• Aucun résultat trouvé

Weakly nonlinear stochastic CGL equations

N/A
N/A
Protected

Academic year: 2022

Partager "Weakly nonlinear stochastic CGL equations"

Copied!
24
0
0

Texte intégral

(1)

www.imstat.org/aihp 2013, Vol. 49, No. 4, 1033–1056

DOI:10.1214/11-AIHP482

© Association des Publications de l’Institut Henri Poincaré, 2013

Weakly nonlinear stochastic CGL equations

Sergei B. Kuksin

1

SNRS and CMLS at Ecole Polytechnique, Palaiseau, France. E-mail:kuksin@math.polytechnique.fr Received 22 June 2011; revised 8 December 2011; accepted 2 February 2012

Dedicated to Claude Bardos on his 70th birthday

Abstract. We consider the linear Schrödinger equation under periodic boundary conditions, driven by a random force and damped by a quasilinear damping:

d dtu+i

+V (x) u=ν

uγR|u|2pu−iγI|u|2qu +√

νη(t, x). (∗)

The forceηis white in time and smooth inx; the potentialV (x)is typical. We are concerned with the limiting, asν→0, behaviour of solutions on long time-intervals 0≤tν1T, and with behaviour of these solutions under the double limitt→ ∞andν→0.

We show that these two limiting behaviours may be described in terms of solutions for thesystem of effective equations for(∗) which is a well posed semilinear stochastic heat equation with a non-local nonlinearity and a smooth additive noise, written in Fourier coefficients. The effective equations do not depend on the Hamiltonian part of the perturbation−iγI|u|2qu(but depend on the dissipative part−γR|u|2pu). Ifpis an integer, they may be written explicitly.

Résumé. Nous considérons l’équation de Schrödinger linéaire avec les conditions aux limites périodiques, perturbée par une force aléatoire et amortie par un terme quasi linéaire:

d dtu+i

+V (x) u=ν

uγR|u|2pu−iγI|u|2qu +√

νη(t, x). (∗)

La forceηest un processus aléatoire blanc en tempstet lisse enx; le potentielV (x)est typique. Nous étudions le comportement asymptotique des solutions sur de longs intervalles de temps 0≤tν1T, quandν→0, et le comportement des solutions quand t→ ∞etν→0. Nous démontrons qu’on peut décrire ces deux comportements asymptotiques en termes des solutions dusystème d’équations effectives pour(∗). Ce dernier est une équation de la chaleur avec un terme quasi linéaire non local et une force aléatoire lisse additive, qui est écrite dans l’espace de Fourier. Les équations ne dépendent pas de la partie hamiltonienne de la perturbation−iγI|u|2qu(mais elles dépendent de la partie dissipative−γR|u|2pu). Sipest un entier, on peut écrire ces équations explicitement.

MSC:35Q56; 60H15

Keywords:Complex Ginzburg–Landau equation; Small nonlinearity; Stationary measures; Averaging; Effective equations

0. Introduction

In [9,10] we considered the KdV equation on a circle, perturbed by a random force and a viscous damping. There we suggested auxiliaryeffective equationswhich are well posed and describe long-time behaviour of solutions for the perturbed KdV through a kind of averaging.

1Supported by l’Agence Nationale de la Recherche through the grant ANR-10-BLAN 0102.

(2)

In this work we apply the method of [9,10] to a weakly nonlinear situation when the unperturbed equation is not an integrable nonlinear PDE (e.g. KdV), but a linear Hamiltonian PDE with a generic spectrum. Since analytic properties of the latter are easier and better understood then those of the former, in the weakly nonlinear situation we understand better properties of the effective system and its relation with the original equation. Accordingly we can go further in analysis of long time behaviour of solutions.

More precisely, we are concerned withν-small dissipative stochastic perturbations of the space-periodic linear Schrödinger equation

d dtu+i

u+V (x)

u=0, x∈Td, (0.1)

i.e. with equations d

dtu+iAu=ν

uγRfp

|u|2

uIfq

|u|2 u

+√

νη(t, x), x∈Td, (0.2)

whereη(t, x)=dtd

j=1bjβj(t)ej(x). HereAu=AVu= −u+V (x)uand the potentialV (x)≥1 is sufficiently smooth; the real numbers p, q are nonnegative, the functions fp(r) and fq(r) are the monomials |r|p and |r|q, smoothed out near zero, and the constantsγR, γI satisfy

γR, γI≥0, γR+γI=1. (0.3)

IfγR =0, then due to the usual difficulty with the zero-mode of a solution u, the term uin the r.h.s. should be modified tou. The functions{ej(x), j ≥1}in the definition of the random force form the real trigonometric base ofL2(Td), the real numbersbjdecay sufficiently fast to zero whenj grows, and{βj(t), j≥1}, are the standard complex Wiener processes. So the noiseηis white in time and sufficiently smooth inx. It is convenient to pass to the slow timeτ=νtand write the equation as

˙

u+ν1iAu=uγRfp

|u|2

u−iγIfq

|u|2

u+η(τ, x), (0.4)

whereu˙=du/dτ. The equation is supplemented with the initial condition

u(0, x)=u0(x). (0.5)

It is known that under certain restrictions onp, qandd the problem (0.4), (0.5) has a unique solutionuν(τ, x),τ≥0, and Eq. (0.4) has a unique stationary measureμν. We review these results in Section1(there attention is given to the 1d case, while higher-dimensional equations are only briefly discussed).

Let{ϕk, k≥1}, and{λk, k≥1}, be the eigenfunctions and eigenvalues ofAV, 1≤λ1λ2≤ · · ·. We say that a potentialV isnonresonantif

j=1λjsj =0 for every finite nonzero integer vector(s1, s2, . . .). In Sections1.4,1.5 we show that nonresonant potentials are typical both in the sense of Baire and in the sense of measure. Assuming that V is nonresonant we are interested in two questions:

Q1. What is the limiting behaviour asν→0 of solutionsuν(τ, x)on long time-intervals 0≤τT? Q2. What is the limiting behaviour of the stationary measureμν asν→0?

For any complex function u(x), x∈Td, denote by Ψ (u)=v=(v1, v2, . . .)the complex vector of its Fourier coefficients with respect to the basis{ϕk}, i.e.u(x)=

vjϕj. Denote Ij=1

2|vj|2, ϕj=Argvj, j≥1. (0.6)

Then(I, ϕ)∈R+ ×Tare the action-angles for the linear Eq. (0.1). Thev- and(I, ϕ)-variables are convenient to study the two questions above. Writing (0.4) in the(I, ϕ)-variables we arrive at the following system:

d

Ij= · · ·, d

ϕj=ν1λj+ · · ·, (0.7)

(3)

where the dots stand for terms of order one (stochastic and deterministic). We have got slow/fast stochastic equations to which the principle of averaging is formally applicable (e.g., see [2,14] for the classical deterministic averaging and [4,8] for the stochastic averaging). DenotingIjν(τ )=Ij(uν(τ ))and averaging inϕtheI-equations in (0.7), using the rules of the stochastic calculus [4,8] and following the arguments in [10], we show in Section2that along sequences νj→0 we have the convergences

D Iνj(·)

D

I0(·)

, (0.8)

where the limiting processI0(τ ), 0τT, is a weak solution of the averagedI-equations. As in the KdV-case the averaged equations are singular and we do not know if their solution is unique. So we do not know if the convergence (0.8) holds asν→0. To continue the analysis we write Eq. (0.4) in thev-variables

˙

vk+iν1λkvk=Pk(v)+

j1

Bkjβ˙j(τ ), (0.9)

where the driftPk and the dispersionBkj are constructed in terms of the r.h.s. of Eq. (0.4) and the transformationΨ. It turns out that the Hamiltonian term−iγIfq(|u|2)ucontributes toP (v)a term which disappears in the averagedI- equations. We remove it fromP (v)and denote the restP (v). For any vector˜ θ=1, θ2, . . .)∈Tdenote byΦθthe linear transformation of the space of complex vectorsvwhich multiplies each componentvjby ej. Following [9] we average the vector fieldP˜by actions of the transformationsΦθ and get theeffective driftR(v)=

TΦθP (Φ˜ θv)dθ.

In Section3.1we show that

Rk(v)= −λkvk+Rk0(v), (0.10)

whereR0(v)is a smooth locally Lipschitz nonlinearity.

Since the noise in (0.9) is additive (i.e., the matrixB is v-independent), then the construction of the effective dispersion, given in [9] for a non-additive noise, simplifies significantly and defines the effective noise for Eq. (0.9) whosekth component equals(

lb2lΨkl2)1/2k(τ ).Accordingly the effective equations for (0.4) become

˙

vk=Rk(v)dτ +

l

b2lΨkl2 1/2

k(τ ), k≥1. (0.11)

By construction this system is invariant under rotations: if v(τ ) is its weak solution, then Φθv(τ ) also is a weak solution. Due to (0.10) this is the heat equationu˙= −Aufor a complex functionu(τ, x), perturbed by a non-local smooth nonlinearity and a nondegenerate smooth noise, written in terms of the complex Fourier coefficients vj. It turns out to be a monotone equation, so its solution is unique (see Section3.2).

In particular, if in (0.4)p=1, then the system of effective equations takes the form

˙

vk= −vk kMk)+γR

|vl|2Lkl

dτ+

l

b2lΨkl2 1/2

k(τ ), k≥1, (0.12)

whereMk=

V (x)ϕk2(x)dxandLkl=(2δkl)

ϕk2(x)ϕl2(x)dx. See Example3.1(the calculations, made there for d=1, remain the same ford≥2).

It follows directly from the construction of effective equations that actions{I (vk(τ ))=12|vk(τ )|2, k≥1}of any solutionv(τ )of (0.11) is a solution of the system of averagedI-equations. On the contrary, every solutionI0(τ )of the averagedI-equations, obtained as a limit (0.8), can be lifted to a weak solution of (0.11). Using the uniqueness we get

Theorem 0.1. LetIν(τ )=I (uν(τ )),whereuν(τ ), 0τT,is a solution of(0.4), (0.5).Thenlimν0D(Iν(·))= D(I0(·)),whereI0(τ ), 0τT,is a weak solution of the averagedI-equations.Moreover,there exists a unique solutionv(τ )of(0.11)such thatv(0)=v0=Ψ (u0)andD(I (v(·))=D(I0(·)),whereI (v(τ ))j=12|vj(τ )|2.

(4)

The solutionsI0(τ )andv(τ )satisfy some apriori estimates, see Theorem3.5. Concerning distribution of the angles ϕ(uν(τ ))and their joint distribution with the actions see Section2.4.

Now let μν be the unique stationary measure for Eq. (0.4) and uν be a corresponding stationary solution, D(uν(τ ))μν. As above, along sequencesνj →0 the actionsIνj(τ )=I (uνj(τ ))converge in distribution to a stationary solutionI(τ )of the averagedI-equations. This solution can be lifted to a stationary weak solutionsv(τ ) of effective Eqs (0.11). Since that system is monotone, then its stationary measuremis unique. So the limit above holds asν→0. As the effective system is rotation invariant, then in the(I, ϕ)-variables its unique stationary measure has the form dm=mI(dI )×dϕ, where dϕ=dϕ12· · ·is the Haar measure onT. It turns out that the measure limν0μνalso has the rotation-invariant form and we arrive at the following result (see Theorem4.3,4.4for a precise statement):

Theorem 0.2. When ν→0 we have the convergences D(I (uν(·)) DI (v(·)) and Ψμν m,where dm= mI(dI )×dϕ.

Accordingly every solutionuν(τ )of (0.2) obeys the following double limit

νlim0 lim

t→∞D uν(t)

=Ψ1m. (0.13)

By Theorems0.1and0.2, the actionsI (uν(τ ))of a solutionuν of (0.4), (0.5) converge in distribution to those of a solutionv(τ )of the effective system (0.11) withv(0)=Ψ (u0), both for 0τT and whenτ→ ∞. We conjecture that this convergence hold for eachτ ≥0, uniformly in τ (the space of measures is equipped with the Wasserstein distance).

In Example4.6we discuss Theorem0.2for equations withp=1, when the effective equations become (0.12). In particular, we show that Theorem0.2implies that in Eqs (0.2) with smallν there is no direct or inverse cascade of energy.

In Example4.5we discuss Theorem0.2for the caseγR=0 (when the nonlinear part of the perturbation is Hamil- tonian) and its relation to the theory of weak turbulence.

We note that the effective Eqs (0.11) depend on the potentialV (x)in a regular way and are well defined without assuming thatV (x)is nonresonant (cf. Eqs (0.12)). In particular, ifVM(x)→1 asM→ ∞, where eachVM(x)≥1 is a nonresonant potential, then in (0.13)mM m(1), wherem(1)is a unique stationary measure for Eq. (0.12) with V (x)≡1. In this equationΨkl=δk,l,Mk≡1 and the constantsLklcan be written down explicitly.

In Section5we show that Theorems0.1,0.2remain true for 1d equations with non-viscous damping (whenuin the l.h.s. of (0.2) is removed, butγR>0).

Inviscid limit

A stationary measureμν for Eq. (0.4) also is stationary for the fast-time Eq. (0.2). LetUν(t)be a corresponding stationary solution,DUν(t)μν. It is not hard to see that the system of solutionsUν(t)is tight on any finite time- interval[0,T˜]. Let{Uνj, νj→0}, be a converging subsequence, i.e.

D(Uνj) Q, μνj μ.

Thenμis an invariant measure for the linear Eq. (0.1) andQ=D(U(·)), whereU(t),0t≤ ˜T, is a stationary process such thatD(U(t))μand every trajectory ofUis a solution of (0.1). The limitD(Uνj) D(U)is the inviscid limit for Eq.(0.2). Equation (0.1) has plenty of invariant measures: if we write it in the action-angle variables (0.6), then every measure of the formm(dI )×dϕis invariant (see [12] for the more complicated inviscid limit for nonlinear Schrödinger equation). Theorem0.2explains which one is chosen by Eq. (0.2) for the limit limν0μν.

The inviscid limit for the damped/driven KdV equation, studied in [9,10] is similar: the limit of the stationary measures for the perturbed equations is a stationary measure of the corresponding effective equations. Due to a com- plicated structure of the nonlinear Fourier transform which integrates KdV, uniqueness of their invariant measure is not proved yet. So the final results concerning the damped/driven KdV are less complete than those for the weakly perturbed CGL equation in Theorem0.2.

(5)

Finally consider the damped/driven 2d Navier–Stokes equations with a small viscosity ν and a random force, similar to the forces above and proportional to√

ν:

vtνv+(v· ∇)v+ ∇p=√

νη(t, x); divv=0, v∈R2, x∈T2. (0.14) It is known that (0.14) has a unique stationary measureμν, the family of measures{μν,0< ν≤1}is tight, and every limiting measure limνj0μνjis a non-trivial invariant measure for the 2d Euler Eq. (0.14)ν=0, see Section 5.2 of [13].

Hovewer it is nonclear if the limiting measure is unique and how to single it out among all invariant measures of the Euler equation. The research [9,10] was motivated by the belief that the damped/driven KdV is a model for (0.14).

Unfortunately, we still do not know up to what extend the description of the inviscid limit for the damped/driven KdV and for weakly nonlinear CGL in terms of the effective equations is relevant for the inviscid limit of the 2d hydrodynamics.

Agreements

Analyticity of mapsB1B2 between Banach spacesB1 andB2, which are the real parts of complex spaces B1c andB2c, is understood in the sense of Fréchet. All analytic maps which we consider possess the following additional property: for anyRa map analytically extends to a complexR>0)-neighbourhood of the ball{|u|B1< R}inB1c. Notations

χA stands for the indicator function of a setA (equal 1 inAand equal 0 outside A). By κ(t)we denote various functions oft such thatκ(t)→0 whent→ ∞, and byκ(t)denote functionsκ(t)such thatκ(t)=o(tN)for eachN. We writeκ(t)(t;R)to indicate thatκ(t)depends on a parameterR.

1. Preliminaries 1.1. Apriori estimates

We consider the 1d CGL equation on a segment [0,π] with a conservative linear part of order one and a small nonlinearity. The equation is supplemented with Dirichlet boundary conditions which we interpret as odd 2π-periodic boundary conditions. Introducing the slow timeτ=νt(cf. theIntroduction) we write the equation as follows:

˙

u+iν1

uxx+V (x)u

uxxγR|u|2pu−iγI|u|2qu+ d dτ

i=1

bjβj(τ )ej(x),

(1.1) u(x)u(x+2π)≡ −u(x).

Hereu˙=du,p, q∈Z+:=N∪ {0}(only for simplicity, see next section),κ>0, constantsγRandγI satisfy (0.3) andRV (x)≥0 is a sufficiently smooth even 2π-periodic function,{ej, j≥1}is the sine-basis,

ej(x)= 1

√πsinj x,

andβj, j≥1, are standard independent complex Wiener processes. That is,βj(τ )=βj(τ )+iβj(τ ), whereβ±j(τ ) are standard independent real Wiener processes. Finally, the real numbersbjall are nonzero and decay whenj grows in such a way thatB1<∞, where

Br :=2 j=1

j2rb2j≤ ∞ forr≥0.

By Hr,r∈Rwe denote the Sobolev space of orderr of complex odd periodic functions and provide it with the homogeneous norm · r,

u2r = l=1

|ul|2l2r foru(x)= l=1

ulel(x), u0= u

(6)

(ifr∈N, thenur= |∂xrur|L2).

Letu(t, x)be a solution of (1.1) such thatu(0, x)=u0.Applying Ito’s formula to12u2we get that d

1 2u2

=

γR|u|2p2p++22−κu21+1 2B0

dτ+dM(τ ), (1.2)

whereM(τ )is the martingaleτ

0

bjuj·dβj(τ ). Here |u|r stands for theLr-norm, 1≤r≤ ∞, and for complex numbersz1,z2we denote byz1·z2their real scalar product,

z1·z2=Rez1z2.

So(uj+iuj)·(dβj+i dβj)=ujj+ujj. From (1.2) we get in the usual way (e.g., see Section 2.2.3 in [13]) that

Eeρκu(τ )2C

κ, B0,u0

τ≥0 (1.3)

for a suitableρκ>0, uniformly inν >0.

Denoting E(τ )=1

2u(τ )2+γR τ

0

|u|2p2p++22ds+κ 2

τ

0

u21ds

and noting that the characteristic of the martingaleMisM(τ )=

bj2|uj|2τb2Mu2τ, wherebM=max|bj|, we get from (1.2) that

E(τ )≤ 1

2u02+1

2B0τ+M(τ )−κ 2

τ

0 u21ds

≤ 1

2u02+1

2B0τ1b2M

κbM2M(τ )

−1 2

κbM2M (τ )

.

Applying in a standard way the exponential supermartingale estimate to the term in the square bracket in the r.h.s.

(e.g., see [13], Section 2.2.3 ), we get that P

sup

τ0

E(τ )−1 2B0τ

≥1

2u02+ρ

≤e2κρbM2 (1.4)

for anyρ >0.

Now let us re-write Eq. (1.1) as follows:

˙

u+iν1

uxx+V (x)u+νγI|u|2qu

uxxγR|u|2pu+ d dτ

bjβj(τ )ej. (1.5)

The l.h.s. is a Hamiltonian system with the hamiltonian−ν1H (u), H (u)=1

2Au, u +γI

ν 2q+2

|u|2q+2dx, A= − 2

∂x2+V (x).

For anyj ∈Nwe denote ur

2= Aru, u

.

Then dH (u)(v)= Au, v +γIν|u|2qu, vand 1

2 ·2 j=1

b2jd2H (u)(ej, ej)=1

2B1+γIνX(τ ),

(7)

where

Br =2

b2jej2r=2

bj2λrjr, and

X(τ )=2qRe |u|2q2u2

j

b2jej(x)2

dx+

|u|2q

j

b2jej(x)2dx

CB0u(τ )2q

2q.

Therefore applying Ito’s formula we get that dH

u(τ )

=

γR

Au,|u|2pu

Au, uxxγIνγR

|u|2p+2q+2dx +κγIν

|u|2qu, uxx

+B1 +γIνX(τ )

dτ+dM(τ ), (1.6)

where dM(τ )=

bjAu+γIν|u|2qu, ej ·dβj(τ ).

DenotingUq(x)=q+11uq+1andUp(x)=p+11up+1,we have |u|2qu, uxx

≤ −

|ux|2|u|2qdx= −

∂xUq 2, and a similar relation holds forq replaced byp. Accordingly,

dH u(τ )

≤ −1 2

κu22+γR

∂xUp

2γIν

∂xUq

2 +νγIγR

|u|2p+2q+2dx−Cκu2−2B1

dτ+dM(τ ), (1.7)

whereCκmay be chosen independent fromκifγR>0. Considering relations onH (u)m,m≥1, which follow from (1.7) and (1.6), using (1.4) and arguing by induction we get that

E

sup

0tT

H u(t )m

+κ 2

T 0

Hm1(u)u22ds

H (u0)m+Cm(κ, T , B1)

1+ u0cm

, (1.8)

EH u(t )m

Cm(κ, B1)

1+H (u0)m+ u0cm

t >0, (1.9)

for anym. Estimates (1.8) in a traditional way (cf. [5,12,16,18]) imply that Eq. (1.1) isregular in spaceH1in the sense that for anyu0H1it has a unique strong solution, satisfying (1.4), (1.8).

1.2. Stationary measures

The a-priori estimates on solutions of (1.1) and the Bogolyubov–Krylov argument (e.g., see in [13]) imply that Eq. (1.1) has a stationary measureμν, supported by spaceH2. Now assume that

bj =0 ∀j. (1.10)

Then the approaches, developed in the last decade to study the 2d stochastic Navier–Stokes equations, apply to (1.1) and allow to prove that under certain restrictions on the equation the stationary measureμν is unique. In particular

(8)

this is true ifγI =0 (the easiest case), or ifpq andγR =0 (see [16]), or ifγR=0 andp=1 (see [18]). In this case any solutionu(t )of (1.1) withu(0)=u0H1satisfies

Du(t ) μν ast→ ∞. (1.11)

This convergence and (1.3), (1.9) imply that

eρκu2μν(du)C(κ, B), (1.12)

u2m1 μν(du)Cm(κ, B1)m. (1.13)

1.3. Multidimensional case

In this section we briefly discuss a multidimensional analogy of Eq. (1.1):

˙

u+iν1Au=uγRfp

|u|2

u−iγIfq

|u|2 u + d

j=1

bjβj(τ )ej(x), u=u(τ, x), x∈Td. (1.14)

HereAu= −u+V (x)u,VCN(Td,R)andV (x)≥1. The numbersγI, γR satisfy (0.3). Functionsfp≥0 and fq≥0 are real-valued smooth and

fp(t)=tp fort≥1, fq(t)=tq fort≥1,

wherep, q≥0. IfγR=0, then the termuin the r.h.s. should be modified tou. By{ek, k≥1}, we denote the usual trigonometric basis of the spaceL2(Td)(formed by all functionsπd/2fs1(x1)· · ·fsd(xd), where eachfs(x)is sinsx or cossx), parameterised by natural numbers. These are eigen-functions of the Laplacian,−er =λrer. We assume that

BN

1=2

k

λNk1b2k<, (1.15)

whereN1=N1(d)is sufficiently large. In this section we denote by(Hr, · r)the Sobolev spaceHr =Hr(Td,C), regarded as a real Hilbert space, and·,·stands for the realL2-scalar product.

Noting that(fp(|u|2)u− |u|2pu)and(fq(|u|2)u− |u|2qu)are bounded Lipschitz functions with compact support we immediately see that the a-priori estimates from Section1.1remain true for solutions of (1.14). Accordingly, for anyu0H1L2q+2Eq. (1.1) has a solutionu(t, x)such thatu(0, x)=u0, satisfying (1.3), (1.8), (1.9).

Now assume that

p, q <∞ ifd=1,2, p, q < 2

d−2 ifd≥3. (1.16)

Applying Ito’s formula to the processesAmu(τ ), u(τ )n,m, n≥1, using (1.3), (1.8), (1.9) and arguing by induction (first innand next inm) we get that

E

sup

0τT

u(τ )2n

2m+ T

0

u(s)2

2m+1u(s)2n2

2m ds

u02n2m+C(m, n, T )

1+ u0cm,n

, (1.17)

Eu(τ )2n

2mC(m, n)τ ≥0, (1.18)

(9)

for each m andn, whereC(m, n, T ) andC(m, n) also depends on |V|CN andBN1 (see (1.15)), and N =N (m), N1=N1(m).

Relations (1.17) withm=m0≥1 in the usual way (cf. [5,12,16,18]) imply that Eq. (1.14) isregular in the space Hm0L2q+2in the sense that for anyu0Hm0L2q+2 it has a unique strong solutionu(t, x), equalu0att=0, and satisfying estimates (1.3), (1.17) withm=m0 for anyn. By the Bogolyubov–Krylov argument this equation has a stationary measureμν, supported by the spaceHm0L2q+2, and a corresponding stationary solutionuν(τ ), Duν(τ )μν, also satisfies (1.3) and (1.18) withm=m0.

If (1.10) holds and (1.16) is replaced by a stronger assumption, then a stationary measure is unique. IfγI=0, the uniqueness readily follows, for example, from the abstract theorem in [13]. In [18] this assertion is proved if

γR=0 and q≤1 ifd=1, q <1 ifd=2, q≤2/d ifd=3. (1.19) In [16] it is established if

p=q, γR, γI>0 ifd=1,2, and p=q < 2

d−2, γR, γI>0 ifd≥3; (1.20) the argument of that work also applies ifp > q.

Note that whenγR=0 or whenp < q (i.e., when the nonlinear damping is weaker than the conservative term), the assumptions (1.19), (1.20), needed for the uniqueness of the stationary measure, are much stronger than the assumptions (1.16), needed for the regularity. This gap does not exist (at least it shrinks a lot) if the random force in Eq. (1.14) is not white in time, but is a kick-force. See in [11] the abstract theorem and its application to the CGL equations.

1.4. Spectral properties ofAV:One-dimensional case

As in Section1.1we denoteAV =A= −2/∂x2+V (x), where the potentialV (x)≥0 belongs to the spaceCeN ofCN-smooth even and 2π-periodic functions,N≥1. Letφ1, φ2, . . .be theL2-normalised complete system of real eigenfunctions ofAV with the eigenvalues 1≤λ1< λ2<· · ·. Consider the linear mapping

Ψ:Hu(x)v=(v1, v2, . . .)∈C, defined by the relationu(x)=

vkφk(x). In the space of complex sequencesvwe introduce the norms

|v|2hm=

k1

|vk|2λmk, m∈R,

and denotehm= {v| |v|hm<∞}. Due to the Parseval identity,Ψ:Hh0is a unitary isomorphism. By{Ψkm, k, m≥1}we denote the matrix ofΨ with respect to the basis{ej}inHand the standard basis inh0. SinceΨ maps real vectors to real, its matrix has real entries.

For anym∈Nwe have|v|2hm= Amu(x), u(x). So the norms|v|hm andumare equivalent form=0, . . . , N.

SinceΨ=Ψ1, then the norms are equivalent for integer|m| ≤N. By interpolation they are equivalent for all real

|m| ≤N. So

the maps Ψ:Hmhm, |m| ≤N, are isomorphisms. (1.21)

DenoteG=Ψ1:hmHm. Then ΨAG=diag{λk, k≥1} =:A.

Consider the operator

L:=Ψ()G=Ψ(AV )G=AΨVG=:AL0. (1.22)

(10)

By (1.21)L0=ΨVGdefines bounded linear maps

L0:hmhm ∀|m| ≤N, (1.23)

and in the spaceh0it is selfadjoint.

For any finiteMconsider the mapping

ΛM:CeN→RM, V (x)1, . . . , λM).

Since the eigenvaluesλj are different, this mapping is analytic. As∇λj(V )=φj(x)2and the functionsφ21, φ22, . . . are linearly independent by the classical result of G. Borg (1946), then for anyVCeNthe linear mapping

M(V ):CNe →RM is surjective (1.24)

(all this result may be found in [17]; e.g. see there p. 46 for Borg’s theorem). In the spaceCeN consider a Gaussian measureμKwith a nondegenerate correlation operatorK(so for the quadratic functionf (V )= V , ξL2V , ηL2 we have

f (V )μK(dV )= Kξ, η). Relation (1.24) easily implies

Lemma 1.1. For anyM≥1the measureΛMμKis absolutely continuous with respect to the Lebesgue measure on RM.

We will call a vectorΛ∈Rnonresonantif for any nonzero integer vectorsof finite length we have

Λ·s =0. (1.25)

A potentialV (x)is called nonresonant if its spectrumΛ(V )=1, λ2, . . .)is nonresonant. The nonresonant potentials are defined inCeN by a countable family of open dense relations (1.25). So

the nonresonant potentials form a subset ofCeNof the second Baire category. (1.26) Applying Lemma1.1we also get

the nonresonant potentials form a subset ofCeNof fullμKmeasure (1.27) for any Gaussian measureμKas above.

The nonresonant vectorsΛare important because of the following version of the Kronecker–Weyl theorem:

Lemma 1.2. LetfCn+1(Tn)for somen∈N.Then for any nonresonant vectorΛwe have

Tlim→∞

1 T

T 0

f

q0+t Λn

dt=(2π)n

fdx, Λn=1, . . . , Λn), uniformly inq0∈Tn.The rate of convergence depends onn,Λand|f|Cn+1. Proof. Let us write f (q) as the Fourier series f (q) =

fseis·q. Then for each nonzero s we have |fs| ≤ Cn+1|f|Cn+1|s|n1.So for anyε >0 we may findR=Rεsuch that|

|s|>Rfseis·q| ≤ε2for eachq. Now it suffices to show that

1 T

T

0

fR

q0+t Λn

dt−f0ε

2 ∀TTε (1.28)

for a suitableTε, wherefR(q)=

|s|≤Rfseis·q. But 1

T T

0

eis·(q0+t Λn)dt ≤ 2

T|s·Λn|

(11)

for each nonzeros. Therefor the l.h.s. of (1.28) is

≤ 2 T inf

|s|≤R

Λn1

|fs| ≤T1|f|C0C(R, Λ).

Now the assertion follows.

1.5. Spectral properties ofAV:Multi-dimensional case

Now let, as in Section 1.3,A=AV be the operator A= −+V (x), x∈Td, where 1≤V (x)CN(Td). Let {φk(x), k≥1} be its L2-normalised eigenfunctions and {λk, k≥1}, be the corresponding eigenvalues, 1≤λ1λ2≤ · · ·. For anyM≥1 denote byFMCN(Td)the open domain

FM= {V |λ1< λ2<· · ·< λM}.

Its complementFMc is a real analytic variety inCN(Td)of codimension≥2, soFM is connected (see [6] and refer- ences therein). The functionsλ1, . . . , λM are analytic inFM. Let us fix any nonzero vectors∈Zsuch thatsl=0 forl > M. The set

Qs=

VFM|Λ(V )·s=0

clearly is closed inFM. Since the functionΛ(V )·sis analytic inFM, then eitherQs=FM, orQs is nowhere dense inFM. Theorem 1 from [6] immediately implies thatQs =FM, so (1.26) also holds true in the case we consider now.

LetμK be a Gaussian measure with a nondegenerate correlation operator, supported by the spaceCN(Td). As Λ(V )·s is a non-trivial analytic function on FM and FMc is an analytic variety of positive codimension, then μK(Qs)=0 (e.g., see Theorem 1.6 in [1]). Since this is true for anyM and any s as above, then the assertion (1.27) also is true.

2. Averaging theorem

The approach and the results of this section apply both to Eqs (1.1) and (1.14). We present it for Eq. (1.1) and at Section2.5discuss small changes, needed to treat (1.14). Everywhere belowT is an arbitrary fixed positive number.

2.1. Preliminaries

In Eq. (1.1) withuH1we pass to thev-variables,v=Ψ (u)h1:

˙

vk+iν1λkvk=Pk(v)dτ+

j1

Bkjj(τ ), k≥1. (2.1)

HereBkj =Ψkjbj (a matrix with real entries, operating on complex vectors), and

Pk=Pk1+Pk2+Pk3, (2.2)

whereP1, P2andP3are, correspondingly, the linear, dissipative and Hamiltonian parts of the perturbation:

P1(v)Ψ2

∂x2u, P2(v)= −γRΨ

|u|2pu

, P3(v)= −iγIΨ

|u|2qu , whereu=G(v). We will refer to Eqs (2.1) as to thev-equations.

Fork≥1 let us denoteIk=I (vk)=12|vk|2andϕk =ϕ(vk)=ArgvkS1, whereϕ(0)=0∈S1. Consider the mappings

ΠI:hrvI=(I1, I2, . . .)hrI+, Πϕ:hrvϕ=1, ϕ2, . . .)∈T.

Références

Documents relatifs

It is thus not evident to prove the uniform in ε wellposedness of the BOB equation in low regularity spaces and , as a consequence, to study its inviscid limit behavior.. However,

Proposition 5.1 of Section 5 (which is the main ingredient for proving Theorem 2.1), com- bined with a suitable martingale approximation, can also be used to derive upper bounds for

Les biais de perception ont donc une grande influence sur la nature et la qualité des échanges entre conjoints, ils déterminent avec d’autres facteurs cognitifs

Miller, Semiclassical nonlinear Schr¨ odinger equations with potential and focusing initial data, Osaka J.. Cazenave, Semilinear Schr¨ odinger equations, Courant Lecture Notes

In order to conclude this study, we shall present some particular example of self-stabilizing diffusion which presents the following property: any invariant symmetric measure

Abstract: We study a damped stochastic non-linear Schr¨odinger (NLS) equation driven by an additive noise.. It is white in time and smooth

More precisely, it is possible to perform formal asymptotic expansions for weakly nonlinear surface waves, and to derive a well-posed amplitude equation even though the

The fixed point technique was used in [13] to estimate the perturbation to the standing soli- tary wave of the Nonlinear Schr¨odinger (NLS) equation when either the external