• Aucun résultat trouvé

Vortex Analysis of the Periodic Ginzburg-Landau Model

N/A
N/A
Protected

Academic year: 2022

Partager "Vortex Analysis of the Periodic Ginzburg-Landau Model"

Copied!
14
0
0

Texte intégral

(1)

www.elsevier.com/locate/anihpc

Vortex analysis of the periodic Ginzburg–Landau model

Hassen Aydi

a

, Etienne Sandier

b,c,

aInstitut supérieur d’informatique de Mahdia, route de Réjiche Km 4, BP 53, Mahdia 5121, Tunisie bUniversité Paris-Est, LAMA – CNRS UMR 8050, 61, Avenue du Général de Gaulle, 94010 Créteil, France

cInstitut Universitaire de France, France

Received 8 January 2008; received in revised form 19 September 2008; accepted 19 September 2008 Available online 29 October 2008

Abstract

We study the vortices of energy minimizers in the London limit for the Ginzburg–Landau model with periodic boundary condi- tions. For applied fields well below the second critical field we are able to describe the location and number of vortices. Many of the results presented appeared in [H. Aydi, Doctoral Dissertation, Université Paris-XII, 2004], others are new.

©2008 Elsevier Masson SAS. All rights reserved.

Résumé

Nous étudions les tourbillons des minimiseurs de l’énergie de Ginzburg–Landau en supraconductivité dans la limite de London, et pour des conditions aux limites périodiques. Lorsque le champ magnétique appliqué est petit devant le second champ critique, nous décrivons le nombre et la localisation de ces tourbillons. Certains de ces résultats étaient présents dans [H. Aydi, Doctoral Dissertation, Université Paris-XII, 2004], d’autres sont nouveaux.

©2008 Elsevier Masson SAS. All rights reserved.

Keywords:Superconductivity; Ginzburg–Landau equations; Vortices; Periodic boundary conditions

Mots-clés :Supraconductivité ; Équations de Ginzburg–Landau ; Tourbillons ; Conditions aux limites périodiques

1. Introduction

Periodic solutions of the Ginzburg–Landau equations of superconductivity with vortices arranged in a lattice were first introduced in the famous work of A. Abrikosov [1], based on an analysis of the linearized equations about the normal solution, where the order parameter is 0. Since then many contributions to the study of this type of solutions have appeared both from physicists and mathematicians, establishing rigorously the existence of Abrikosov type so- lutions ([20], more recently [5] established the existence of many other families of periodic solutions) or investigating the energy or the minimality of these solutions [18,17] or their numeric analysis [12]. We may refer to the review paper [10] for a broad overview of the subject from a physicist’s point of view.

This research was supported by the ANR, project ANR-05-JCJC-0213.

* Corresponding author at: Université Paris-Est, LAMA – CNRS UMR 8050, 61, Avenue du Général de Gaulle, 94010 Créteil, France.

E-mail address:sandier@univ-paris12.fr (E. Sandier).

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

doi:10.1016/j.anihpc.2008.09.004

(2)

There is a convenient variational setting for the periodic Ginzburg–Landau equations, we may refer to [20,13] for a mathematically oriented presentation of this setting, the existence of minimizers is proved in [20], the regularity of solutions is established [13] together with other properties of minimizers or critical points of the Ginzburg–Landau functional. Let us also cite the recent works [6] and [14], without going into further detail.

Opposite to the case of solutions close to 0, the so-called London limit or London approximation deals with so- lutions where the order parameter has modulus close to 1, except in small areas (the vortex cores). This limit was investigated by A. Abrikosov from the beginning, see also [10] for the use of this approximation in numerous situa- tions, or the classical textbook [25]. The mathematical justification of this approach following the methods introduced in [8] may be found in [22], where it is in addition applied to describe the vortices of minimizers of the Ginzburg–

Landau energy in a fixed domain Ω (in units of the penetration depth), when the Ginzburg–Landau parameterκ is large and the applied field is small compared to theR2-upper critical field, i.e. in the parameter region where mini- mizers may indeed be analyzed using the London approximation.

In [7], the first author carried out for the case of periodic boundary conditions the analysis carried out in [22] in the case of natural boundary conditions in a bounded domain. It was established among other things that in the periodic case, the regime of the lower critical field isHc1(κ)=logκ/2 to leading order asκ→ +∞, see below for a precise statement. Note that in the units used in [7], and in most regimes where vortices are present, there is a divergent number of vortices asκ→ ∞in each periodicity cell. The periodicity cell is large compared to the intervortex distance, and thus the periodicity constraint is not a strong one.

The present paper contains both results present in [7] and new results which complete them, in order to give a fairly unified description of the vortices of minimizers of the Ginzburg–Landau functional in the limitκ→ ∞, for applied fields well below the upper critical field.

2. Statement of the results

Notation. Throughout,K will denote a parallelogram with area 1 generated by two vectors(u, v). We will denote byLthe group of translations generated by(u, v).

The parameter of our asymptotic analysis will be the inverse of the Ginzburg–Landau parameterκ, denoted byε.

The applied fieldhexis a positive function ofε∈R+, and we define Δex=hex−1

2|logε|.

As noted in the introduction, 12|logε|is the regime of the so-called lower critical fieldHc1(ε), to be defined below, this motivates the notationΔex.

Definition 1.We defineHper1 to be the set of(u, A)inHloc1 (R2)such that for any integersk, ∈Zthe configuration (u(· +ku+v), A( · +ku+v)) is gauge-equivalent to(u, A).

In a more geometrical language,uis a section of a complex line bundle over the torusR2/L, andAis a connection.

Then curlAis the curvature of the connection and 1

K

curlA

is an integer, the first chern class of the line bundle.

Given(u, A)inHper1 , and an for eachε >0 an applied magnetic fieldhex(ε), we define for anyε >0 Gε(u, A)=1

2

K

|∇uiAu|2+1

2(curlAhex)2+ 1 2ε2

1− |u|22 .

Then the problem of minimizingGε(u, A)overHper1 is well posed.

Proposition 2.1.The minimum ofGε(u, A)overHper1 is achieved.

(3)

We refer to [20] for the proof.

A useful fact is

Proposition 2.2.Given any(u, A)Hper1 , then 1

K

curlA

is an integer. Moreover, if(u1, A1)minimizes the Ginzburg–Landau energy with parametersε >0andhex=h1, and if(u2, A2)minimizes the Ginzburg–Landau energy with parametersε >0andhex=h2> h1, thenn2n1, where we have set fori=1,2

ni= 1 2π

K

curlAi.

Proof. The fact that

KcurlA∈2πZwas already mentioned.

For the second statement, denoteG1(resp.G2) the Ginzburg–Landau functional with parameters(ε, h1)(resp.

(ε, h2)), thenG1(u1, A1)G1(u2, A2)andG2(u1, A1)G2(u2, A2), hence (G1G2)(u1, A1)(G1G2)(u2, A2),

which translates as (h2h1)

K

curlA1+1 2

K

h22h21

(h2h1)

K

curlA2+1 2

K

h22h21 ,

hence the result. 2

Remark 2.1.As a consequence of the above proposition, for eachε >0 there is a well-defined valueHc1(ε)which we call the first critical field, and such that the minimizers of the Ginzburg–Landau functional with parameters(ε, hex) satisfy n=0 if hex< Hc1, and n=0 if hex> Hc1. Note that in the former case, the minimizers are necessarily gauge-equivalent to the constant superconducting solutionu=1,A=0, see below.

Theorem 1.Denote by(uε, Aε)any minimizer ofGε and lethε=curlAε. We define the integernεby 2π nε=

K

hε.

Then the following behaviour ofhε,nεholds, according to the regime considered for the applied fieldhex. (1) If1 Δex 1/ε2, then, asε→0,

hε

2π nε →1 inW1,pfor anyp <2, andnεΔex

. (2.1)

(2) IfΔexis bounded independently ofεthen so ishεW1,p, for anyp <2. In particularnεis bounded independently ofε. If{ε}is a subsequence such that{hε}εconverges tohandΔexconverges to a valueΔex, thennεn∈N along the same subsequence, thus in particularnε=nfor small enoughε, and there arendistinct points{ai}i

inKsuch that

Δh+h=2π

n

i=1

δai. (2.2)

Moreover, denote byP the finite families of points inKand forp=(p1, . . . , pk)Plet W (p)=lim

ρ0

π nlogρ+1 2

K\ iB(pi,ρ)

|∇hp|2+h2p

+n

γ−2π Δex

, (2.3)

(4)

where hp is the unique K-periodic solution ofΔhp+hp=2πk

i=1δpi. Then (a1, . . . , an)minimizes W overP.

The numberγin(2.3)was introduced in[8], we define it as in[22], Proposition3.11, as γ= lim

R→+∞πlogR+1 2

B(0,R)

|∇u0|2+(1− |u0|2)2

2 , (2.4)

whereu0is the unique solution ofΔu0=u0(1−|u0|2)inR2of the formu0(r, θ )=f (r)e, withf:R+→R+. (3) There exists a possibly negativeΔ1∈Rsuch that ifΔex< Δ1andεis small enough, thenn=0. In this case

(uε, Aε)is gauge-equivalent to the Meissner solution(1,0).

Remark 2.2.The first item in the above theorem was proved initially in [7].

The main difference between the periodic case and the case of a bounded domain is that in the former the distribu- tion of vortices is always uniform, and hence may be described by a unique number, namely the number of vortices, at least if it tends to+∞asε→0. This is a big simplification over the case of a bounded domain, together with the fact that here the number of vortices is given by the integral ofh.

Remark 2.3.The fact that the minimum ofW onP is achieved is a consequence of the above theorem, it could also be derived directly from (2.3). From Proposition 2.2, ifΔ2> Δ1∈Rand ifp1(resp.p2) minimizesWforΔex=Δ1 (resp.Δex=Δ2), then the number of points inp2is larger than the number of points inp1. This shows that there are critical values ofΔexfor which the number of points for a minimizer experiences a jump, and it would be interesting to show that it jumps by one unit only. This would show the existence of an increasing sequence of critical values k)k∈Nfor which the number of vortices jumps fromk−1 tok.

Note that at a jump there exists minimizers with different numbers of vortices, hence the minimizer of W need not be unique. Even ifΔexis not a critical value, i.e. if minimizers ofW all have the same number of points, if this number is 2 for instance then the symmetry of the periodicity is broken and there are several minimizers as well.

Remark 2.4.Regarding the finer structure of vortices, they are expected in general to arrange themselves in periodic lattices, and the hexagonal lattice is supposedly optimal. Several rigorous mathematical results in this direction have been proved (see for instance [3,2,23]). In the periodic setting, the strongest version of this conjecture should be true: If the lattice generated by(u, v) is the hexagonal lattice and ifnε=k2for some integerk, then the minimizing configuration should be periodic w.r.t. the vectors(u/k, v/k). A limiting form of this conjecture would be that – still in the case of a hexagonal lattice – in the case of a bounded number of vortices, and assumingn=k2, thenh is periodic w.r.t. the vectors(u/k, v/k).

Remark 2.5.Note that for a minimizer(uε, Aε)ofGεand arbitrary positive values of the parametersε,hex, we have 1

2

K

(hεhex)2Gε(uε, Aε)Gε(1,0)=1 2h2ex, wherehε=curlAε. Therefore the following holds

0nε, Gε(uε, Aε)1

2h2ex. (2.5)

Moreover, ifnε=0, thenhε=0 as well. Then(uε, Aε)is gauge equivalent to a configuration(uε,0)andGε(uε,0) Gε(1,0). But, since(1,0)is clearly the unique minimizer ofGε among configurations(u, A)such thatA=0, we deduce thatuε=1 and that(uε, Aε)is gauge-equivalent to the Meissner solution.

The paper is organized as follows: In Section 3 we construct a test configuration which will be useful in the proof of case (1) of the theorem. In Section 4, matching the upper bound of the previous section with appropriate lower bounds, we prove case (1) of the theorem, with anL2instead ofW1,pconvergence, we also prove anL2bound forhε in case (2), and we prove case (3) completely. In Section 5 we discuss the improvement fromL2toW1,pconvergence without going into the details, since similar arguments appear in [22] in the context of natural, instead of periodic,

(5)

boundary conditions. Finally in the last section we finish the proof of case (2), i.e. identify the limithas a minimizer ofW. This involves arguments from [8,9] adapted to the periodic setting, that are in part sketched.

3. Upper bound

The theorem is proved by energy comparison with an appropriate test configuration(vε, Bε), which will be peri- odic. It is defined below, rather quickly since this construction appears elsewhere (see for instance [21,7] or [4]). In the case whereΔexC, which corresponds to cases (2) and (3) of the theorem, the upper bound in (2.5) will suffice, hence we assume from now on that

1 Δex 1/ε2.

Then, we definemε as the integer such that√mε is the integer part of Δex

.

SinceΔextends to+∞asε→0 we have mεΔex

.

We then divideKintomεidentical parallelograms generated by the vectorsuε= u/

mε andvε= v/

mε. Denote byKεone of these parallelograms, byaεits center, and byfεthe solution inKε, with periodic boundary conditions, of the equation

Δfε=2π δaε− 2π

|Kε|.

Note that such a solution exists since the integral of the right-hand side overKεis 0. It is then naturally extended by periodicity to a functionfεdefined in all ofR2and satisfying

Δfε=2π

k,∈Z

δaε+kuε+vε− 2π

|Kε|. Then, we letBε:R2→R2be such that

curlBε= 2π

|Kε|

and writevε=ρεeε, whereρε is periodic with respect to the vectors (uε,vε) and is defined inKε by ρε(x)= min(|xaε|/ε,1), and whereϕεis such that−∇fε= ∇ϕεBε. This latter equation can be solved in the sense that the curl of(Bε− ∇fε)is

k,∈Z

δaε+kuε+vε,

hence there exists a functionϕε defined modulo 2π except at the pointsaε+kuε+vε which satisfies the identity.

Then lettingvε=ρεeε is legitimate, since precisely from the definition ofρε we haveρε(aε+kuε+vε)=0 for any integersk, .

To estimate the energy, note that the integrand ofGε(vε, Bε)is precisely 1

2

ρε2|∇fε|2+ |∇ρε|2+ 1 2ε2

1−ρε22

+ 2π

|Kε|−hex 2

. (3.1)

Its integral overKεis easily estimated (see [21,7] or [4]). Putting aside the last term, the contribution from the region whereρε=1, which isB(aε, ε)is bounded by a constant independent ofεfrom scaling arguments. The integrand outsideB(aε, ε)reduces to 12|∇fε|2which, with the price of an error bounded independently ofε, may be replaced by 12|∇log|xaε||2, whose integral overKεisπlogm1εε +O(1)asε→0. Then, integrating (3.1) overKεyields

πlog 1

mεε+|Kε| 2

|Kε|−hex

2+O(1) asε→0.

(6)

Since there aremε=1/|Kε|squares inK, we deduce Gε(vε, Bε)=π mεlog 1

mεε+1

2|2π mεhex|2+O(mε).

After expanding and replacinghex=Δex+ |logε|/2 we find Gε(vε, Bε)=2π2m2ε−2π mεΔex+1

2h2ex+o m2ε

.

Finally, using the fact thatmεΔex/2πand tends to+∞asε→0 we get Gε(vε, Bε)=1

2

h2exΔ2ex +o

Δ2ex

. (3.2)

4. Lower bound, identification of the limits

High fields. The case where Δexhex, or equivalently|logε| hex ε2 is particularly simple. In fact in this case one could even get more detailed information (see [21,22]). Indeed the existence of a test configuration satisfying (3.2) implies thatGε(uε, Aε)is smaller than the right-hand side of (3.2), which is itself o(h2ex). it follows in particular that hεhex2L2(K) is o(h2ex)and therefore, dividing byh2ex, thathε/ hex→1 inL2(K), hence also inL1(K). In particular

nεhex 2π ≈Δex

,

and of coursehε/(2π nε)→1 inL2(K). In fact in this case we even have strong convergence ofhε/(2π nε)inH1. Indeed, any critical point ofGεsatisfies the so-called second Ginzburg–Landau equation

−∇hε=jε, wherejε=(iuε,uεiAεuε)=ρε2(ϕεAε), (4.1) where the last equality is an alternative expression forjεwhenuε=ρεeεis not zero. A consequence of (4.1) is that pointwise|∇hε||∇uεiAεuε|. Therefore

1 2

K

|∇hε|2+ |hhex|2

Gε(uε, Aε),

and we are able to deduce as above, since the right-hand side is o(h2ex), thathε/ hex→1 inH1.

Low fields.We now deal with the rest of the cases in the theorem, namely case (1) withhex=O(|logε|)and cases (2) and (3). We thus assume the estimatehexC|logε|.

First, we recall the following construction which can either be adapted from [15] or [22], Theorem 4.1 to this periodic setting. We first define for anyr(0,1)thefree energyof(u, A)Hper1 as

Fr,ε(u, A)=1 2

K

|∇uiAu|2+r2h2+ 1 2ε2

1− |u|22

, (4.2)

with the notationh=curlA. In particular we have the following relation betweenFr,εandGε Gε(u, A)=Fr,ε(u, A)+1−r2

2

K

h2hex

K

h+1

2h2ex. (4.3)

Proposition 4.1.AssumeGε(uε, Aε)C|logε|2for anyε >0, where(uε, Aε)Hper1 . Then there existsε0>0and r0>0such that, for anyε < ε0and any1> rε1/3, the following holds.

There exists a collection of disjoint closed ballsB= {Bi}iI which is periodic with respect toKand finite in any compact subset ofR2. Moreover, denoting byB0a minimal subset ofBsuch thatB= τ∈LτB0,

(1) Denoting byr(B)the sum of the radii of balls belonging to a collectionB, we haver(B0)=r.

(7)

(2) Abusing notation,{||uε| −1|1/2} ⊂B. (3) WritingdB=deg(u, ∂B),

Fr,ε(uε, Aε, KB)π D

log r C

, (4.4)

whereD=

B∈B0|dB|is assumed to be nonzero andCis a universal constant.

(4) Finally,

DCFε(uε, Aε, K)

|logε| , (4.5)

whereCis a universal constant.

Finally, ifr1> r2andB1,B2are the corresponding families of balls, then every ball inB2is included in one of the balls ofB1.

The above result is adapted from [22], Theorem 4.1, applied withα=2/3. The proof of the equivalent of Theo- rem 4.1 in the periodic setting poses no difficulty, the way to do it is to consideruε as a section of a complex line bundle over the torusT =R2/L, whereLis the group of translations generated by(u, v), and Aε as a connection over this bundle. This is described somewhat in [4], Proposition 4.7. Note that we must impose a boundr < r0which is here to ensure that any ball of radiusr < r0onT is indeed a topological ball, it suffices to taker0smaller than the injectivity radius ofT.

We also note the following

Lemma 4.1.For any(u, A)Hper1 , if{Bi}iI is a finite collection of disjoint closed balls inR2such that|u|>0in K\ iBi and such that iBi does not intersect∂K, then

i

deg

u/|u|, ∂Bi

=

K

curlA.

Proof. This is standard. By periodicity, there exists two functionsf, g:R2→Rsuch that u(x+ u)=u(x)eif (x), A(x+ u)=A(x)+ ∇f (x),

and the same relations hold, replacingubyvandf byg.

The sum D of the degrees is the degree of u/|u| restricted to∂K. Letting ϕ be the phase of u, we thus have, denoting byτ the positively oriented unit vector tangent to∂K,

D=

∂K

τϕ= 1 0

1 u

d ds

ϕ(su)ϕ(su+ v) +

1 0

1 v

d ds

ϕ(u+sv)ϕ(sv) ds.

Then sinceϕ(x+ u)ϕ(x)=f (x)andϕ(x+ v)ϕ(x)=g(x)it follows that

D= 1 0

1 u

d

dsg(su) + 1 v

d dsf (sv),

and then, sinceA(x+ u)A(x)= ∇f (x)andA(x+ v)A(x)= ∇g(x), D=

∂K

A·τ =

K

curlA. 2

Let us now consider for anyε >0 the minimizer (uε, Aε)of Gε, and let hε=curlAε. For any r(0,1)we may construct using Proposition 4.1 a collection of ballsB of total radiusr. Moreover ifr is small enough, then a translationτ can be chosen so that∂(τ K)does not intersectB. We may then choose as the minimal subsetB0of

(8)

Proposition 4.1 the collectionBrconsisting of the balls inBwhich are included inτ K. The previous lemma combined with Proposition 4.1 then implies that

Fr,ε(uε, Aε,Br)π dε

log r dεεC

, 2π dε

τ K

hε=

K

hε=nε,

wheredε=

B∈Br|dB|, and the equality between the integrals overτ K andK follows from the periodicity of the configurations. Note that sinceFr,ε(uε, Aε)Ch2exandhexC|logε|, the a priori bound (4.5) reads

dεC|logε|.

we deduce from the above a lower bound forGε: Gε(uε, Aε|logε|nε−2π hexnε+1

2h2ex+1−r2 2

K

h2ε+Rε, (4.6)

where

Rε=π|logε|(dεnε)+π dε

log r dεC

. (4.7)

We now choose a fixed radiusr <1/4.SincedεC|logε|, it is easy to check, that ifdε>2nεandεis small enough, thenRε0 while ifnεdε2nε, then clearly

RεCnε(lognεCr), (4.8)

whereCr depends onr. Returning to the lower bound forGε, we writehex=Δex+ |logε|/2 and obtain Gε(uε, Aε)−2π Δexnε+1

2h2ex+1−r2 2

K

h2ε+Rε. (4.9)

Now we distinguish two cases.

ΔexC. This corresponds to cases (2) and (3) of the theorem. Combining (4.9) with the upper bound (2.5) and the estimate (4.8) gives

1−r2 2

K

h2εCnε(lognεCr)2π ΔexnεCnε (4.10)

but the integral of h2ε overK is bounded below by 4π2n2ε using Cauchy–Schwarz, and thus the above inequality implies that for somec, C >0 independent ofε,

cn2εCnε(lognε+1)2π ΔexnεCnε.

Thusnεis bounded independently ofε, and going back to (4.10)hεis bounded inL2(K)independently ofε.

Moreover, sincenε0, the above inequality implies that ifΔexis smaller than a certain, possibly negative value Δ1∈Rwhich could be expressed in terms of the constantsc, Cappearing in the inequality, thennε=0 ifεis small enough. Then the inequality

1 2

K

(hεhex)2Gε(uε, Aε)Gε(1,0)=1 2h2ex

implies thathε=0 and|uε| =1, thus(uε, Aε)is the Meissner solution, proving case (3) of the theorem.

1 ΔexC|logε|. In this case as in the previous one we first combine (4.9) with the upper bound (2.5) to obtain 1−r2

2

K

h2εCnε(lognεCr)2π Δexnε,

(9)

which allows to conclude thatnεexremains bounded asε→0, and then thathεexis bounded inL2(K)inde- pendently ofε. It remains to identify the limit of bothnεexandhεex. More precisely, from any sequence{εn}n

tending to zero we may extract a subsequence such that the limits exist, let us call themnandh. If we compute handnindependently of the particular subsequence, then we will have proved the convergence, and identified the limits. Since we have the relation

2π n=

K

h, (4.11)

it suffices, to finish the proof of the theorem, to prove thath=1.

First we may combine the more precise upper bound (3.2) with (4.9) and (4.8) to obtain 1−r2

2

K

h2εΔex

K

hεCnε(lognεCr)−1

2Δ2ex+o Δ2ex

.

Then dividing the above byΔ2exand lettingε→0 we find 1−r2

2

K

h2

K

h−1 2.

The left-hand side is bounded below by the minimum of the function−x+x2/2, i.e. by−1/2. It follows thath=1.

5. Improved convergence

To improve the convergence ofhεex(resp.hε) in case (1) (resp. case (2)) of the theorem, we invoke the classical Jacobian estimate of [16], together with a small-large ball type argument. Note that in the casehex |logε|, we have already proved theH1convergence, hence we assume below thathex< C|logε|.

We recall the London equation satisfied by critical points of the functionalGε:

Δhε+hε=με, whereμε=curljε+hε, (5.1)

where the superconducting currentjε is defined in (4.1).

The proof for the improved convergence is then the following. We construct for anyε >0 using Proposition 4.1 a collection of vortex balls with total radiusrε=ε13, which we denoteB, and we let

νε=2π

B∈B

dBδaB, (5.2)

whereaBdenotes the center ofB.

It then follows from the Jacobian estimate (see for instance [22], Theorem 6.2) that for anyβ(0,1), we have

εlim0ενε)=0, in the dual ofC0,β. Now we letdε=

B∈B|dB|, and we claim that

dε=O(nε), asεtends to 0. (5.3)

Assume a moment that this is true (sincenε=

B∈BdB, it amounts to proving that the balls have mostly positive degrees), then writing

με=νε+ενε),

we find that in the case 1 ΔexC|logε|and sincenε=O(Δex), the sequence{μεex}ε is bounded in the dual ofC0,β for anyβ(0,1). By compact embedding ofW1,q intoC0,β for anyq >2 and well chosenβ, we deduce that for anyp <2 we may extract a convergent subsequence inW1,p. Using (5.1), this gives the convergence of {hεex}ε inW1,p. The same argument holds, without normalizing byΔex in caseΔexis bounded independently ofε.

(10)

It remains to prove (5.3). This is done by inspecting carefully (4.6) and (4.7), lettingr=ε13 there, and comparing with the upper bound (2.5). These yield after some simplifications

−2π nεΔex+π dεnε

|logε| −dε

3 |logε| −dεlogdεCdε

0.

Since we have assumedΔexC|logε|, the above may be written π2

3dε|logε| −Cnε|logε| +o

dε|logε|

0,

which implies thatdε=O(nε)asε→0, and concludes the proof of theW1,pconvergence.

6. The case of a finite number of vortices

We now finish the proof of case (2) of the theorem, i.e. identifyhas a minimizer of some renormalized energy, in the terminology of [8].

The proof of this fact relies first on the construction of test configurations with prescribed vortices, a rather straight- forward task. The second part is to compute a precise expansion of the energy of a minimizer(uε, Aε)in terms of the vortex locations. This expansion can be found in [8] for the model without magnetic field and Dirichlet bound- ary condition, or in the case with magnetic field in [9], the definition of γ there is different from (2.4), but a result of Mironescu [19] shows the two are equivalent. The vortex locations in these references is to be understood as the limiting locations asε→0. A similar expansion may be found in [24] for configurations which are not necessarily minimizers of the energy. Another option is to give an expansion of the energy in terms of the actual vortex locations for fixedε. This is the approach used in [11] in the case without magnetic field or in [22] for the case with magnetic field. The latter approach gives information for eachε >0, but is less elementary than the former, that we adopt.

The proof we sketch for the convenience of the reader is borrowed from [22,9] and [8].

Upper bound.Givenp=(a1, . . . , an)P, the test configuration is constructed as follows. First we lethbe the solution of−Δh+h=2π

iδai inKwith periodic boundary conditions. We still denote byhthe extension ofhto R2by periodicity (note thathdoes not depend onε). Letting

A= {ai+ku+v|k, ∈Z, 1in}, we have−Δh+h=2π

a∈AδainR2.

ThenA is chosen such that curlA=h andϕ is defined modulo 2π inR2\Aand such that∇ϕA= ∇h.

Such aϕexists precisely because curl(A+ ∇h)= −Δh+his equal to 2π

a∈Aδa.Finally we define, for a fixed, arbitrarily chosenR >0,

ρε(x)=

1 ifx /a∈AB(a, Rε),

f (|xa|/ε)

f (R) ifxB(a, Rε),

wheref (r)= |u0|(r), andu0is the radial vortex which appears in (2.4), the definition ofγ.

Then we let vε =ρεe. It is clear that (vε, A) is K-periodic since the gauge-invariant quantities h=curlA, ρε= |vε|are and since∇ϕA= ∇h. It remains to evaluate the energy of(vε, A)overK. First we note that, since the integral ofhoverKis 2π n,

Gε(vε, A)=h2ex

2 −2π nεhex+Gε(vε, A), (6.1)

where

Gε(vε, A)=1 2

K

|∇vεiAvε|2+h2+ 1 2ε2

1− |vε|22

. (6.2)

We evaluateGε(vε, A). LetBi=B(ai, Rε)andKR=K\ iBi. OnKRwe have|vε| =1 and∇ϕA= ∇hfrom which it follows that the energy density there reduces to(|∇h|2+ |h|2)/2. Therefore, from the definition (2.3), we may write

εlim0

Gε(vε, A, KR)+π nlog(Rε)

=W (p)n

γ−2π Δex .

(11)

On the other hand, from [22], Chapter 10, we have

R→+∞lim lim

ε0

Gε

vε, A,

i

Bi

π nlogR

=0.

It follows that

εlim0

Gε(vε, A)+π nlogε

=W (p)+2π nΔex and therefore

εlim0

Gε(vε, A)h2ex 2

=W (p, n). (6.3)

Lower bound.Now we assume that(uε, Aε)is a minimizer ofGεand we try to compute a matching lower bound for (6.3).

First, from (5.1), we know thathεconverges inW1,pfor anyp <2 to the solution of

Δh+h=μ,

whereμis the common limit of{με}εand{νε}εasε→0. Then (5.2), (5.3) and the fact thatnεnimply thatμ is of the form 2πk

i=1diδai,wheredi∈Zand

idi=n. Note that the pointsai need not (yet) be distinct.

From the lower and upper bounds (4.6) and (2.5), we draw some further consequences, in view of (4.7). Indeed, in the lower bound (4.6), we have not used all the terms in the integrand ofGε. More precisely, we have only taken into account outside the ballsBr the term(hεhex)2/2. It follows that a by-product of the upper bound (2.5) is that, as ε→0,

1 2

K\Br

|∇uεiAεuε|2+ 1 2ε2

1− |uε|22

C.

Since this is true for anyr (the constantC then depending onr), we obtain in particular, lettingρε= |uε|and since

|∇uεiAεuε||∇ρε|, that

|∇ρε|2+ 1 2ε2

1−ρε22

is bounded inL1loc(K\ {a1, . . . , an}).

Letr0be half the minimal distance between two pointsai,aj, and fixr(0, r0). It follows from the above and from the strong convergence ofhεtohinW1,p, using a mean-value argument that for anyε >0, there existsr/2< rε< r such that for every 1inand lettingγi,ε=∂B(ai, ε),

γi,ε

|∇ρε|2+ 1 2ε2

1−ρε22

=O(1), hεhW1,pi,ε)=o(1), (6.4)

asε→0.

We bound from belowGε(uε, Aε, Bi,ε), whereBi,ε=B(ai, rε). To this aim we assume that we are in the Coulomb gauge. Then (see [9]),{Aε}εis bounded inW2,p(K)for anyp <2, hence inL. Moreover, since

|∇uεiAεuε|2|∇uε|2−2

Aε·jε+ |Aε|2|uε|2 , we deduce thatGε(uε, Aε, Bi,ε)is bounded below by

1 2

Bi,ε

|∇uε|2+(1− |uε|2)22

C

Aε2L2(Bi,ε)+ AεL4(Bi,ε)jεL4(Bi,ε) ,

wherejε=(iuε,uεiAεuε). But from (4.1) we havejε→ −∇hinLp for anyp <2. It follows easily (using also theW2,pbound forAε), that

Gε(uε, Aε, Bi,ε)1 2

Bi,ε

|∇uε|2+ 1 2ε2

1− |uε|22

δr,ε,

Références

Documents relatifs

The aim of this final section is to use the structure of the minimizers of the limiting functional F 0,g given by Theorem 4.3 to prove that minimizers of F ε,g 0 have the same

We consider a spherical superconductor in a uniform external field, and study vortices which appear near the lower critical field, the smallest value of the external field strength

For a sufficiently large applied magnetic field and for all sufficiently large values of the Ginzburg–Landau parameter κ = 1/ε, we show that minimizers have nontrivial

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

We find a vanishing gradient type property and then we obtain the renormalized energy by Bethuel, Brezis and Helein’s shrinking.. holes

We will also show that the solutions obtained minimize the original GL energy among all field configurations in the Coulomb gauge when the admissible space of the gauge vector

(1.14) By the dissipativity method (see [4], Section 11.5), (ii) implies exponential convergence to equilibrium for the second equation if X is fixed.. Whilst the coupling

The relation of the (possible) asymptotic behavior of solutions to (PGL) ε with motion by mean curvature, in Brakke’s (weak) formulation, has been shown in [4], under some