable rescalings of the energies of finite elasticity. The quadratic control from below of the energy densityW (∇v)for large values of the deformation gradient∇vis replaced here by the weaker conditionW (v)|∇v|p, for somep >1. Energies of this type are commonly used in the study of a large class of compressible rubber-like materials.

©2012 Elsevier Masson SAS. All rights reserved.

1. Introduction

Consider an elastic body occupying a reference configurationΩ⊆Rn, withn2, subject to some deformation v:Ω→Rn. Assuming that the body is homogeneous and hyperelastic, the stored energy can be written as

Ω

W (v) dx,

where∇vis the deformation gradient and the energy densityW (F )0 is defined for everyF ∈Rn×nand is finite only for detF >0. We assume that the energy densityWis minimized at the value 0 by the identity matrixI, which amounts to saying that the reference configuration is stress free. We assume also thatW is frame indifferent, i.e., W (F )=W (RF )for everyF ∈Rn×nand everyRin the spaceSO(n)of rotations.

Since the deformationv(x)=xis an equilibrium when no external loads are applied, we expect that small external loadsεl(x)will produce deformations of the formv(x)=x+εu(x), so that the total energy is given by

Ω

W (I+εu) dxε2

Ω

lu dx. (1.1)

In the case∇ubounded, by Taylor-expandingW (I+εu)aroundI and rescaling (1.1) byε2, we obtain in the limitε→0 the formula

* Corresponding author.

E-mail addresses:vagostin@sissa.it(V. Agostiniani),dalmaso@sissa.it(G. Dal Maso),desimone@sissa.it(A. DeSimone).

0294-1449/$ – see front matter ©2012 Elsevier Masson SAS. All rights reserved.

http://dx.doi.org/10.1016/j.anihpc.2012.04.001

(2)

1 2

Ω

D2W (I )[∇u]2dx

Ω

lu dx, (1.2)

whereD2W (I )[F]2 is the second differential ofW atI applied to the pair[F, F]. By frame indifference, the first summand in (1.2) depends only on the symmetric parte(u)of the displacement gradient∇u, i.e.,

1 2

Ω

D2W (I )[∇u]2dx=1 2

Ω

D2W (I ) e(u)2

dx.

This functional is the linearized elastic energy associated with the displacementu.

This elementary derivation of linear elasticity requires onlyC2regularity ofW nearI, and hence in a neighbour- hood ofSO(n), by frame indifference. However, it does not guarantee that the minimizers of the most natural boundary value problems for (1.1) converge to the minimizer of the corresponding problems for the limit functional (1.2).

Convergence of minimizers has been established in[5]in the framework ofΓ-convergence, under the assumption W (F )d

F,SO(n)2

, (1.3)

whered(F,SO(n))is the distance ofF fromSO(n). The main result of the present paper is that the same conclusion holds if (1.3) is satisfied only in a neighbourhood ofSO(n), while the weaker condition

W (F )cd

F,SO(n)p

, for some 1< p2 andc >0, (1.4)

is assumed far fromSO(n). Similar results have been obtained in[11]assuming also a bound of orderpfrom above.

The reason for considering energies satisfying (1.4) without any bound from above is not purely academic. Indeed, for a large class of compressible rubber-like materials, (1.4) is the appropriate behaviour (see Remark2.8and the discussion in[1]for a multiwell case).

The main tool for the proof of the compactness of the minimizers considered in [5] is theGeometric Rigidity Lemmaof[8]. To obtain the same result when (1.3) holds only nearSO(n), while (1.4) holds far fromSO(n), we need a version with two exponents of the Geometric Rigidity Lemma, similar to those used in[3,10,11].

The proof of theΓ-convergence in the present paper has been renewed with respect to[5], and also with respect to the further improvements introduced in[12]. The main simplification relies on some arguments developed in[8]for the rigorous proof of dimension reduction results.

Moreover, the strong convergence inW1,pof the minimizers is obtained by adapting to our techniques some ideas introduced in[12]for the case of multiwell energies satisfying the analog of (1.3). We hope that all our results can be extended to multiwell energies satisfying only (1.4) far from the wells.

2. Setting of the problem and main results

Throughout the paper,d(·,·)denotes the Euclidean distance both between two points and between a point and a set.

The space ofn×nreal matrices is identified withRn×n;SO(n)is the set of rotations,Sym(n)andSkw(n)the sets of symmetric and skew-symmetric matrices, respectively, Psym(n) the set of positive definite symmetric matrices, Lin+(n) the set of invertible matrices with positive determinant. Given M∈Rn×n,symM andskwM denote the symmetric and the skew-symmetric part ofM, respectively.

The reference configuration Ω is a bounded connected open set of Rn with Lipschitz boundary ∂Ω. We will prescribe a Dirichlet condition on a part DΩ of ∂Ω with Lipschitz boundary in ∂Ω, according to the following definition.

Definition 2.1.Let us define

Q:=(−1,1)n, Q+:=(−1,1)n1×(0,1),

Q0:=(−1,1)n1× {0}, Q+0 :=(−1,1)n2×(0,1)× {0}.

We say that E∂Ω has Lipschitz boundary in ∂Ω if it is nonempty and for everyx in the boundary ofE for the relative topology of∂Ω there exist an open neighbourhood U of x inRn and a bi-Lipschitz homeomorphism ψ:UQsuch that

ψ (UΩ)=Q+, ψ (U∂Ω)=Q0, ψ (UE)=Q+0.

(3)

The Sobolev spaceW1,p(Ω,Rn)will be denoted byW1,p. To deal with the Dirichlet boundary condition, for every hW1,pwe introduce the set

Wh1,p:=

uW1,p: u=hHn1-a.e. onDΩ

, (2.1)

where the equality onDΩrefers to the traces of the functions on the boundary∂Ω, andHn1denotes the(n−1)- dimensional Hausdorff measure.

We suppose the material to be hyperelastic and that the stored energy densityW:Ω×Rn×n→ [0,∞]isL ×B- measurable, whereL andB are theσ-algebras of the Lebesgue measurable subsets of Rn and Borel measurable subsets ofRn×n, respectively. We assume thatW satisfies the following properties for a.e.xΩ:

(i) W (x,·)is frame indifferent;

(ii) W (x,·)is of classC2 in some neighbourhood ofSO(n), independent ofx, where the second derivatives are bounded by a constant independent ofx;

(iii) W (x, F )=0 ifFSO(n);

(iv) W (x, F )gp(d(F,SO(n))), for some 1< p2, wheregp: [0,∞)→Ris defined by gp(t):=

t2

2, if 0t1,

tp

p +12p1, ift >1. (2.2)

Observe that these assumptions are compatible with the conditionW (x, F )= ∞, if detF 0, which is classical in the context of finite elasticity. Also, observe thatgp is a convex function. In what followsD2 denotes the second differential with respect to the variable F ∈Rn×n, so that D2W (x, F )[M]2 means the second differential of the functionW with respect toF, evaluated at the point(x, F )and applied to the pair[M, M]. By frame indifference, for a.e.xΩwe have that

D2W (x, I )[M]2=D2W (x, I )[symM]2, for everyM∈Rn×n. (2.3) Together with assumption (iv), this implies that the quadratic formD2W (x, I )[·]2is null onSkw(n)and satisfies the coerciveness condition

D2W (x, I )[symM]2|symM|2, for a.e.xΩand everyM∈Rn×n. (2.4) Energy densities for which estimate (iv) holds only with 1< p <2 are commonly used in the study of compressible elastomers (see Remark2.8).

The load is modelled by a continuous linear functionalL :W1,p→R. IfvW1,prepresents the deformation of the elastic body, the stable equilibria of the elastic body are obtained by minimizing the functional

Ω

W (x,v) dxL(v),

under the prescribed boundary conditions. We are interested in the case where the load has the formεL and we want to study the behaviour of the solution asεtends to zero. We write

v=x+εu

and we assume Dirichlet boundary condition of the form v=x+εh Hn1-a.e. onDΩ,

with a prescribedhW1,. The corresponding minimum problem forubecomes min

W1,p Ω

W (x, I+εu) dxεL(εu)

, (2.5)

where the termεL(x)has been neglected since it does not depend onu. The following theorem is the main result of the paper. It describes the behaviour of the minimizers of (2.5).

(4)

Theorem 2.2. Assume that W :Ω ×Rn×n → [0,∞] satisfies conditions (i)–(iv) for some 1 < p2, and let hW1,. For everyε >0let

mε:= inf

uWh1,p

1 ε2

Ω

W (x, I+εu) dxL(u)

, (2.6)

and let{uε}ε>0be a sequence such that 1

ε2

Ω

W (x, I+εuε) dxL(uε)=mε+o(1). (2.7)

Then,{uε}converges strongly inW1,pto the unique solution of the problem m:= min

uWh1,2

1 2

Ω

D2W (x, I ) e(u)2

L(u)

, (2.8)

wheree(u):=sym(u). Moreover,mεm.

In the case 1< p <2, Theorem2.2asserts that a sequence of “almost minimizers” inWh1,p for theε-problems converges to a minimizer for the limit problem in a different Sobolev space: indeed, the limit problem is formulated inWh1,2.

In the case p=2, weak convergence of the “almost minimizers” has already been proved in[5]. Theorem2.2 extends this result to the case 1< p2 and provides also strong convergence. The proof is based on the following three results which are proved in Sections3,4, and5, respectively. These involve the functionalsFε,F :W1,p→ [0,∞]defined by

Fε(u):=

1

ε2

ΩW (x, I+εu) dx, ifuWh1,p,

, otherwise, (2.9)

and

F(u):=

1

2

ΩD2W (x, I )[e(u)]2dx, ifuWh1,2,

, otherwise, (2.10)

and the functionalsGε,G :W1,p(−∞,∞]defined by

Gε:=FεL, G:=FL. (2.11)

Observe that, due to the growth property (iv) ofW, the functionalsGεandG are bounded from below.

Theorem 2.3.Assume thatW:Ω×Rn×n→ [0,∞]satisfies conditions(i)–(iv)for some1p2. There exists a constantC >0depending onΩ,∂DΩ, andpsuch that for everyhW1,pand every sequence{uε} ⊆Wh1,pwe have

Ω

|∇uε|pdxC

1+Fε(uε)+

DΩ

|h|dHn1 2

, (2.12)

for everyε >0sufficiently small.

The previous theorem ensures that, if {uε}is a sequence in Wh1,p such that{Fε(uε)}is bounded, then {uε}is bounded inW1,p, hence a subsequence converges weakly inW1,p.

Theorem 2.4.Under the hypotheses of Theorem2.2, for everyεj→0we have that Fεj

−→Γ F, asj → ∞, in the weak topology ofW1,p.

(5)

Theorem2.4, together with the compactness result provided by Theorem2.3, implies the convergence of minima and the weak convergence of minimizers, using standard results onΓ-convergence. The next theorem and the previous remarks allow us to obtain the strong convergence of minimizers.

Theorem 2.5.Under the hypotheses of Theorem2.2, letεj→0and let{uj}be a recovery sequence foruWh1,2, that isuj uweakly inW1,pandFεj(uj)F(u). Then{uj}converges strongly inW1,p.

Remark 2.6 (On the condition DΩ = ∅). Observe that in Theorem 2.3 the assumption DΩ = ∅ is crucial.

When DΩ = ∅, inequality (2.12) is false, as the following example shows. Consider the simple case W (F ):=

gp(d(F,SO(n)))for everyFLin+(n). For everyε >0 and someRSO(n)\ {I}, set uε(x):=RI

ε x, xΩ.

In this case, we have that

Ω

|∇uε|pdx=|Ω||RI|p

εp → ∞, asε→0+, whereas

Fε(uε)= 1 ε2

Ω

gp d

I+εuε,SO(n)

dx=0, for everyε >0.

Remark 2.7(On the conditionhW1,).In Theorems2.2,2.4and2.5the hypothesishW1,cannot be replaced byhW1,2, unlessW satisfies suitable bounds from above, which are not natural in the context of finite elasticity.

Consider the simple caseDΩ=∂Ω,L =0, and assume that for somer >2 we have W (F )|F|r for|F|large enough.

By well-known properties of the images of Sobolev spaces under the trace operator, there exists hW1,2 such that

uW1,r: u=hHn1-a.e. on∂Ω

= ∅. (2.13)

Let us prove that Fε(u)= ∞ for every uW1,p. Assume by contradiction that there exists uW1,p with Fε(u) <∞. By (2.9) we have that ∇uLr, hence uW1,r, because Ω has Lipschitz boundary. This contra- dicts (2.13). Therefore{Fε}cannotΓ-converge toF, becauseF(h) <∞.

Remark 2.8(Model energy densities).A large class of models where the energy density grows quadratically near the wells and less than quadratically elsewhere is provided by rubber elasticity, when one wishes to take into account the compressibility of the material. We recall that we have formalized this growth behaviour by introducing, as bound from below of our energies, the function

gp d

·,SO(3)

, for some 1< p <2,

wheregpis the function defined in (2.2). For simplicity, we focus on the homogeneous case.

A common practice to pass from an incompressible model, with associated energy density W˜ defined on {F∈R3×3: detF=1}, to a corresponding compressible modelW(see, e.g.,[2,7,9]) is to define

W (F ):= ˜W

(detF )1/3F

+Wvol(detF ), for everyFLin+(3), whereWvolis such that

Wvol0 and Wvol(t)=0 if and only if t=1.

For example, we can takeWvolof the form Wvol(t)=c

t2−1−2 logt

, for everyt >0,

(6)

forc >0. Consider first theNeo-Hookeanincompressible model for hyperelastic materials, where the energy density is of the form

W˜N(F ):=a

|F|2−3

, for everyF∈R3×3with detF =1,

for a certaina >0. Following the procedure described above, we consider the corresponding compressible energy density defined for everyFLin+(3)by

WN(F ):= ˜WN F

(detF )1/3

+Wvol(detF )

=a

|F|2 (detF )2/3−3

+Wvol(detF ).

Let us check thatWN has “gp-growth”. By using the well-known inequality between arithmetic and geometric mean, it is easy to see that

WN 0 and WN(F )=0 if and only if FSO(3). (2.14)

Moreover, recalling the Green–St. Venant strain tensorE=12(FTFI )and using simple rules of tensor calculus, it turns out that in the small strain regime (that is, the regime of the deformation gradients which vary nearSO(3)), W has the expression

WN(F )=μ|E|2+λ

2tr2E+o

|E|2

, (2.15)

where

μ=2a, λ=4

a 3 +c

.

The parametersμandλ+23μhave the physical meaning of a shear modulus and a bulk modulus, respectively. Since

|E|213tr2Efor everyESym(3), from (2.15) we obtain that WN(F )min{μ,6c}|E|2+o

|E|2 , and in turn,

WN(F )1

2min{μ,6c}|E|2, (2.16)

if |E| is small enough, that is, ifd(F,SO(3))is small enough. Since|√

CI||CI|for everyCPsym(3), from (2.16) we obtain that

WN(F )1

8min{μ,6c}FTFI2=1

8min{μ,6c}d2

F,SO(3)

, (2.17)

ifd(F,SO(3))is sufficiently small. Now, we want to study the growth ofW in the regime|F| → ∞. In this case, if detF is bounded, then

WN(F )C|F|2−3aCd˜ 2

F,SO(3)

(detF bounded), (2.18)

for someC,C >˜ 0. In the case detF → ∞, we have that WN(F )K

|F|2

det2/3F +det2F

, for someK >0. By using Young’s inequality

xyxp p +yq

q 1

p+1 q =1

withx=(detF|F|3)1/2andy=(detF )1/2, it is easy to show that WN(F )K|F|3/2Kd˜ 3/2

F,SO(3)

(detF → ∞), (2.19)

(7)

for someK >˜ 0. (2.14), (2.17), (2.18) and (2.19) show thatWN hasgp-growth from below withp=32. It is important to notice thatWN has not quadratic growth everywhere. In particular,WN has not quadratic growth in the regime detF → ∞. This can be checked by taking into account deformation gradients of the type

F =

λ2 0 0 0 1λ 0

0 0 1

, withλ0. (2.20)

As a second example, we consider theMooney–Rivlincompressible model given, for somea,b >0, by WM(F ):=a

|F|2 (detF )2/3−3

+b

(detF )2/3F12−3

+Wvol(detF )

=WN(F )+b

(detF )2/3F12−3

, (2.21)

for everyFLin+(3), and derived from the corresponding incompressible version as explained before. The inequality between arithmetic and geometric mean implies that the second summand in (2.21) is nonnegative, so that, from (2.14), we have that

WM0 and WM(F )=0 if and only if FSO(3).

The formula for the small strain regime is given by (2.15), with μ=2(a+b), λ=4

a+b 3 +c

.

From the fact that WN has gp-growth and from the positiveness of the second summand of (2.21) thegp-growth of WM trivially follows. Also in this case, deformation gradients of the type (2.20) show thatWM does not grow quadratically everywhere.

Finally, we mention someOgden-typecompressible energy densities:

WO(F ):=

m i=1

ai

tr((FTF )γi/2) (detF )γi/3 −3

+Wvol(detF ),

defined for everyFLin+(3), for somem1 andai, γi>0,i=1, . . . , m. The formula forWO in the small strain regime is again given by (2.15), with

μ=2 m i=1

ai, λ=4

−1 3

m i=1

ai+c

.

Arguing similarly to the Neo-Hookean and the Mooney–Rivlin models, we obtain thatWO attains its minimum 0 atSO(3). By using Young’s inequality and proper counterexamples, it is possible to show thatWO hasgp-growth for some 1< p <2 (p depending on the exponentsγi), but not a quadratic growth in general, if 0< γi <3 for every i=1, . . . , mandγi>65for at least one indexi∈ {1, . . . , m}.

3. Compactness

In this and in the next sections we give the proofs of the results stated in Section2. To simplify the exposition, the proofs are given only whenW does not depend explicitly onx. The proofs in the general case require only minor modifications.

The compactness result requires the following extension of the well-known geometric rigidity result of[8], where a power ofd(v,SO(n))is replaced bygp(d(v,SO(n))).

Lemma 3.1(Geometric rigidity). Letgp be the function defined in(2.2). There exists a constantC=C(Ω, p) >0 with the following property:for everyvW1,pthere exists a constant rotationRSO(n)satisfying

Ω

gp

|∇vR| dxC

Ω

gp d

v,SO(n)

dx. (3.1)

(8)

Similar versions of Lemma 3.1 can be found in [3,10,11]. For sake of completeness, we give the proof in Appendix A.

We need two more lemmas in order to prove Theorem2.3.

Lemma 3.2.LetS⊆Rnbe a boundedHm-measurable set with0<Hm(S) <for somem >0. Then

|F|S:=min

ζ∈Rn

S

|F xζ|dHm

is a seminorm onRn×n. Define S0:=

xS: Hm

SBρ(x)

>0for everyρ >0 ,

and letaff(S0)be the smallest affine space containingS0. LetK⊆Rn×nbe a closed cone such that dim

Ker(F )

<dim aff(S0)

, for everyFK\ {0}. (3.2)

Then, there exists a constantC=C(S) >0such that C|F||F|S, for everyFK.

Proof. It is enough to repeat the proof of[5, Lemma 3.3], replacing theL2norm with theL1norm. 2

We will use the next lemma also in the proof of the Γ-convergence result. In what follows and in the rest of the paper we denote byCa positive constant which may change from line to line.

Lemma 3.3.Letε >0anduεWh1,p. Under the hypotheses of Theorem2.3, letRεSO(n)be a constant rotation satisfying(3.1)withv=x+εuε. Then,

|IRε|22

Fε(uε)+

DΩ

|h|dHn1 2

,

whereCdepends only onΩ,∂DΩ, andp.

Proof. Consider the deformationvε:=x+εuε. Lemma3.1tells us that there exists a constant rotationRεSO(n) such that

Ω

gp

|∇vεRε| dxC

Ω

gp d

vε,SO(n) dx,

whereCdepends only onΩandp. Then, by assumption (iv) onW, we have that

Ω

gp

|∇vεRε| dxC

Ω

W (vε) dx=2Fε(uε).

Jensen inequality thus implies gp

1

|Ω|

Ω

|∇vεRε|dx

2Fε(uε). (3.3)

Poincaré–Wirtinger inequality and the continuity of the trace operator give

DΩ

|vεRεxζε|dHn1C

Ω

|∇vεRε|dx,

whereζε:=|Ω1|

Ω(vεRεx) dxandC depends onΩ, so that, sincevε=x+εhHn1-a.e. onDΩ, we obtain

DΩ

(I−Rε)xζεdHn1C

Ω

|∇vεRε|dx+ε

DΩ

|h|dHn1

. (3.4)

(9)

Now, let us use Lemma3.2withS=DΩand withKequal to the closed cone generated byISO(n). Showing first that everyFKbelongs to the cone generated byISO(n)or toSkw(n), it is easy to prove that everyFK\ {0}

is such that dim

Ker(F )

< n−1.

On the other hand,∂Ω Lipschitz implies that the right-hand side of (3.2) is equal ton−1. Thus, we can apply Lemma3.2to(IRε)Kand write that

C|IRε|min

ζ∈Rn

DΩ

(I−Rε)xζdHn1, (3.5)

whereC depends onDΩ and not onε. From (3.4) and (3.5) we obtain that

|IRε|2C 1

|Ω|

Ω

|∇vεRε|dx 2

+ε2

DΩ

|h|dHn1 2

. (3.6)

We conclude the proof by distinguishing two cases. If

Ω|∇vεRε|dx|Ω|, then (3.3) and the definition ofgptell us that

1 2

1

|Ω|

Ω

|∇vεRε|dx 2

2Fε(uε).

Using this last inequality in (3.6), it turns out (2.12). If

Ω|∇vεRε|dx >|Ω|, again (3.3) and the definition ofgp tell us that

2Fε(uε) > 1 2.

This bound from below ofε2Fε(uε)gives trivially (2.12), in view of the fact that|IRε|2√ n. 2 For the proof of Theorem2.3we will need the following estimate

gp(s+t )C

gp(s)+t2

, for everys, t0, (3.7)

for a certainC depending onp. This estimate can be easily deduced from the convexity ofgpand from the growth properties ofgpwhich give

gp(t)1 pmin

tp, t2

and gp(2t )Cgp(t), for everyt0, for someCdepending onp.

Proof of Theorem2.3. LetRε be given by Lemma3.1forvε:=x+εuε, for everyε >0. By using (3.7), we have that

Ω

gp

|εuε| dxC

Ω

gp

|∇vεRε|

+ |IRε|2 dx

C

Ω

gp d

vε,SO(n)

dx+ |IRε|2

,

where in the last inequality we have used Lemma3.1. Assumption (iv) onW and Lemma 3.3then imply that for someC, depending onΩ,DΩ, andp,

Ω

gp

|εuε|

dxCε2

Fε(uε)+

DΩ

|h|dHn1 2

. (3.8)

(10)

In particular, from (3.8) and from the definition ofgpwe obtain

{xΩ:|εuε(x)|1}

|εuε|2dx2

Ω

gp

|εuε| dx

2

Fε(uε)+

DΩ

|h|dHn1 2

,

so that, by Hölder inequality, it turns out

{xΩ:|εuε(x)|1}

|εuε|pdx

{xΩ:|εuε(x)|1}

|εuε|2dx p/2

|Ω|1(p/2)

p

Fε(uε)+

DΩ

|h|dHn1 2p/2

p

1+Fε(uε)+

DΩ

|h|dHn1 2

. (3.9)

Note that in (3.9) we have used the fact that tp/21+t, for everyt0.

On the other hand, from (A.2) and again from (3.8) we obtain that

{xΩ:|εuε(x)|>1}

|εuε|pdxC

{xΩ:|εuε(x)|>1}

gp

|εuε| dx

2

Fε(uε)+

DΩ

|h|dHn1 2

. (3.10)

Inequalities (3.9) and (3.10) imply that (2.12) holds. 2

In the next remark we construct a counterexample which shows that Theorem 2.3 is not true in general for p(0,1).

Remark 3.4.Letp(0,1)and consider the simple case in whichΩis the open unitary ballB(0,1)inR2,W (F ):=

gp(d(F,SO(2)))for everyFLin+(2),h=0, andL =0. For anyε >0 and someα >0 to be chosen, we introduce the set

Sε:=

x∈R2: 1

2<|x|<1 2+εα

.

For every ε >0 sufficiently small, Sε is an open annulus strictly included in Ω. We want to define a sequence {uε} ⊆W01,psuch that the valuesFε(uε)are equibounded and

Ω|∇uε|pdx→ ∞asε→0+. In order to do this, we consider for everyε >0 arbitrarily small a functionϕεCc(Ω,R)such that supp(ϕε)B(0,12)Sε, 0ϕε1, ϕε≡1 onB(0,12)and

|∇ϕε| C

εα for someCindependent ofε. (3.11)

Then, we chooseRSO(2)\ {I}and define the function uε(x):=ϕε(x)RI

ε x, xΩ,

(11)

which belongs toCfor everyε >0 sufficiently small. Observe that

Ω

|∇uε|pdx

B0,12

|∇uε|pdx=π|RI|pp , so that

Ω|∇uε|pdx→ ∞asε→0+(for every choice ofα >0). Now, let us compute

uε(x)=1 ε

ϕε(x)(RI )+

(RI )x

⊗ ∇ϕε(x)

and observe that∇uε≡0 onΩ\ B

0,12

Sε

, so thatd(I+εuε,SO(2))≡0 on the same set. Thus, recalling that gpis increasing, it turns out that

ε2Fε(uε)

B0,12Sε

gp

|I+εuεR| dx

Sε

gp

|RI|

1+ |x||∇ϕε|

dx, (3.12)

where in the last inequality we have also used the fact thatϕε≡1 onB(0,12). Therefore, from (2.2) and (3.12) we obtain that

Fε(uε) C ε2

Sε

1+ |∇ϕε|p

dx, (3.13)

for someCindependent ofε. Using (3.11) and noticing that|Sε| =π εα+o(εα), (3.13) implies that Fε(uε) C

ε2

π εα+o εα

1+ 1 εαp

,

so that{Fε(uε)}turns out to be bounded wheneverα >12p. We end this section with the following corollary.

Corollary 3.5. Under the hypotheses of Theorem 2.3, the functionals Gε are equicoercive in the weak topology ofW1,p.

Proof. Lett∈Rand{uε}be a sequence withGε(uε)t, so that{uε} ⊆Wh1,p. Thus, by the definition ofGε (2.11), we have

Fε(uε)t+L(uε).

Theorem2.3implies that for everyεsufficiently small

Ω

|∇uε|pdxC

1+L(uε)+

DΩ

|h|dHn1 2

,

for someCindependent ofε. By Poincaré inequality, this gives uεpW1,pC

uεW1,p +1

, (3.14)

whereCnow depends also onhandL. Therefore, sincep >1, from (3.14) we obtain thatuεW1,pis bounded. 2 Observe that the proofs of Theorem2.3, Lemma3.3and Corollary3.5do not use the fact thatDΩ has Lipschitz boundary in∂Ω(see Definition2.1): actually, these results hold under the weaker hypothesisHn1(∂DΩ) >0.

(12)

4. Γ-convergence

Consider a sequenceεj →0+asj→ ∞. By Theorem2.3, we can characterize theΓ-limit in the weak topology ofW1,pin terms of weakly converging sequences (see[6, Proposition 8.10]). In particular, we have that

F(u):=Γ-lim inf

j→∞ Fεj(u)=inf

lim inf

j→∞ Fεj(uj): uj uweakly inW1,p ;

F(u):=Γ-lim sup

j→∞ Fεj(u)=inf

lim sup

j→∞ Fεj(uj): uj uweakly inW1,p

. (4.1)

Thus, in order to prove Theorem2.4, we will show thatF(u)F(u)andF(u)lim infj→∞Fεj(uj), for every uW1,pand everyuj uweakly inW1,p.

Proof of Theorem2.4. (I) We want to show thatF(u)F(u). Consider the nontrivial caseF(u) <∞, so that uWh1,2and

F(u)=1 2

Ω

D2W (I ) e(u)2

dx.

Suppose firstuW1,. The boundedness of∇uand assumption (ii) onW, together with the fact thatW (I )=0 and DW (I )=0, imply that

jlim→∞

1 εj2W

I+εju(x)

=1

2D2W (I )

u(x)2

, for a.e.xΩ, and that there existsC >0 such that for everyεj>0 sufficiently small

W (I+εju)ε2jC|∇u|2, a.e. inΩ.

Then, by dominated convergence and by (2.3), we obtain

jlim→∞

1 εj2

Ω

W (I+εju) dx=1 2

Ω

D2W (I ) e(u)2

dx.

Therefore, by (4.1), F(u)= lim

j→∞Fεj(u)F(u). (4.2)

Consider now the general caseuWh1,2. SinceDΩhas Lipschitz boundary in∂Ω, from PropositionA.2we have that there exists a sequence{uk} ⊆Wh1,such thatukustrongly inW1,2, ask→ ∞. Observe that by (4.2) we haveF(uk)F(uk)for everyk. Thus, by the weak lower semicontinuity ofFinW1,pand the strong continuity ofF inWh1,2, it turns out that

F(u)= lim

k→∞F(uk)lim inf

k→∞ F(uk)F(u).

(II) We want to prove that, if uj u weakly in W1,p, then F(u)lim infjFεj(uj). Consider the nontrivial case lim infj→∞Fεj(uj) <∞so that, up to a subsequence, we can suppose{Fεj(uj)}bounded and, in particular, {uj} ⊆Wh1,p. Let 1Bj be the characteristic function ofBj, where

Bj:=

xΩ:uj(x) 1

εj

. (4.3)

Claim 1.We have that{1Bjuj}is bounded inL2.

(13)

Proof. By Lemma3.1and by the growth hypothesis onW we have that for everyj there existsRjSO(n)such that

Ω

gpI+εjuj(x)Rjdxεj2CFεj(uj)Cεj2, (4.4)

where the last inequality follows from the boundedness of{Fεj(uj)}. Considering the set Aj:=

xΩ: I+εjuj(x)Rj3√ n

,

it is easy to check thatBjAj for everyj large enough, so that

Bj

|∇uj|2dx 2 ε2j

Aj

|εjuj+IRj|2+ |IRj|2

dx. (4.5)

Therefore, by using (A.1) and the definition ofAj, from (4.5) we obtain that

Bj

|∇uj|2dx C ε2j

Aj

gp

|εjuj+IRj|

+ |IRj|2 dx

C

1+|IRj|2 εj2

, (4.6)

where in the last inequality we have used (4.4) andC depends onΩ andp. Since{F(uj)}is bounded, Lemma3.3 tells us that|IRj|22jis bounded. This fact, together with (4.6), gives the claim. 2

Claim 2.uL2and, up to a subsequence, we have that 1Bjuju weakly inL2.

Proof. By Claim 1, we have that, up to a subsequence,

1Bjuj v weakly inL2, (4.7)

for somevL2. Let us prove that

1Bjcuj→0 strongly inLα, (4.8)

for everyα∈ [1, p). We first observe that|Bjc| →0, by Chebyshev inequality. Taking into account the boundedness of{uj}inW1,p, by Hölder inequality we obtain

Ω

|1Bjcuj|αdx

Ω

|∇uj|pdx α/p

Bjc(pα)/pCBjc(pα)/p→0, which proves (4.8).

The weak convergence ofuj touinW1,p implies also that∇ujuweakly inLα, for everyα∈ [1, p). This fact, together with (4.8), gives that

1Bjuj=(uj−1Bc

juj) u weakly inLα, (4.9)

for everyα∈ [1, p). By (4.7) and (4.9) we conclude that∇u=vL2and Claim 2 follows. 2 From assumptions (ii) and (iii) onW it is easy to show that

W (I+F )1

2D2W (I )[F]2η

|F|

|F|2, for everyF ∈Rn×n,

whereηis an increasing function on[0,∞)such thatη(t )→0 ast→0+. Therefore, we can write

Références

Documents relatifs

We prove an extended continuous mapping theorem for outer almost sure weak convergence in a metric space, a notion that is used in bootstrap empirical processes theory.. Then we

It has been pointed out in a couple of recent works that the weak convergence of its centered and rescaled versions in a weighted Lebesgue L p space, 1 ≤ p &lt; ∞, considered to be

Abstract. Since its introduction, the pointwise asymptotic properties of the kernel estimator f b n of a probability density function f on R d , as well as the asymptotic behaviour

Poggiolini, Lower semicontinuity for quasiconvex integrals of higher order, Nonlinear Differential Equations Appl.. Kristensen, Lower semicontinuity of quasi-convex integrands in

about the central limit theorem for a sequence of Hilbert space valued continuous local

Remark also that ℓ − 0 is by no means unique, as several profiles may have the same horizontal scale as well as the same vertical scale (in which case the concentration cores must

About the convergence of PSO, Van Den Bergh and Engelbrecht (2006) looked at the trajectories of the particles and proved that each particle converges to a stable point

Central limit theorems for the Wasserstein distance between the empirical and the true distributions. Correction: “Central limit the- orems for the Wasserstein distance between