• Aucun résultat trouvé

A classification of semilocal vortices in a Chern–Simons theory

N/A
N/A
Protected

Academic year: 2022

Partager "A classification of semilocal vortices in a Chern–Simons theory"

Copied!
21
0
0

Texte intégral

(1)

ScienceDirect

Ann. I. H. Poincaré – AN 33 (2016) 575–595

www.elsevier.com/locate/anihpc

A classification of semilocal vortices in a Chern–Simons theory

Jann-Long Chern

, Zhi-You Chen, Sze-Guang Yang

DepartmentofMathematics,NationalCentralUniversity,Chung-Li32001,Taiwan Received 5October2011;receivedinrevisedform 29August2013;accepted 25November2014

Availableonline 3December2014

Abstract

WeconsideraChern–SimonstheoryofplanarmatterfieldsinteractingwiththeChern–SimonsgaugefieldinaSU(N )global⊗ U(1)localinvariantfashion.Weclassifytheradiallysymmetricsolitonsolutionsofthesystemintermsoftheprescribedvalueof magneticfluxassociatedwiththismodel.Wealsoprovetheuniquenessofthetopologicalsolutioninacertaincondition.

©2014ElsevierMassonSAS.All rights reserved.

Résumé

NousétudionsunethéorieChern–SimonsdechampsdematièreplansinteragissantaveclechampdejaugedeChern–Simons d’unemanièreinvarianteparlegroupeSU(N )global⊗U(1)local.Nousclassonslessolutionssolitonsradialementsymétriquesdu systèmeenfonctiondelavaleurprescrited’unfluxmagnétiqueassociéàcemodèle.Nousprouvonségalementl’unicitédela solutiontopologiquesousunecertainecondition.

©2014ElsevierMassonSAS.All rights reserved.

MSC:35J60;35J55

Keywords:Chern–Simons–Higgsmodel;Classificationofnontopologicalsolutionsforellipticsystem;Uniquenessresultoftopologicalsolutions

1. Introduction

The theory of semilocal vortices has a global SU(N ) symmetry in addition to the local U(1) symmetry, SU(N )global⊗U(1)local, which is described by the Lagrangian:

L=κ

4μναFμνAα+(Dμφ) Dμφ

− 1 κ2|φ|2

|φ|2−12

,

Work supported by Ministry of Science and Technology, Taiwan (grants NSC101-2115-M-008-009-MY3 and MOST103-2115- M-008-011-MY3).

* Correspondingauthor.

E-mailaddresses:chern@math.ncu.edu.tw(J.-L. Chern),zhiyou@math.ncu.edu.tw(Z.-Y. Chen),sgyang@math.ncu.edu.tw(S.-G. Yang).

http://dx.doi.org/10.1016/j.anihpc.2014.11.007

0294-1449/©2014ElsevierMassonSAS.All rights reserved.

(2)

where Aμ, μ =0, 1, 2, is the gauge field on the (2 +1)-dimensional space R2,1 of metric tensor (gμν) = diag(1, −1, −1), Fμν=(∂/∂xμ)Aν(∂/∂xν)Aμis the corresponding gauge curvature tensor, Dμ=(∂/∂xμ) −iAμ

is the gauge covariant derivative, φ=1, φ2, . . . , φK) is K-component scalar field with φ=1, φ2, . . . , φK) (here ∗denotes complex conjugation), μνα is the totally skew-symmetric tensor with 012=1, and κ >0 is the Chern–Simons coupling constant. We mention that there are more interesting background and further studies of the corresponding Abelian Higgs model with SU(N )global⊗U(1)localsymmetry, e.g., in Refs.[2,7,8]and [13].

In the static case, when the Bogomol’nyi bound is saturated, we deduce the following self-dual equations:

(D1±iD2j=0, j=1, . . . , K, F12± 2

κ2|φ|2

|φ|2−1

=0. (1)

Without loss of generality, we will only take the upper (plus) sign into account. The flux corresponding to the magnetic field F12is given by

Φ=

R2

F12dx.

The system (1)is associated with one of the following boundary conditions that either (i) |φ(x)|→1 as |x|→ +∞, or

(ii) |φ(x)|→0 as |x|→ +∞.

The first is calledtopologicalboundary condition and the second is nontopological. The point vortices of the system are identified with zeros of φj, j=1, . . . , K. We remark that the existence of the topological as well as nontopological (multi-vortex) solutions has been proved by Chae in [2]. In the present article we set forth a classification of the radially symmetric solutions to the system.

We locate the zeros of φjat the origin with the multiplicities Nj(positive integers) for j=1, . . . , K. For simplicity, we set κ=2. With the substitution uj=log|φj|2, j=1, . . . , K, or equivalently

φj(z)=exp uj

2 +iNjArg(z)

, z=x1+ix2, (x1, x2)=x∈R2, the system (1)is reduced to the following semilinear elliptic system:

⎧⎪

⎪⎪

⎪⎪

⎪⎨

⎪⎪

⎪⎪

⎪⎪

uj=

K

k=1

euk K

k=1

euk−1

+4π Njδ0 inR2, j=1, . . . , K,

β=β(u)= 1 2π

R2

K

k=1

euk

1− K k=1

euk

dx(0,+∞),

(2)

where u =(u1, . . . , uK), ∂x22 1 +∂x22

2

and δ0denotes the Dirac delta function with point mass at the origin. Note that Φ=πβby definition. The topological and nontopological boundary conditions (i) and (ii) above are equivalent respectively to the following statements:

eujσj as|x| → +∞forj=1, . . . , K withσj≥0 and K j=1

σj=1; (3)

euj→0 as|x| → +∞forj=1, . . . , K. (4) For simplicity, through a rearrangement we may assume that N1N2≥ · · · ≥NK≥0, and for purely mathematical sake, that each Nj is not necessarily an integer. Our first consequence is stated as follows.

(3)

Theorem 1.1. Let u(r) =(u1, . . . , uK)(r)be a radially symmetric solution of Eq. (2)with r= |x|, and let α1=max

2(N1+1),4(NK+1)

, α2=max

4(N1NK),4(NK+1) .

Then βE:= {2N1} ∪1, +∞). Conversely, for a prescribed βE, the following statements are true:

(a) If β=2N1, then uis topological and can be characterized by uj(r)=v(r)+2Njlogr+logcj, cj>0, j=1, . . . , K, where v(|x|) Cloc2 (R2)is a solution of

⎧⎪

⎪⎩ v+1

rv=f ev

f ev−1

, f (r)= K j=1

cjr2Nj, r >0, v(r)= −2N1logr+O(1) asr→ +∞.

(5)

(b) If β2, +∞)then uis nontopological and can be characterized by uj(r)=v(r)+2Njlogr+logcj, cj>0, j=1, . . . , K, where v(|x|) Cloc2 (R2)is a solution of

⎧⎪

⎪⎩ v+1

rv=f ev(f ev−1), f (r)= K j=1

cjr2Nj, r >0, v(r)= −βlogr+O(1) asr→ +∞.

(6)

Moreover, α1=α2=4NK+4when N1≤2NK+1.

(c) If N1>2NK+1, then there exists a constant α3=α3(N1, . . . , NK) 1, α2)such that (2)possesses a nontopo- logical solution (u1, . . . , uK)which is characterized in (b) for each β∈ [α3, ). Moreover, β(u) ∈ [α3, )for any nontopological solution u =(u1, . . . , uK)of(2).

Remark 1.1. As shown in Appendix A, each solution of (2) satisfying the (topological) boundary condition (3)is automatically radially symmetric. So that the radial symmetry premise can be removed from Theorem 1.1for the topological solutions.

Remark 1.2. In the case N1>2NK+1, it turns out that α1=2(N1+1)and α2=4(N1NK)with α1< α2. Remark 1.3. When K=1, the semilocal system (2) is reduced to the well-known Abelian Chern–Simons Higgs model

⎧⎪

⎪⎨

⎪⎪

u=eu eu−1

+4π N δ0 inR2, β=β(u)= 1

R2

eu 1−eu

dx(0,+∞), (7)

in which the sufficient and necessary condition for Eq.(7)possessing a radially symmetric solution is that β∈ {2N}∪ (4N+4, +∞). This specializes the consequence of Theorem 1.1.

In the case K=1 we remark that radial solutions of (7)depend uniquely on the values of β; please see e.g. [5]and [4]for the such uniqueness results. In the case K≥2, there can be two solutions u1=u2such that β(u1) =β(u2). In fact, as shown at the beginning of Section 2, any solution u =(u1, . . . , uK)of Eq. (2)can be written as

uj(r)=v(r)+2Njlogr+logcj, cj>0, j=1, . . . , K, where v=v(β, c1, . . . , cK)satisfies the equation

(4)

⎧⎪

⎪⎩ v+1

rv=f ev

f ev−1

, f (r)= K j=1

cjr2Nj, r >0, v(r)= −βlogr+O(1) asr→ +∞.

(8)

In other words, u depends on β and the K-tuple parameter τ =(c1, . . . , cK). Let β be fixed. Assume that v= v(c1, . . . , cK)and v˜=v(c˜1, . . . , ˜cK)are solutions of (8)corresponding parameters (cj)Kj=1and (c˜j)Kj=1respectively.

So we have two solutions u, ˜uof (2)given by uj(r) =v(r) +2Njlogr+logcj and u˜j(r) = ˜v(r) +2Njlogr+logc˜j, j =1, . . . , K. If u = ˜u, then there must be a constant λ >0 such that

logcj

˜

cj =v− ˜v=λ, j=1,2, . . . , K.

Consequently, we may select two different parameters (cj)Kj=1, (c˜j)Kj=1on the spherical region Σ=

(c1, . . . , cK):c21+ · · · +cK2 =1 withc1, . . . , cK>0

, (9)

so that u = ˜u. Although a solution of (2)cannot be uniquely determined by any prescribed βin the case K≥2, the following theorem shows that a solution of (5)is unique under a certain condition. Such a result represents some kind of uniqueness for the model. We state the theorem as follows:

Theorem 1.2. If N1NK≤1, then for any given cj>0, j=1, . . . , K, the solution of (5)is unique. Moreover, the topological solution of (2), which is denoted by uτ,2N1=uτ, is uniquely determined by the parameter τ belonging to the region Σgiven in (9).

By the way, we remark that Theorem 1.2can be equivalently characterized from an alternative viewpoint by means of the initial value at the origin. Since all topological solutions of (2) are radially symmetric with respect to the origin as we have shown in Appendix Aof the present paper, we may rewrite the system (2)as the following system associated the initial value at the origin:

uj(r)+1 ruj(r)+

K

j=1

euj

1− K j=1

euj

=0, r >0, (10)

with behavior at the origin that

uj(r)=2Njlogr+ρj+o(1) asr→0+. (11)

According the uniqueness result in Theorem 1.2, we obtain the following assertion about the structure of topological solutions of (10)in terms of (ρ1, · · ·, ρK).

Corollary 1.3. Suppose that N1NK ≤ 1. Then for any =(1, . . . , K1) ∈ RK1, there exists a unique 1(), . . . , ρK()) ∈RK such that (u11, . . . , ρK), . . . , uK1, . . . , ρK)) is a topological solution of (10) with ρj() ρ1() =j1for j=2, . . . , K.

The paper is organized as follows. In Section2we show that the solutions of (2)can be reduced to a single profile equation, and furthermore make an analysis on its solution structure. In addition, we carry out a uniqueness result under a certain condition. In Sections 3and 4we exhibit the existence of the nontopological solutions with respect to prescribed values of magnetic flux, in order to detect an optimal lower bound of the possible values. Finally, in Appendix Aof the present paper, we include a proof about the radial symmetry for the topological solution of the system(2).

2. Analysis of the solution structure

In view of the K-component system (2), we see that (uiuj) =4π(NiNj0for any i, j ∈ {1, . . . , K}, which indicates that

(5)

ui(r)=uj(r)+2(NiNj)logr+hij(r),

where hij is a harmonic function. Since β <+∞, we have limr→∞rui(r) =2Niβ for i=1, . . . , K and it is possible to extract a positive constant Csuch that

u1(r)+ · · · +uK(r)Clog(1+r), r >0.

It follows that hij is in fact a constant. For the sake, through a simple substitution that

uj(r)=v(r)+logcj+2Njlogr, (12)

where cj>0 for j =1, . . . , K, we obtain a further reduction of the system (2), namely

⎧⎪

⎪⎪

⎪⎨

⎪⎪

⎪⎪

v+1

rv=f ev

f ev−1

, r >0, β=

0

rf ev

1−f ev

dr(0,+∞),

(13)

where f (r) =K

j=1cjr2Nj with N1N2≥ · · · ≥NK≥0. We set off a structure analysis for the solutions of (13).

First of all, for any given K-tuple (c1, c2, . . . , cK)with cj>0, we consider the solution v(r) =v(r; s) solving the initial value problem

v+1

rv=f ev

f ev−1 , v(0)=s, v(0)=0, s∈R.

(14) The solution of (14)can be classified into three types; in fact, we have:

Theorem 2.1. Let vbe a solution of (14). Then vbelongs to one of the following classes given in the sense that 0-Type: f (r)ev(r)→0 asr→ +∞;

1-Type: f (r)ev(r)→1 asr→ +∞;

∞-Type: v(r)blows up at a finiter. (15) Before proving Theorem 2.1, we introduce properties that 2NKrf/f≤2N1and that

rf f

(r)= 1 f2(r)

i<j

4Ni(NiNj)cicjr2(Ni+Nj)1≥0. (16) They can be obtained by simple calculations. Note that the properties are independent of the choice of (c1, c2, . . . , cK).

So in the following content we assume c1=c2= · · · =cK=1 for simplicity.

Lemma 2.1. Let vsolve the problem (14). Then the function f ev−1cannot attain a nonnegative local maximum at any finite r >0.

Proof. Assume g=f ev−1 has a nonnegative local maximum at r=r1. Then g(r1) =0 and there is a small >0 such that (rv)≥0 on [r1, r1+). Note that

g(r)=f ev

r h(r), whereh(r)=rf f +rv. Hence h(r1) =r1g(r1)/f (r1)ev(r1)=0, and furthermore,

h(r)= rf

f

+ rv

≥0

on [r1, r1+). It follows that h(r) ≥0 on [r1, r1+). Therefore, g(r) ≥0 on [r1, r1+). This contradicts the assumption that g(r1)is a local maximum. 2

(6)

Lemma 2.2. Let u1(r) =2N1logr+v(r)where vsolves (14)in an interval I= [0, ω), ω∈R ∪ {∞}. If u1(r1) <0 for some r1I, then vmust be of the 0-Type.

Proof. It suffices to show that f ev<1 for any ras far as v(r)is defined. In fact, since u1is positive near r=0 and negative at r=r1, the function u1attains its maximum sup0<r<r1u1(r)at some locus r0(0, r1). From (14),

2N1

r1

0

rf ev

1−f ev

dr=r1u1(r1) <0. (17)

Applying the fact u1(r0) =0 and using r0in substitution for r1in the equality (17), we see that the derived integral over (0, r0)has value 2N1; this indicates

r1

r0

rf ev

1−f ev

dr >0. (18)

We conclude f ev<1 in (0, r1), because, by Lemma 2.1, once f (r0)ev(r0)≥1, the function f (r)ev(r)−1 is positive for r > r0and thus (18)cannot happen. Assume f (r)e¯ v(r)¯ =1 for some r¯≥r1with the property f (r)ev(r)<1 for r(0, r)¯ and f (r)ev(r)>1 for r >r¯. Then u1(r) <0 whenever r0< r <r¯and therefore f evis strictly decreasing on the interval (r0, ¯r). Since f ev=1 at r= ¯r in our assumption, we come to the conclusion that f (r0)ev(r0)>1, a contradiction. 2

Proof of Theorem 2.1. From Lemma 2.1, if f (r0)ev(r0)≥1 for some r0, then f ev is increasing in the interval (r0, +∞)and then v≥λf evfor r≥r0where λis a positive constant. Hence vblows up at a finite r(cf. e.g. [12]).

Otherwise, vis an entire solution of (14)satisfying f ev<1 all the time; in the meanwhile vis a decreasing function which approaches to −∞as r→ +∞. Note that in the last case we have

β= 0

rf ev

1−f ev

dr <+∞. (19)

For, if the integral β= +∞, it is possible to select r1>0 such that r1v(r1) ≤ −4N1−4. By Eq.(14), rv(r)=r1v(r1)

r r1

rf ev

1−f ev

dr≤ −4N1−4, rr1,

implying v(r) ≤ −(4N1+4) logr+C for large r; this yields that β <+∞, a contradiction. Now by the fact that f ev<1 for the entire solutions, there are only two possibilities that either (i) r2N1ev(r) increases for all r >0, or (ii)r2N1ev(r)decreases in an interval (r0, +∞)for some r0>0. In the case (i), we have β=2N1and f (r)ev(r)→1 as r→ +∞where we note that if f (r)ev(r)c <1, it will contradict (19). In the case (ii), from Lemma 2.2we have f (r)ev(r)→0 as r→ +∞, which indicates that β >2N1+2 by (19). By the way, in either case of the entire solution, we have

v(r)= −βlogr+O(1), r→ +∞. (20)

The classification (15)is concluded. 2

Taking account of the initial value sfor the solution v(r) =v(r; s)of (14)depending on s, we set E(r, s)=

rv(r;s)2

r2f2(r)e2v(r;s)+2r2f (r)ev(r;s).

Multiplying (14)by the factor rvand taking integration both sides, we obtain the following variational identity:

1

2E(r, s)= r 0

tg(t)f (t)ev(t;s)

1−f (t )ev(t;s) dt+

r 0

tf2(t)e2v(t;s)dt, (21)

(7)

where g(t) = [tf(t)/f (t)] +2. We define Ji=

s∈R:v(r;s)is of thei-Type according to(15)

, i=0,1,∞.

Clearly, R=J0J1J. E(r, s) >0 near r=0 for any s∈R. We remark that s∈Jif and only if E(r0, s) =0 for some r0>0. In fact, if sJ0J1, then

∂E

∂r r, s

=rg(r)f (r)ev(r;s)

1−f (r)ev(r;s)

+rf2(r)e2v(r;s) (22) is positive for all r >0, where we apply the fact that f ev<1. Thus E(r, s)is positive whenever r >0. On the other hand, letting sJ, it is not hard to see that E(r, s) → −∞as r→ωin the maximal existence interval Imax= [0, ω)of the solution v(r; s). Hence E(r, s)must vanish at some site r=r0>0. In addition, we conclude the following lemma.

Lemma 2.3. Both J0and Jare open sets. Moreover, J0J1is bounded from above.

Proof. The assertion of the lemma about J0can be concluded readily from Lemma 2.2by means of the continuous dependence of v(r; s)on the parameter s. We omit the details and straightforwardly account for the set J. To show that Jis open, we take advantage of the variational identity (21). Let s0J. Then there exists r0>0 such that E(r0, s0) =0. By definition, we have

r0v(r0)2

+r02f (r0)ev(r0)

2−f (r0)ev(r0)

=E(r0, s0)=0, v(r)=v(r;s0), implying that f (r0)ev(r0)>2. Hence by a rearrangement in the formula (22),

∂E

∂r(r, s0)=2r02f(r0)ev(r0)

1−f (r0)ev(r0)

+2r0f (r0)ev(r0)

2−f (r0)ev(r0)

<0.

Using the Implicit Function Theorem, r can be represented as a C1-function of s in a neighborhood I(s0) = (s0, s0+)of s0, with a small >0, such that r(s0) =r0and

E r(s), s

=0, sI(s0).

Therefore I(s0) J. As for the remaining part of the lemma, we first show that the set J0 has an upper bound.

Assume contrarily that there exists a sequence of real numbers sjJ0such that sj→ +∞as j→ +∞. Through an adaptation of the device in [4], we let

wj(r)=v

esj/(2NK+1)r;sj

sj.

Assume N1> N2>· · ·> NK (without loss of generality). Then, by a direct computation from (14), wj solves the equation

⎧⎨

wj(r)+1

rwj(r)=G2j(r) e2wj(r)esj/(2Nk+1)Gj(r) ewj(r), r >0, wj(0)=0, wj(0)=0,

(23) where

Gj(r)= K m=1

e2sj(NmNK)r2Nm.

Note that v(r;sj)is decreasing in r; wjis thus decreasing and the right hand side of Eq. (23)converges uniformly on any compact interval. By passing to a subsequence, it follows that wjapproaches to a function w, which solves˜

w˜(r)+1

rw˜(r)=r4NKe2w(r)˜ , r >0,

˜

w(0)=0, w˜(0)=0.

(24) This is impossible because any solution of (24)is increasing and blows up at finite r. So that J0 is bounded from above. The argument for sjJ1is the same. We omit it here. 2

(8)

To make out how the solution vdepends on s, we consider the function ϕ(r) =ϕ(r; s) =(∂v/∂s)(r; s). Directly taking derivative on Eq.(14), ϕsolves the following equation

ϕ(r)+1

(r)=f (r)ev(r)

2f (r)ev(r)−1

ϕ(r), v(r)=v(r;s), ϕ(0)=1, ϕ(0)=0.

(25)

Lemma 2.4. The functions ϕ(r;s)and (∂ϕ/∂r)(r;s)are continuously dependent on sfor any fixed r.

Proof. Let s∈R. Consider the function w(r) =w(r;s) =ϕ(r; s) −ϕ(r;s)for |ss|≤δwith a small δ >0. To conclude the lemma, it suffices to show that w(r;s) →0 as s→sfor any fixed r. In fact, by a direct computation from (25),

w(r)+1

rw(r)=P (r;s)w(r)+Q(r;s),

where P , Qare functions of r, depending on sand given by P (r;s)=f (r)ev(r;s)

2f (r)ev(r;s)−1 , Q(r;s)=f (r)

2f (r)

ev(r;s)+ev(r;s)

−1

ev(r;s)ev(r;s) ϕ

r;s . So that wsatisfies the integral equation

w(r)= r 0

t

logr t

P (t;s)w(t) dt+ r 0

t

logr t

Q(t;s) dt.

Therefore, by the continuity of v(r; s), we have w(r;s)

r 0

tQ(t;s)logr t dt

exp

r 0

tP (t;s)logr t dt

C r 0

Q(t;s)dt

C sup

t∈[0,r]

v(t;s)v

t;s→0 as s→s. This concludes the proof. 2

Lemma 2.5. If s0J0, then there exist numbers C, R, >0such that

ϕ(r;s)Clogr (26) for all rRand sJ0∩ {s∈R: |ss0| < }.

Proof. Let s0J0. We show that ϕ0(r) =ϕ(r; s0)does not change sign for rsufficiently large. In fact, by comparison ϕwith the function wc(r) =rv(r) +c, here v(r) =v(r; s0)and cbeing a constant, we have

r

ϕ0wcwcϕ0r2

r1=

r2

r1

ϕ0 rwc

wc 0

dr

=

r2

r1

r 2f ev

f ev−1

+gcf ev

2f ev−1

ϕ0dr, (27)

(9)

where gc(r) = [(rf(r))/f (r)] −c. Since rv(r)vanishes at r=0 and decreases to −β with β=β(s0) >2N1+2 for the 0-Type solution, there can be λ ∈(2N1+2, β) such that wλ(rλ) =0 for some rλ>0, whereas wλ>0 in the interval (0, rλ)and wλ<0 in (rλ, +∞). If ϕ0changes sign no matter how large r is, it is possible to select two successive zeros r1, r2> rλof ϕ0such that r1< r2, ϕ0(r1) >0, ϕ0(r2) <0 and ϕ0>0 in between. Note that gλ<−2 and from (20)we have f evr2σ near infinity, σ >0. Hence the left hand side of (27)is positive, but the right hand side is negative. This is a contradiction. Now we consider the case ϕ0>0 in (r0, +∞)for some large r0; the argument for its negative counterpart is similar. From the expression

0(r)=r0ϕ0(r0)+ r r0

tf (t)ev(t )

2f (t )ev(t )−1

ϕ0(t) dt,

we have r0ϕ0(r0) −Crσ0(r) ≤r0ϕ0(r0)for r > r0and some positive constant C, where we use the fact that f ev1 in (r0, +∞)for r0sufficiently large. Thus, there is C1>0 such that

ϕ0(r)C1logr, rr0.

To conclude the lemma, we need to show that there exist σ, c0>0 such that f (r)ev(r;s)c0r2σ

for any sin a neighborhood of s0and rsufficiently large. In fact, taking σ >0 so small that

β=β(s0):=

0

rf (r)ev(r;s0)

1−f (r)ev(r;s0)

dr >2(N1+1)+3σ, we have

rv(r;s0)= −β+ r

tf (t)ev(t;s0)

1−f (t )ev(t;s0) dt

<−2(N1+1)−3σ+c1r2N1+2β

<−2(N1+1)−2σ, rr1

for some r1r0. Assume J0(s00, s0+0)for some 0>0. By the continuity of v(r; ·)as a function of s, it is possible to extract a small 1(0, 0)such that

rv(r;s)rv(r1;s) < rv(r1;s0)+σ <−2(N1+1)−σ,

and hence v(r; s) ≤c2(2N1+2 +σ ) logrfor all |ss0| < 1and r≥r1. This concludes the proof. 2 Theorem 2.2. The function

β(s)= 0

rf (r)ev(r;s)

1−f (r)ev(r;s) dr

is differentiable on J0and supsJ0β(s) = +∞.

Proof. The differentiability of β(s)follows readily from Lemma 2.4by a direct computation. To see the unbounded- ness of β, we pick s0J0and define s=sup{s: [s0, s) J0}. By Lemma 2.3, sJ1. We show that β(s)goes to infinity as s→s. As a matter of fact, if we consider the sequence vj(r) =v(r;sj)with sj(s0, s)and sjsas j→ +∞, from the variational identity (21), we have

(10)

1

2β2(sj)= 0

tg(t)f (t)evj(t)

1−f (t )evj(t) dt+

0

tf2(t)e2vj(t)dt

0

tf2(t)e2vj(t)dt.

Assume β(sj)is bounded. Then rf2(r)e2v(r;s)L1(R). This contradicts the fact that f (r)ev(r;s)→1 as r→ +∞.

So that β(s)is not bounded. 2

Lemma 2.6. Assume N1NK≤1. If sJ1, then there exists μ >0such that ϕ

r;s

μ

for all r >0. Moreover, limr→+∞ϕ(r;s) → +∞.

Proof. Note first that, from Theorem 2.1, the function rv(r; s) is strictly greater than −2N1 for all r >0 and approaches to −2N1as r→ +∞. So that the function

w(r)=rv(r)+2N1, v(r)=v r;s

,

is positive all the time and w(r) →0 as r→ +∞. Applying the identity (27)between the functions w(r)and ϕ(r) = ϕ(r;s)with the substitution r1=0, r2=Rand c=2N1, gives

R

ϕ(R)w(R)w(R)ϕ(R)

= R 0

ϕ rw

w

dr

= R 0

r

(gc+2)f ev

f ev−1

+gcf2e2v

ϕ dr <0 (28)

for all R >0, where we make use of the fact that −2 ≤gc≤0 for c=2N1. Applying (28), we show that ϕ >0 as follows. In fact, assuming contrarily that ϕ has the first zero at r=σ >0 and letting R=σ, it follows that the left hand side (LHS) of the identity (28)is positive while the right hand side (RHS) is negative. This contradiction indicates that ϕ(r)must be positive for all r >0. To conclude the proof, it suffices to show that ϕis not a bounded function. Assume contrarily that ϕis bounded. Then from (25)it is not hard to see that

r→+∞lim (r)=0.

Now applying (28)again and letting R→ +∞, the LHS of (28)approaches to 0, while the RHS approaches to a negative number; that is a contradiction. Hence the function ϕis unbounded and thus from (25)we note, in addition, that ϕcannot oscillate near infinity. So that ϕ(r) → +∞as r→ +∞. Therefore, ϕhas a positive lower bound. 2 Remark 2.1. For sJ1, from Lemma 2.5and (25)it follows that

ϕ(r)=ϕ(r0)+ r r0

t

logr t

f (t )ev(t )

2f (t )ev(t )−1 ϕ(t ) dt

ϕ(r0)+1 2

r r0

t

logr t

ϕ(t ) dt, rr0,

for r0>0 sufficiently large, where v(r) =v(r;s)and ϕ(r) =ϕ(r;s). Thus, we may conclude further that ϕ(r;s) grows exponentially near infinity; in fact, the inequality

ϕ r;s

c1exp c2r2

holds true in a neighborhood of +∞for which c1, c2>0 are suitable constants.

(11)

Proof of Theorem 1.2. From Lemma 2.3, it is clear that the set J1is nonempty. To conclude the theorem, we are going to show that J1consists of a single point. Let sJ1. In view of Lemmas 2.4 and 2.6, as well as the continuity of v(r; s), it is possible to select r0>0 and δ >0 such that

(a) 2f (r)ev(r,s)>1 for r≤r0;

(b) 2f (r0)ev(r0,s)>1 and ϕ(r0; s) >0 for s∈Iδ=(sδ, s+δ);

(c) ϕ(r; s) ≥λ >0 for (r, s) ∈ [0, r0] ×Iδ.

We set Iδ+=(s, s+δ)and Iδ=(sδ, s). Letting [0, ωs)with ωs∈R∪{+∞}be the maximal existence interval for the solution v(·) =v(·;s)that solves (14), we have the following conclusion:

Claim 1. ϕ(r, s) ≥λand 2f (r)ev(r,s)>1for all sIδ+and r∈ [r0, ωs).

Assume our assertion fails. Then by the statements (b) and (c), we may extract sIδ+and r1(r0, ωs)such that ϕ(r; s) >0 for r∈ [r0, r1)and ϕ(r1; s) =0. Thus, from the statement (a), we have 2f (r)ev(r,s)>1 for r∈ [r0, r1).

So that 0=r1ϕ

r1;s

=r0ϕ r0;s

+

r1

r0

f (t )ev(t;s)

2f (t )ev(t;s)−1 ϕ

t;s

dt >0, (29)

which is a contradiction. So we conclude Claim 1. Since f (r)ev(r;s0)→0 as r→ +∞for any s∈J0, it follows from Claim 1that Iδ+J0is empty. Now we continue to the following claims.

Claim 2. If there exist two points s1, s2J1Iδ, then there must be a point sJwhich lies between s1and s2. To see this, we assume, in contrast to the assertion, that v(r;s) is an entire solution of (14)for all s(s1, s2) where we suppose s2> s1. This implies that (s1, s2) J0is empty. Indeed, if there is s0(s1, s2) J0, then, since f (r)ev(r;s0)→0 as r→ +∞, we can find (r, s) ∈ [r0, +∞) ×Iδ such that ϕ(r, s) =0. Consequently, as have done with (29), we may arrive at a contradiction. Now we conclude that (s1, s2) J1. Therefore, we have

v(r;s2)v(r;s1)=ϕ(r; ¯s)(s2s1)λ(s2s1)c >0, s¯∈(s1, s2),

for all r≥r0, where we apply the statement (c) and the formula (29)again. However, this contradicts the fact that both f (r)ev(r;s1)and f (r)ev(r;s2)approach to 1 as r→ +∞.

Claim 3. If sIδJ, then [s, s+δ) J.

The proof of Claim 3 is essentially the same as that of Claim 1. We omit it here. From Claims 1, 2 and 3, we conclude that Iδ+Jand IδJ0. If J1consists of more than one elements, then, in terms of Lemma 2.3, it is possible to select s1, s2J1with s1< s2such that either the interval (s1, s2) J0or (s1, s2) J. In either case, it contradicts the conclusion we have just mentioned. Therefore, the proof is complete. 2

3. Sharp lower bound of flux for N1≤2NK+1

In the preceding section we have proved that it is possible to find a radially symmetric nontopological solutions of the system (2)whose flux assumes an arbitrarily large value. We are going to make an effort to grasp the optimal lower bound for the value β associated with the (radial) nontopological solutions. Notice that, letting r→ +∞in the identity (21), we have β >4NK+4. Combining this with the constraint (19)thus gives

β >max{2N1+2,4NK+4}. (30)

To sharpen the lower bound, we take advantage of the radially symmetric solutions of the Liouville system in corre- spondence with (2), namely

Références

Documents relatifs

On the other hand, we also establish the sharp region of the flux for non-topological solutions and provide the classification of radial solutions of all types in the case of one

The set of simple solutions to the unperturbed problem is given by circles of latitude of radius (1 + k 2 0 ) −1/2 with respect to an arbitrarily chosen north pole. Together with

(1.6) Secondly, we will establish the minimization characterization for the first eigenvalue λ 1 (μ) of the fourth order measure differential equation (MDE) (1.6) with

- We prove the existence of a minimizer for a multidimen- sional variational problem related to the Mumford-Shah approach to computer vision.. Key words :

The existence of a positive lower bound of the compressi- bility is proved in the case of general quantum lattice systems at high temperature.. The compressibility

By considering an Abelian Chern-Simons model, we are led to study the existence of solutions of the Liouville equation with singularities on a flat torus.. A non-existence and

at infinity is required instead of strict convexity, but its conclusions are weaker too: the solutions are only asserted to be Lipschitz instead of smooth. We

In addition, it was proved in the same work that vortexless solution to Ginzburg-Landau equation continue to exists for h ex higher than the critical field (up to h ex ≤ Cε − α )