• 検索結果がありません。

2 Basic properties of the regularizable minimiz- ers

N/A
N/A
Protected

Academic year: 2022

シェア "2 Basic properties of the regularizable minimiz- ers"

Copied!
28
0
0

読み込み中.... (全文を見る)

全文

(1)

Asymptotic behavior of regularizable minimizers of a Ginzburg-Landau functional in higher

dimensions

Yutian Lei

Abstract

We study the asymptotic behavior of the regularizable minimizers of a Ginzburg-Landau type functional. We also dicuss the location of the zeroes of the minimizers.

1 Introduction

LetG⊂Rn (n≥2) be a bounded and simply connected domain with smooth boundary ∂G. Let g be a smooth map from ∂G into Sn1 satisfying d = deg(g, ∂G)6= 0. Consider the Ginzburg-Landau-type functional

Eε(u, G) = 1 p Z

G

|∇u|p+ 1 4εp

Z

G

(1− |u|2)2, (p >1)

with a small parameter ε > 0. It is known that this functional achieves its minimum on

Wp={v∈W1,p(G,Rn) :v|∂G=g}

at a functionuε. We are concerned with the asymptotic behavior ofuεand the location of the zeroes ofuεas ε→0.

The functionalEε(u, G) was introduced in the study of the Ginzburg-Landau vortices by F. Bethue, H. Brezis and F. Helein [1] in the casep=n= 2. Similar models are also used in many other theories of phase transition. The minimizer uεofEε(u, G) represents a complex order parameter. The zeroes ofuεand the module |uε| both have physics senses, for example, in superconductivity |uε|2 is proportional to the density of supercoducting electrons, and the zeroes ofuε

are the vortices, which were introduced in the type-II superconductors.

In the case 1< p < n, it is easily seen that Wg1,p(G, Sn1)6=∅. It is not difficult to prove that the existence of solutionupfor the minimization problem

min{ Z

G

|∇u|p :u∈Wg1,p(G, Sn1)}

Mathematics Subject Classifications: 35J70.

Key words: Ginzburg-Landau functional, module and zeroes of regularizable minimizers.

2001 Southwest Texas State University.c

Submitted December 13, 2000. Published February 23, 2001.

1

(2)

by taking the minimizing sequence. This solution is called a map of the least p-energy with boundary valueg. Using the variational methods, we can proved that the solutionup is also p-harmonic map onG with the boundary data g, namely, it is a weak solution of the following equation

−div(|∇u|p2∇u) =u|∇u|p.

Asε→0, there exists a subsequenceuεk ofuε, the minimizer ofEε(u, G), such that

uεk→up, in W1,p(G,Rn).

In the case p > n, Wg1,p(G, Sn1) = ∅. Thus there is no map of least p-energy on G with the boundary value g. It seems to be very difficult to study the convergence for minimizers of Eε(u, G) in Wp. Some results on the asymptotic behavior of the radial minimizers ofEε(u, G) were presented in [7].

Whenp=n, this problem was introduced in [1] (the open problem 17). M.

C. Hong studied the asymptotic behavior for the regularizable minimizers of Eε(u, G) inWn [6]. He proved that there exist{a1, a2, . . . , aJ} ⊂G,J ∈N and a subsequenceuεk of the regularizable minimizersuε such that

uεk

w un, inWloc1,n(G\ {a1, a2, . . . , aJ},Rn) (1.1) asεk →0, whereun is an n-harmonic map.

In this paper we shall discuss the asymptotic behavior for the regularizable minimizers of Eε(u, G) on Wn in the case p = n. Without loss of generality, we may assumed >0. Recalling a minimizer ofEε(u, G) on Wn be called the regularizable minimizer, if it is the limit of the minimizer of the regularized functional

Eετ(u, G) = 1 p Z

G

(|∇u|2+τ)p/2+ 1 4εp

Z

G

(1− |u|2)2, (τ∈(0,1)) onWn in W1,p. It is not difficult to prove that the regularizable minimizer is also a minimizer ofEε(u, G). In order to find the zeroes of the minimizers, we should first locate the singularities of the n-harmonic mapun.

Theorem 1.1 If aj ∈ G, j = 1,2, . . . , J are the singularities of n-harmonic map un, then J =d, the degree deg(un, aj) = 1, and {aj}dj=1 ⊂G. Moreover, for everyj, there exists at least one zero of the regularizable minimizeruεnear toaj.

Because the module of the minimizer has the physics sense, we have also studied its asymptotic behavior.

Theorem 1.2 Let uε be a regularizable minimizer of Eε(u, G),ρ=|uε|, then there exists a constant C independent ofε such that

Z

G

|∇ρ|n≤C, and 1 εn

Z

G

(1−ρ2)≤C(1 +|lnε|).

(3)

For any given η >0, denote Gη=G\ ∪dj=1B(aj, η), then asε→0, 1

εn Z

Gη

(1−ρ2)2→0, ρ→1, inCloc(Gη, R).

At last, we develop the conclusion of (1.1) into following

Theorem 1.3 There exists a subsequenceuεk of uε such that as ε→0, uεk→un, in Wloc1,n(G\ ∪dj=1{aj},Rn).

We shall prove Theorems 1.2 and 1.3 in§5 and§7 respectively, and the proof of Theorem 1.1 will be given in §6.

2 Basic properties of the regularizable minimiz- ers

First we recall the minimizer of the regularized functional Eετ(u, G) = 1

n Z

G

(|∇u|2+τ)n/2+ 1 4εn

Z

G

(1− |u|2)2, τ ∈(0,1) on Wn, denoted by uτε. As τ → 0, there exists a subsequence uτεk of uτε such that

τlimk0uτεk =uε, inW1,n(G,Rn), (2.1) and the limituεis one minimizer ofEε(u, G) onWn, which is named the regu- larizable minimizer. It is not difficult to prove thatuτε solves the problem

−div[(|∇u|2+τ)(n2)/2∇u] = 1

εnu(1− |u|2), onG, (2.2) u|∂G=g

and satisfies the maximum principle: |uτε| ≤1 on G. Moreover

Proposition 2.1 (Theorem 2.2 in [6]) For any δ > 0, there exists a con- stant C independent ofε such that

limτ0|∇uτε| ≤Cε1, onGδε, (2.3) where Gδε ={x∈G: dist(x, ∂G)≥δε}.

In this section we shall present some basic properties of the regularizable minimizeruε. Clearly it is a weak solution of the equation

−div(|∇u|n2∇u) = 1

εnu(1− |u|2), onG, (2.4) and it is known that|uε| ≤1 a.e. onG[6]. We also have

Proposition 2.2 For any δ > 0, there exists a constant C independent of ε such that

k∇uεkL(B(x,δε/8,Rn)≤Cε1, ifx∈Gδε.

(4)

Proof. Let y=xε1 in (2.4) and denote v(y) =u(x),Gε={y=xε1 :x∈ G}, Gδ = {y ∈ Gε : dist(y, ∂Gε) > δ}. Since that u is a weak solution of (2.4), we have

Z

Gε

|∇v|n2∇v∇φ= Z

Gε

v(1− |v|2)φ, φ∈W01,n(Gε,Rn).

Takingφ=vζn, ζ ∈C0(Gε, R), we obtain Z

Gε

|∇v|nζn≤n Z

Gε

|∇v|n1ζn1|∇ζ||v|+ Z

Gε

|v|2(1− |v|2n.

Setting y ∈ Gδ, B(y, δ/2) ⊂ Gε, and ζ = 1 in B(y, δ/4), ζ = 0 in Gε\ B(y, δ/2),|∇ζ| ≤C(δ), we have

Z

B(y,δ/2)

|∇v|nζn ≤C(δ) Z

B(y,δ/2)

|∇v|n1ζn1+C(δ).

Using Holder inequality we can derive R

B(y,δ/4)|∇v|n ≤C(δ). Combining this with the theorem of [9] yields

k∇vknL(B(y,δ/8))≤C(δ) Z

B(y,δ/4)

(1 +|∇v|)n ≤C(δ) which implies

k∇ukL(B(x,εδ/8))≤C(δ)ε1.

Proposition 2.3 (Lemma 2.1 in [6]) There exists a constantCindependent of εsuch that forε∈(0,1),

Eε(uε, G)≤d(n−1)n/2

n |Sn1||lnε|+C. (2.5) Proposition 2.4 There exists a constant C independent ofε such that

1 εn

Z

G

(1− |ue|2)2≤C. (2.6)

Proof. By (3.6) in [6], Z

G

|∇uε|n≥d(n−1)n/2|Sn1||lnε| −C.

Applying Proposition 2.3 we may obtain (2.6).

(5)

3 A class of bad balls

Fixρ >0. For the regularizable minimizeruε, from Theorem 2.2 in [6] we know

|uε| ≥ 1

2, onG\Gρε, (3.1)

where Gρε ={x ∈G : dist(x, ∂G) ≥ρε}. Thus there exists no zero of uε on G\Gρε.

Proposition 3.1 Let uε be a regularizable minimizer of Eε(u, G), There exist positive constants λ, µ which are independent ofε∈(0,1)such that if

1 εn

Z

GρεB2lε

(1− |uε|2)2≤µ, (3.2) where B2lε is some ball of radius2lεwithl≥λ, then

|uε| ≥ 1

2, ∀x∈Gρε∩B. (3.3)

Proof. First it is known that there exists a constantβ >0 such that for any x∈Gρεand 0< r≤1,

|Gρε∩B(x, r)| ≥βrn. Next we take

λ= min( 1 4C,1

8ρ), µ= βλn 16 where Cis the constant in Proposition 2.2.

Suppose that there is a pointx0∈Gρε∩B such that|uε(x0)|<1/2, then applying Proposition 2.2 we have

|uε(x)−uε(x0)| ≤Cε1|x−x0|=1

4, x∈B(x0, λε)∩Gρε. Hence

(1− |uε(x)|2)2> 1

16, ∀x∈B(x0, λε)∩Gρε, Z

B(x0,λε)Gρε

(1− |uε|2)2> 1

16|Gρε∩B(x0, λε)| ≥β 1

16(λε)n =µεn. (3.4) Since x0 ∈ B∩Gρε, we have (B(x0, λε)∩Gρε) ⊂ (B2lε∩Gρε), thus (3.4) implies

Z

B2lεGρε

(1− |uε|2)2> µεn which contradicts (3.2) and thus the proposition is proved.

To find the zeroes of the regularizable minimizer uε based on Proposition 3.1, we may take (3.2) as the ruler to distinguish the ball of radius λε which contain the zeroes.

(6)

Letλ, µ be constants in Proposition 3.1. If 1

εn Z

GρεB(xε,2λε)

(1− |uε|2)2≤µ,

thenB(xε, λε) is called good ball. OtherwiseB(xε, λε) is called bad ball. From Proposition 3.1 we are led to

|uε| ≥ 1

2, onGρε\ ∪xεΛB(xε, λε), (3.5) where Λ is the set of the centres of all bad balls. (3.5) and (3.1) imply that the zeroes ofuεare contained in these bad balls.

Now suppose that{B(xεi, λε), i∈I} is a family of balls satisfying (i)xεi ∈Gρε, i∈I

(ii)Gρε⊂ ∪iIB(xεi, λε) (iii)

B(xεi, λε/4)∩B(xεj, λε/4) =∅, i6=j . (3.6) LetJε={i∈I:B(xεi, λε) is a bad ball}.

Proposition 3.2 There exists a positive integer N which is independent of ε such that the number of bad balls cardJε≤N.

Proof. Since (3.6) implies that every point in Gρε can be covered by finite, say m (independent of ε) balls, from (2.6) and the definition of bad balls,we have

µεncardJε ≤ X

iJε

Z

B(xεi,2λε)Gρε

(1− |uε|2)2

≤ m Z

i∈JεB(xεi,2λε)Gρε

(1− |uε|2)2

≤ m Z

G

(1− |uε|2)2≤mCεn and hence cardJεmCµ ≤N.

Similar to the argument of Theorem IV.1 in [1], we have

Proposition 3.3 There exist a subsetJ ⊂Jε and a constant h≥λsuch that

iJεB(xεi, λε)⊂ ∪iJB(xεj, hε),

|xεi−xεj|>8hε, i, j∈J, i6=j. (3.7)

(7)

Proof. If there are two pointsx1, x2 such that (3.7) is not true withh=λ, we take h1 = 9λ and J1 =Jε\ {1}. In this case, if (3.7) holds we are done.

Otherwise we continue to choose a pair pointsx3, x4which does not satisfy (3.7) and takeh2= 9h1andJ2=Jε\ {1,3}. After at mostN steps we may conclude this proposition.

Applying Proposition 3.3 we may modify the family of bad balls such that the new one, denoted by {B(xεi, hε) :i∈J}, satisfies

iJεB(xεi, λε)⊂ ∪iJB(xεi, hε),

λ≤h; cardJ ≤cardJε, (3.8)

|xεi −xεj|>8hε, i, j∈J, i6=j.

The last condition implies that every two balls in the new family do not intersect.

Asε→0, there exist a subsequencexεik ofxεi and ai ∈Gsuch that xεik→ai, i= 1,2, . . . , N1= cardJ.

Perhaps there may be at least two subsequences converge to the same point, we denote by

a1, a2, . . . , aN2, N2≤N1

the collection of distinct points in{ai}N11.

To proveaj∈∂G, it is convenient to enlarge a littleG. AssumeG0⊂Rn is a bounded, simply connected domain with smooth boundary such that G⊂G0, and take a smooth map ¯g : (G0 \G) → Sn1 such that ¯g = g on ∂G. We extend the definition domain of every element in {u:G→Rn :u|∂G =g} to G0 such thatu=gonG0\G. In particular, the regularizable minimizeruεcan be defined onG0.

Fix a small constantσ >0 such that

B(aj, σ)⊂G0, j= 1,2, . . . , N2; 4σ <|aj−ai|, i6=j; 4σ <dist(G, ∂G0).

Writing Λj={i∈J:xεik→aj}, j= 1,2, . . . , N2, we have

iΛjB(xεik, hεk)⊂B(aj, σ), j= 1,2, . . . , N2

jJB(xεjk, hεk)⊂ ∪Nj=12 B(aj, σ/4) B(xεik, hεk)∩B(xεjk, hεk) =∅, i, j∈J, i6=j

as long asεk is small enough. Letuεis the regularizable minimizer ofEε(u, G) and denote dki =deg(uεk, ∂B(xεik, hεk)), ljk=deg(uεk, ∂B(aj, σ)), thus

lkj = X

iΛj

dki, d=

N2

X

j=1

lkj. (3.9)

To prove that the degrees dki and lkj are independent of εk, we recall a proposition stated in [6] (Lemma 3.3) or [2] (Theorem 8.2).

(8)

Proposition 3.4 Let φ:Sn1→Sn1 be aC0-map withdegφ=d. Then Z

Sn−1

|∇τφ|n1dx≥ |d|(n−1)(n1)/2|Sn1|.

Proposition 3.5 There exists a constant C which is independent of εk such that

|dki| ≤C, i∈J; |lkj| ≤C, j= 1,2, . . . , N2.

Proof. Sinceu=uεis a weak solution of (2.4), applying the theory of the local regularity in [9], we know u∈C(∂B(xεik, hεk)). Since (3.5) implies |u| ≥ 1/2 on∂B(xεik, hεk), thusφ= u

|u| ∈C(∂B(xεik, hεk), Sn1). From Proposition 3.4, we have

|dki| ≤ |Sn1|1(n−1)(1n)/2 Z

∂B(xεki ,hεk)

|( u

|u|)τ|n1. Since|u| ≥ 12 onG0\Gρε, there is no zero ofuεin it. Thus

deg(uεk, ∂B(xεik, hεk)) = deg(uεk, ∂(B(xεik, hεk)∩Gρεk)) and

|dki| ≤ |Sn1|1(n−1)(1n)/2 Z

∂[B(xεki ,hεk)Gρε]

|( u

|u|)τ|n1. (3.10) Substituting (2.3) and the fact|uεk| ≥12 on∂[B(xεik, hεk)∩Gρε] into (3.10), we obtain

|dki| ≤Cε1kn|Sn1|1(n−1)(1n)/2(hεk)n1≤C,

where C is a constant which is independent ofεk. Combining this with (3.9) we can complete the proof of the proposition.

Proposition 3.5 implies that there exist a number kj which is independent ofεk and a subsequence oflkj denoted itself such that

lkj →kj, as k→ ∞.

Since lkj, kj ∈N,{ljk} must be constant sequence for any fixed j, namely lkj = kj. The same reason shows dki can be writen as di which is also a number independent ofεk later.

4 An estimate for the lower bound

Write Ω0 =G0\ ∪Nj=12 B(aj, σ). Fixingj∈ {1,2, . . . , N2} and takingi0∈Λj, we havexi0 →aj as ε→0. Thus

iΛjB(xεi, hε)⊂B(xi0, σ/4)⊂B(aj, σ) (4.1) holds withεsmall enough.

Denote Ωj =B(aj, σ)\ ∪iΛjB(xεi, hε),Ω =B(xi0, σ/4)\ ∪iΛjB(xεi, hε).

To estimate the lower bound of k∇uεkLn(Ωj), the following proposition is nec- essary that was given by Theorem 3.9 in [6].

(9)

Proposition 4.1 Let As,t(xi) = (B(xi, s)\B(xi, t))∩G with ε≤t < s≤R.

Assume thatu∈Wg1,n(G,Rn)and12 ≤ |u| ≤1onAs,t(xi). If there is a constant C such that

1 εn

Z

As,t(xi)

(1− |u|2)2≤C.

Then for ε < ε0 there holds Z

As,t(xi)

|∇u|n≥ |di|n/(n1)(n−1)n/2|Sn1|lns t −C,

where C is a constant which is independent of ε and di is the degree of u on each ∂(B(xi, r)∩G), t≤r≤s.

Proposition 4.2 AssumeCardΛj =N. Then Z

j

|∇uε|n≥ Z

j,σ

|∇uε|n≥(n−1)n/2|Sn1||kj|lnσ

ε −C (4.2) where C is a constant which is independent ofε.

Proof. We give the proof following that in [6] (see Theorem 3.10), and the idea comes from [8]. Supposex1, x2, . . . , xN converge toaj, anddi,R(i= 1,2, . . . , N) is the degree of uε around ∂B(xi, R). Let Rσε denote the set of all numbers R ∈ [ε, σ] such that∂B(xi, R)∩B(xj, ε) = ∅ for all i 6= j and such that for some collection JR ⊂ {1,2, . . . , N}, satisfying JR ⊂JR0 ifR0 ≤R, the family {B(xi, R)}iJR is disjoint and

Ni=1B(xi, ε)⊂ ∪iJR0B(xi, R0)⊂ ∪iJRB(xi, R), R0≤R.

Note that Rσε is the union of closed intervals [Rl0,Rl],1 ≤l ≤L, whose right endpoints correspond to a number R=Rl such that∂B(xi, R)∩B(xj, R)6=∅ for some pairi6=j∈JR and whose left endpoints correspond to a number Rl0 such thatB(xi,Rl1)\ ∪jJ0B(xj, Rl0)6=∅fori∈JRl

0. JR=Jlis a constant for R ∈[Rl0,Rl] andJl+1 ⊂Jl, Jl+16=Jl. ThusL≤N. Moreover, there exists a constant M =M(h)>0 such that

Rl0≤M ε, RL ≥σ/M, Rl+10 ≤M Rl (4.3) for alll= 1,2, . . . , L−1. Finally, observe that for allR∈Rσε andJ ∈JR,

|kj|=| X

iJR

di,R| ≤ X

iJR

|di,R|n/(n1). (4.4) Applying (4.3)(4.4) and proposition 4.1 we have

Z

j,σ

|∇uε|n

L

X

l=1

X

iJl

| Z

ARl ,Rl

0

(xi)

∇uε|n

(10)

L

X

l=1

X

iJl

|Sn1|(n−1)n/2|di,Rl|ln(Rl/Rl0)−C

≥ |Sn1|(n−1)n/2|kj|X

l

(lnRl−lnRl0)−C

≥ (n−1)n/2|Sn1||kj|lnσ ε −C.

This and (4.1) imply that (4.2) holds.

Remark In fact the following results Z

j

|∇ uε

|uε||n≥(n−1)n/2|Sn1||kj|n/(n1)lnσ ε,

and Z

j

(1− |uε|n)|∇ uε

|uε||n≤C

had been presented in the proof of Theorem 3.9 in [6], where C which is inde- pendent ofε. Noticing

Z

j

|uε|n|∇uε

|uε||n = Z

j

|∇ uε

|uε||n− Z

j

(1− |uε|n)|∇ uε

|uε||n, we have

Z

j

|uε|n|∇ uε

|uε||n≥(n−1)n/2|kj|n/(n1)|Sn1|lnσ ε −C.

Theorem 4.3 There exists a constant C which is independent ofε, σ ∈(0,1) such that

Z

Nj=12j

|∇uε|n≥(n−1)n/2|Sn1|dlnσ

ε −C, (4.5)

1 n

Z

Gσ

|∇uε|n+ 1 4εn

Z

G

(1− |uε|2)2≤ 1

n(n−1)n/2|Sn1|dln 1

σ+C (4.6) whereGσ=G\ ∪Nj=12 B(aj, σ).

Proof. From (4.2) and Proposition 2.3 we have (n−1)n/2|Sn1|(

N2

X

j=1

|kj|) lnσ

ε ≤(n−1)n/2|Sn1|dln1 ε +C or (PN2

j=1|kj| −d) ln1ε ≤C. It is seen asεsmall enough

N2

X

j=1

|kj| ≤d=

N2

X

j=1

kj

(11)

which implies

kj ≥0. (4.7)

This and (3.9) imply

N2

X

j=1

|kj|=

N2

X

j=1

kj =d. (4.8)

Substituting (4.8) into (4.2) yields (4.5), and (4.6) may be concluded from (4.5) and Proposition 2.3.

From (4.6) and the fact |uε| ≤ 1 a.e. on G, we may conclude that there exists a subsequenceuεk ofuε such that

uεk

w u, W1,n(Gσ,Rn) (4.9) as εk→0. Compare (4.9) with (1.1) we knownu=un onGσ, and

{aj}Nj=12 ={aj}Jj=1. (4.10) These points were called the singularities ofun.

To show these singularitiesaj∈∂G, the following conclussion is necessary.

Proposition 4.4 Assumea∈∂Gandσ∈(0, R)with a small constant R. If u∈W1,n(AR,σ(a), Sn1)∩C0, u=g

on (G0\G)∩B(a, R)anddeg(u, ∂B(a, R)) = 1, then there exists a constantC which is independent of σsuch that

Z

AR,σ(a)

|∇u|n ≥21n(n−1)n/2|Sn1|ln1

σ−C . (4.11)

Proof. Similar to the proof of Lemma VI.1 in [1], we may writeGas the half space

{(x1, x2, . . . , xn) :xn>0} locally and aas 0 by a conformal change.

Denote St =∂B(0, t), t∈(σ, R). Noticing that g is smooth on G0\G, we have

sup

G0\G

|gτ| ≤C1. Takingtsufficiently small such that

t≤(n−1)1/2(2n1−1)1/(n1) 2C1

, then

Z

St

|g¯τ|n1≤ |St|C1n1≤ |Sn1|tn1C1n1≤(n−1)(n1)/2|Sn1|(1−21n) (4.12)

(12)

with R <1 small enough, where St =St∩ {xn <0}. On the other hand we can be led to

(n−1)(n1)/2|Sn1| ≤ Z

St

|uτ|n1= Z

St+

|uτ|n1+ Z

St

|g¯τ|n1 from Proposition 3.4. HereSt+=St\St . Combining this with (4.12) yields

Z

S+t

|uτ|n ≥ |St+|1/(n1)( Z

S+t

|uτ|n1)n/(n1) (4.1)

≥ 2n1|Sn1|(n−1)n/2t1. (4.2) Integrating this over (σ, R), we obtain

Z

AR,σ

|∇u|n≥21n|Sn1|(n−1)n/2lnR σ

which implies (4.11). To provekj = 1 for anyj, we supposeR >2σ is a small constant such that

B(aj, R)⊂G0; B(aj, R)∩B(ai, R) =∅, i6=j. (4.13) Denote Π ={v∈W1,n(Ω0, Sn1)∩C0: deg(v, ∂B(aj, r)) =kj, r ∈(σ, R), j = 1,2, . . . , N2}.

Proposition 4.5 For anyv∈Π, ifkj ≥0, j= 1,2, . . . , N2, then there exists a constant C=C(R)which is independent of σsuch that

Z

0

|∇v|n≥(n−1)n/2|Sn1|(

N2

X

j=1

k

n n−1

j ) ln1

σ−C. (4.14)

Proof. WriteAR,σ(aj) =B(aj, R)\B(aj, σ), thus∪Nj=12 AR,σ(aj)⊂Ω0. From Proposition 3.4 we have

kj=|kj| ≤ (n−1)(1n)/2|Sn1|1 Z

Sn−1

|vτ|n1

≤ (n−1)(1n)/2|Sn1|(n1)/n( Z

Sn−1

|vτ|n)(n1)/n namely

Z

Sn−1

|vτ|n≥(n−1)n/2|Sn1|kn/(nj 1). On the other hand, we may obtain

Z

0

|∇v|n

N2

X

j=1

Z

AR,σ(aj)

|∇v|n

(13)

N2

X

j=1

Z R

σ

Z

Sn−1

rn|∇τv|nrn1dζdr

≥ (n−1)n/2|Sn1|

N2

X

j=1

kjn/(n1) Z R

σ

r1dr

= (n−1)n/2|Sn1|(

N2

X

j=1

kn/(nj 1)) lnR σ which implies (4.14).

5 The proof of Theorem 1.2

Let uεbe a regularizable minimizer ofEε(u, G). Proposition 2.4 has given one estimate of convergence rate of |uε|. Moreover, we also have

Theorem 5.1 There exists a constantCwhich is independent ofε∈(0,1)such

that 1

εn Z

G

(1− |uε|2)≤C(1 + ln1

ε). (5.1)

Proof. The minimizer u =uτε of the regularized functional Eετ(u, G) solves (2.2). Taking the inner product of the both sides of (2.2) withuand integrating overGwe have

1 εn

Z

G

|u|2(1− |u|2) = − Z

G

div(v(n2)/2∇u)u

= Z

G

v(n2)/2|∇u|2− Z

∂G

v(n2)/2uun (5.2)

≤ Z

G

v(n2)/2|∇u|2+C Z

∂G

vn/2+C

where n denotes the unit outward normal to ∂G and un the derivative with respect to n.

To estimateR

∂Gvn/2, we choose a smooth vector fieldν such thatν|∂G=n.

Multiplying (2.2) by (ν· ∇u) and integrating overG, we obtain 1

εn Z

G

u(1− |u|2)(ν· ∇u) = − Z

G

div(v(n2)/2∇u)(ν· ∇u)

= Z

G

v(n2)/2∇u·(ν· ∇u)− Z

∂G

v(n2)/2|un|2. Combining this with

1 εn

Z

G

u(1− |u|2)(ν· ∇u) = 1 2εn

Z

G

(1− |u|2)(ν· ∇(|u|2))

= − 1

n Z

G

(1− |u|2)2divν

(14)

and Z

G

v(n2)/2∇u· ∇(ν· ∇u)

= Z

G

v(n2)/2|∇u|2divν+1 n

Z

G

ν· ∇(vn/2)

= Z

G

v(n2)/2|∇u|2divν+1 n

Z

∂G

vn/2− 1 n

Z

G

vn/2divν we obtain

Z

∂G

v(n2)/2|un|2≤ C 4εn

Z

G

(1− |u|2)2+C Z

G

vn/2+ 1 n

Z

∂G

vn/2. Thus

Z

∂G

vn/2 = Z

∂G

v(n2)/2(|un|2+|gt|2+τ)

≤ C Z

∂G

v(n2)/2+1 n

Z

∂G

vn/2+CEετ(uτε, G).

Substituting this into (5.2) yields 1

εn Z

G

|u|2(1− |u|2)≤CEετ(uτε, G).

Letτ→0, applying (2.1) and Proposition 2.3 we have 1

εn Z

G

|uε|2(1− |uε|2)≤CEε(uε, G)≤C(1 +|lnε|) which and (2.6) imply (5.1).

Theorem 5.2 Denoteρ=|uε|. There exists a constantCwhich is independent of ε∈(0,1)such that

k∇ρkLn(G)≤C. (5.3)

Proof. Denote u=uε. From the Remark in§4 we know Z

j

|u|n|∇ u

|u||ndx≥(n−1)n/2|kj|n−1n |Sn1|lnσ ε −C.

Thus we may modify (4.5) as Z

Nj=12j

ρn|∇ u

|u||n≥(n−1)n/2|Sn1|dlnσ ε −C.

Combining this with Z

Nj=12 j

|∇u|n ≥ Z

Nj=12j

ρn|∇ u

|u||n+ Z

Nj=12 j

|∇ρ|n−C

(15)

and Proposition 2.3, we derive Z

Nj=12j

|∇ρ|n≤C. (5.4)

On the other hand, from (2.1) and Proposition 2.1 we are led to Z

GρεB(xi,hε)

|∇uε|n= lim

τk0

Z

GρεB(xi,hε)

|∇uετk

|n ≤C(λε)n(C/ε)n≤C, fori∈Λj. Summarizing fori and using (5.4) we can obtain (5.3).

Theorem 5.3 For theσ >0 in Theorem 4.4, then as ε→0, 1

εn Z

G

(1−ρ2)2→0, (5.5)

where G=G\ ∪Nj=12 B(aj,3σ).

Proof. The regularizable minimizeruεsatisfies Z

Gσ

|∇u|n2∇u∇φ= 1 εn

Z

Gσ

uφ(1− |u|2), (5.6) where φ ∈ W01,n(Gσ,Rn) since uε is a weak solution of (2.4). Denoting u = uτε = ρw, ρ =|u|, w = u

|u| in Gσ and taking φ =ρwζ, ζ ∈ W01,n(Gσ,Rn), we have

Z

Gσ

|∇u|n2(w∇ρ+ρ∇w)(ρζ∇w+ρw∇ζ+wζ∇ρ) = 1 εn

Z

Gσ

ρ2ζ(1−ρ2). (5.7) Substituting 2w∇w=∇(|w|2) = 0 into (5.7), we obtain

Z

Gσ

|∇u|n2(ρ∇ρ∇ζ+|∇u|2ζ) = 1 εn

Z

Gσ

ρ2ζ(1−ρ2). (5.8) SetS={x∈Gσ:ρ(x)>1−εβ}for some fixedβ∈(0, n/2) andρ= max(ρ,1− εβ), thusρ=ρonS. In (5.8) takingζ= (1−ρ)ψ, whereψ∈C(Gσ, R), ψ= 0 onGσ\G,0< ψ <1 onG\G, ψ= 1 onG, we have

Z

Gσ

|∇u|n2ρ∇ρ· ∇ρψ¯ + 1 εn

Z

Gσ

l2(1−ρ2)(1−ρ)ψ¯ (5.9)

= Z

Gσ

|∇u|n2ρ∇ρ∇ψ(1−ρ) +¯ Z

Gσ

|∇u|nψ(1−ρ) Noticing 1/2≤l≤1 inGσ and applying (4.6) we obtain

1 εn

Z

G

(1−ρ)(1−ρ2) + Z

SG

|∇u|n2|∇ρ|2≤Cεβ. (5.10)

(16)

On the other hand, (2.6) implies ε|Gσ\S| ≤

Z

Gσ\S

(1−l2)2≤Cεn,

namely|Gσ\S| ≤Cεn. Then there exists a small constantε0>0 such that G ⊂S∪E

asε∈(0, ε0) whereE is a set, the measure of which converges to zero. Thus lim

ε0

Z

G

(1−ρ2)(1−ρ) = lim

ε0

Z

G

(1 +ρ)(1−ρ)2. By (5.10),

εlim0

1 εn

Z

G

(1 +ρ)2(1−ρ)2

≤ lim

ε0

2 εn

Z

G

(1−ρ)(1−ρ2) = 0 This is our conclusion.

Theorem 5.4 Assume B(x,2σ)⊂Gσ satisfies 1

εn Z

B(x,σ)

(1− |uε|2)2→0, asε→0, (5.11) then|uε| →1 inC(B(x, σ), R).

Proof. SinceB(x,2σ)⊂Gσ, there existsε0sufficiently small so thatB(x, σ)⊂ G2δε0. We always assumeε < ε0. Forx0∈B(x, σ), setα=|uε(x0)|. Proposi- tion 2.2 implies

|uε(x)−uε(x0)|< Cε1τ ε, ifx∈B(x0, τ ε),

whereτ= (1−α)(N C)1, Cis the constant in Proposition 2.2 andN is a large number such thatτ < δ. ThusB(x0, τ ε)⊂B(x, σ) and

|uε(x)| ≤α+Cτ, ifx∈B(x0, τ ε), Z

B(x0,τ ε)

(1− |uε(x)|2)2≥(1−1/N)2(1−α)n+2πεn(N C)n.

Combining this with (5.11) we obtain (1−α)n+2 =o(1) as ε→0. Thus it is not difficult to complete the proof of Theorem.

(17)

6 The proof of Theorem 1.1

It is known that the singularities ofun are inGfrom the discussion in§3. Since deg(g, ∂G) > 0, we can see that the zeroes of uε are also in G . Moreover, the zeroes are contained in finite bad balls, i.e. B(xεi, hε), i ∈ J. As ε → 0, B(xεi, hε)→aj, i∈Λj. This implies that the zeroes ofuεdistribute near these singularities of un as ε→0. Thus it is necessary to describe these singularities {aj}, j= 1,2, . . . , N2.

Proposition 6.1 kj = deg(un, aj).

Proof. Denote Ω0=G0\ ∪Nj=12 B(aj, σ). Combining (4.6) and Z

G0\G

|∇uε|n= Z

G0\G

|∇g¯|n≤C, we have

Z

0

|∇uε|n ≤C+ (n−1)n/2|Sn1|d|lnσ|, (6.1) whereC is a constant which is independent ofε. ForRin (4.13), from (6.1) we

have Z

AR,σ(aj)

|∇uε|n ≤C.

Then we know that there exists a constantr∈(σ, R) such that Z

∂B(aj,r)

|∇uε|n≤C(r)

by using integral mean value theorem. Thus there exists a subsequence uεk of uεsuch that

uεk →un, inC(∂B(aj, r)) as εk→0, which implies

kj= deg(uε, ∂B(aj, σ)) = deg(un, aj).

Proposition 6.2 kj = 0or kj= 1.

Proof. From the regularity results on n-harmonic maps (see [3][5] or [9]), we knowun∈C0(Gσ,Rn). Set

w=

g¯ onG0\G, un onGσ, thenw∈Π. Using Proposition 4.5 and (4.7) we have

Z

0

|∇w|n≥(n−1)n/2|Sn1|(

N2

X

j=1

k

n n−1

j ) ln1

σ−C(R). (6.2)

(18)

On the other hand, (6.1) and (4.9) imply uεk

w w, inW1,n(Ω0,Rn).

Noting this and the weak lower semicontinuity of R

0|∇u|n, applying (6.1) we have

Z

0

|∇w|n ≤limεk0 Z

0

|∇uεk|n≤(n−1)n/2|Sn1|dln 1

σ+C. (6.3) Combining this with (6.2), we obtain

(

N2

X

j=1

k

n n−1

j −d) ln1

σ ≤C or

N2

X

j=1

k

n n−1

j ≤d=

N2

X

j=1

kj

forσsmall enough. Thus (kj1/(n1)−1)kj ≤0 which implies that the Proposition holds.

Proposition 6.3 kj >0,j= 1,2, . . . , N2.

Proof. Supposek1= 0 andk2, k3, . . . , kN2>0. Similar to the proof of Theo- rem 4.3 we have

Z

Nj=22j

|∇uε|n≥(n−1)n/2|Sn1|dlnσ ε −C.

By this we can rewrite (4.6) as Z

G\∪Nj=22 B(aj,σ)

|∇uε|n+ 1 4εn

Z

G

(1− |uε|2)2≤C(σ).

Thus similar to the proof of Theorem 5.3 we may modify (5.5) as 1

εn Z

G\∪Nj=22B(aj,3σ)

(1− |uε|2)2→0 (6.4) asε→0. Noticing

G∩B(a1, σ)⊂G∩B(a1, R)⊂G\ ∪Nj=22 B(aj, R)⊂G\ ∪Nj=22 B(aj,3σ) we have

1 εn

Z

GB(a1,σ)

(1− |uε|2)2→0. (6.5) On the other hand, the definition ofa1implies that there exists at least one bad ballB(xε0, hε) such that

G∩B(xε0, hε)⊂G∩B(a1, σ).

Applying the definition of bad ball we obtain 1

εn Z

GB(a1,σ)

(1− |uε|2)2≥ 1 εn

Z

GB(xε0,hε)

(1− |uε|2)2≥µ >0 which is contrary to (6.5). This contradiction showsk1>0.

参照

関連したドキュメント

Meanwhile, in the scalar method [2–4, 14, 15, 28, 32, 33] the asymptotic behavior of solutions for scalar linear differential equations of Poincaré type is obtained by a change

We derive a high-order topological asymptotic expansion for a Kohn-Vogelius type functional with respect to the presence of a small obstacle inside the fluid flow domain.. An

Trujillo; Fractional integrals and derivatives and differential equations of fractional order in weighted spaces of continuous functions,

In section 3, we state some results on the large time behavior of subsonic solutions (for classical solutions in section 3.1 and for subsonic shock solutions in

On Landau–Siegel zeros and heights of singular moduli Submitted

Next, we prove bounds for the dimensions of p-adic MLV-spaces in Section 3, assuming results in Section 4, and make a conjecture about a special element in the motivic Galois group

Due to this we may also research the asymptotic behavior of minimizers of E ε (u, B) by referring to the p-harmonic map with ellipsoid value (which was discussed in [2]).. In

Our situation is different from the cases studied in [19] or [20], where they have considered the energy J with a ≡ 1 in a multiply connected domain without applied magnetic