• Aucun résultat trouvé

On the Korteweg-de Vries long-wave approximation of the Gross-Pitaevskii equation II

N/A
N/A
Protected

Academic year: 2021

Partager "On the Korteweg-de Vries long-wave approximation of the Gross-Pitaevskii equation II"

Copied!
45
0
0

Texte intégral

(1)

HAL Id: hal-00371344

https://hal.archives-ouvertes.fr/hal-00371344v2

Submitted on 12 Dec 2009

HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

On the Korteweg-de Vries long-wave approximation of the Gross-Pitaevskii equation II

Fabrice Bethuel, Philippe Gravejat, Jean-Claude Saut, Didier Smets

To cite this version:

Fabrice Bethuel, Philippe Gravejat, Jean-Claude Saut, Didier Smets. On the Korteweg-de Vries long- wave approximation of the Gross-Pitaevskii equation II. Communications in Partial Differential Equa- tions, Taylor & Francis, 2010, 35 (1), pp.113-164. �10.1080/03605300903222542�. �hal-00371344v2�

(2)

On the Korteweg-de Vries long-wave approximation of the Gross-Pitaevskii equation II

Fabrice B´ethuel 1, Philippe Gravejat 2, Jean-Claude Saut3, Didier Smets 4 December 12, 2009

Abstract

In this paper, we proceed along our analysis of the Korteweg-de Vries approximation of the Gross-Pitaevskii equation initiated in [6]. At the long-wave limit, we establish that solutions of small amplitude to the one-dimensional Gross-Pitaevskii equation split into two waves with opposite constant speeds ±

2, each of which are solutions to a Korteweg-de Vries equation. We also compute an estimate of the error term which is somewhat optimal as long as travelling waves are considered. At the cost of higher regularity of the initial data, this improves our previous estimate in [6].

1 Introduction

1.1 Statement of the results

In this paper, we proceed along our study initiated in [6] of the one-dimensional Gross-Pitaevskii equation

i∂tΨ +∂x2Ψ = Ψ(|Ψ|2−1) on R×R, (GP) supplemented with the boundary condition at infinity

|Ψ(x, t)| →1, as|x| →+∞.

This boundary condition is suggested by the formal conservation of the Ginzburg-Landau energy E(Ψ) = 1

2 Z

R|∂xΨ|2+ 1 4

Z

R

(1− |Ψ|2)2. In this paper, we will only consider finite energy solutions to (GP).

The Gross-Pitaevskii equation is integrable by means of the inverse scattering method, and it has been formally analyzed within this framework in [16], and rigorously in [14]. Concerning the Cauchy problem, it can be shown (see [18, 13, 6]) that (GP) is globally well-posed in the spaces

Xk(R) =

u∈L1loc(R,C), s.t.1− |u|2 ∈L2(R) and∂xu∈Hk−1(R) ,

1UPMC, Universit´e Paris 06, UMR 7598, Laboratoire Jacques-Louis Lions, F-75005, Paris, France. E-mail:

bethuel@ann.jussieu.fr

2Centre de Recherche en Math´ematiques de la D´ecision, Universit´e Paris Dauphine, Place du Mar´echal De Lattre De Tassigny, 75775 Paris Cedex 16, France, and ´Ecole Normale Sup´erieure, DMA, UMR 8553, F-75005, Paris, France. E-mail: gravejat@ceremade.dauphine.fr

3Laboratoire de Math´ematiques, Universit´e Paris Sud and CNRS UMR 8628, Bˆatiment 425, 91405 Orsay Cedex, France. E-mail: Jean-Claude.Saut@math.u-psud.fr

4UPMC, Universit´e Paris 06, UMR 7598, Laboratoire Jacques-Louis Lions, F-75005, Paris, France, and ´Ecole Normale Sup´erieure, DMA, UMR 8553, F-75005, Paris, France. E-mail: smets@ann.jussieu.fr

(3)

for any k≥1. More precisely, we have

Proposition 1 ([6]). Let k∈N and Ψ0 ∈Xk(R). Then, there exists a unique solution Ψ(·, t) in C0(R, Xk(R)) 1 to (GP) with initial data Ψ0. Furthermore, the energy E is conserved along the flow.

Ifu belongs toX1(R) and satisfies

E(u) < 2√ 2 3 ,

then it does not vanish, and we may write u = |u|expiθ, where θ is continuous (see e.g. [4]).

Here, we will focus on solutions with small energy, so that in view of the conservation of the energy, we may write

Ψ(·, t) =̺(·, t) expiϕ(·, t).

More precisely, we will consider initial data which are small long-wave perturbations of the constant one, namely

̺(x,0) =

1−ε62Nε0(εx)12 , ϕ(x,0) = ε

6

2Θ0ε(εx),

where 0< ε <1 is a small parameter, and Nε0 and Wε0 =∂xΘ0ε are uniformly bounded in some Sobolev spaces Hk(R) for sufficiently large k. We will add two additional assumptions on Θ0ε and Nε0. We will assume that

kNε0kM(R)+k∂xΘ0εkM(R)<+∞, (1) with a uniform bound in ε. Here,k · kM(R) denotes the norm defined onL1loc(R) by

kfkM(R)= sup

(a,b)∈R2

Z b a

f(x)dx

, (2)

so that (1) implies in particular that Θ0ε is uniformly bounded in L(R). In the appendix, we will introduce a notion of mass for Ψ closely related to the M-norm of 1− |Ψ|2, and prove its conservation by the Gross-Pitaevskii flow.

We next introduce the slow coordinates x=ε(x +√

2t), x+ =ε(x−√

2t), and τ = ε3 2√

2t.

The definition of the new coordinates x and x+ corresponds to reference frames travelling to the left and to the right respectively with speed √

2 in the original coordinates (x, t). We define accordingly the rescaled functions Nε± and Θ±ε as follows

Nε±(x±, τ) = 6

ε2η(x, t) = 6 ε2ηx±

ε ± 4τ ε3,2√

2τ ε3

, Θ±ε(x±, τ) = 6√

2

ε ϕ(x, t) = 6√ 2 ε ϕx±

ε ±4τ ε3,2√

2τ ε3

,

(3)

1Here, the spaceXk(R) is endowed with the distance

dXk,A(u, v) =kuvkL([−A,A])+k∂xuxvkHk−1(R)+k|u| − |v|kL2(R), for some givenA >0 (see e.g. [13, 5] for more details).

(4)

where η= 1−̺2. Setting

Uε(x, τ) = 1 2

Nε(x, τ) +∂xΘε(x, τ) , Uε+(x+, τ) = 1

2

Nε+(x+, τ)−∂x+Θ+ε(x+, τ) ,

(4)

our main result is

Theorem 1. Let k≥0 andε >0 be given. Assume that the initial dataΨ0 belongs toXk+6(R) and satisfies the assumption

kNε0kM(R)+k∂xΘ0εkM(R)+kNε0kHk+5(R)+εk∂xk+6Nε0kL2(R)+k∂xΘ0εkHk+5(R)≤K0. (5) Let U and U+ denote the solutions to the Korteweg-de Vries equations

τU+∂x3U+UxU= 0, (KdV) and

τU+−∂x3+U+− U+x+U+ = 0, (6) with the same initial value2 asUε, respectivelyUε+. Then, there exist positive constants ε1 and K1, depending only on k and K0, such that

kUε(·, τ)− U(·, τ)kHk(R)+kUε+(·, τ)− U+(·, τ)kHk(R)≤K1ε2expK1|τ|, (7) for any τ ∈Rprovided ε≤ε1.

Remark 1. In the original time variable, the Korteweg-de Vries approximation is valid on a time interval t∈[0, Tε] with

Tε=o

|log(ε)| ε3

.

Moreover, the approximation error remains of order O(ε2) on a time interval t ∈ [0, Tε] with Tε =O(ε3).

In order to explain the statements of Theorem 1, it is presumably useful to recast them in the context of known results about the long-wave limit of the Gross-Pitaevskii equation. First, we rewrite (GP) in the slow coordinates (y, s) = (εx, εt) and set

̺(x, t) =

1−ε62nε(εx, εt)12 ,

xϕ(x, t) = ε2

6

2wε(εx, εt).

In this setting, (GP) translates into the system fornε and wε,

snε−√

2∂ywε=−3ε22y(nεwε),

swε−√

2∂ynε=−3√

2y 2 ynε

6ε2nε +w362ε +ε22(6(∂yεn2εn)ε2)2

. (8)

It has been shown in [3] that for suitably small data and times, this system is well-approximated by the linear wave equation. More precisely, assume thats≥2 and

K(s)ε2k(Nε0, Wε0)kHs+1(R)×Hs(R)≤1,

2Since their initial data depend on ε, U and U+ do as well. We voluntarily hide this dependence in the notations in order to stress the fact that the equations they satisfy are independent ofε.

(5)

whereK(s) refers to some positive constant depending only ons. Let (n,w) denote the solution of the free wave equation

sn−√

2∂yw= 0,

sw−√

2∂yn= 0, (9)

with initial data (Nε0, Wε0). Then, for any 0≤t≤Tε, we have

k(nε, wε)(·, εt)−(n,w)(·, εt)kHs−2(R)×Hs−2(R)

≤K(s)ε3t

k(Nε0,Wε0)kHs+1(R)×Hs(R)+k(Nε0, Wε0)k2Hs+1(R)×Hs(R)

, whereTε= K(s)ε3k(Nε0, Wε0)kHs+1(R)×Hs(R)

−1

. In particular, whenk(Nε0, Wε0)kHs+1(R)×Hs(R)

remains uniformly bounded, then Tε=O ε3

, and the wave equation is a good approximation for times of order o ε3

. The general solution to (9) may be written as (n,w) = (n+,w+) + (n,w),

where the functions (n±,w±) are solutions to (9) given by the d’Alembert formulae, n+(y, s),w+(y, s)

= N+(y−√

2s), W+(y−√ 2s)

, n(y, s),w(y, s)

= N(y+√

2s), W(y+√ 2s)

,

where the profilesN± and W±are real-valued functions on R. Solutions may therefore be split into right and left going waves of speed √

2. Since the functions (n±,w±) are solutions to (9), it follows that

y N++W+

= 0, and∂y N−W

= 0, so that, if the functions decay to zero at infinity, then

N±=∓W±= Nε0∓Wε0

2 .

Theorem 1 extends our earlier results in [6] (see also [9] for an alternative approach and an extension to the higher dimensional case). It shows that the Korteweg-de Vries equation provides the appropriate approximation for time scales of order O ε3

. The definition of the new coordinate x+, respectively x, corresponds to a reference frame travelling to the left, respectively to the right, with speed √

2 in the original coordinates (x, t). In the frame corresponding to x, the wave (n,w), originally travelling to the left, is now stationary, whereas the right going wave (n+,w+) now has a speed equal to 8ε−2. The coordinate x is therefore particularly appropriate for the study of waves travelling to the left, whereas the coordinate x+ is appropriate for the study of waves travelling to the right. In [6], we imposed some additional assumptions which implied in particular the smallness ofUε+, so that it was only the study of waves going to the left which was addressed. This approach simplifies somehow the analysis.

In this paper, we remove this smallness assumption, at the cost however of new assumption (1), and we analyze both waves at the same time. Finally, we would like to emphasize also that comparing Theorem 1 with Theorem 1.4 in [6], the error term now involves ε2 instead ofε. As explained in [6], the ε2 is somewhat optimal, as the specific examples provided by travelling waves show. This improvement is related to the fact that we use higher order derivatives. As a matter of fact, the same improvement holds in the setting of Theorem 1.4 of [6]. More precisely, we have

(6)

Theorem 2. Let ε >0andk≥0 be given. Assume that the initial dataΨ0 belongs toXk+6(R) and satisfies

kNε0kHk+5(R)+εk∂xk+6Nε0kL2(R)+k∂xΘ0εkHk+5(R)≤K0. (10) Let N± and W± denote the solutions to the Korteweg-de Vries equation3

τU±∓∂x3U±∓ U±xU± = 0,

with initial data Nε0, respectivelyxΘ0ε. There exists positive constants ε2 and K2, depending possibly on K0 and k, such that

kN±(·, τ)−Nε±(·, τ)kHk(R)+kW±(·, τ)−∂xΘ±ε(·, τ)kHk(R)

≤K2 ε2+kNε0±∂xΘ0εkHk(R)

expK2|τ|, (11) for any τ ∈R, provided ε≤ε2.

Remark 2. If the termkNε0±∂xΘ0εkHk(R) is small, then the Korteweg-de Vries approximation is valid on a time interval (in the original time variable) t∈[0, Tε] with

Tε=o

min

|log(ε)|

ε3 ,|log(kNε0±∂xΘ0εkHk(R))| ε3

.

In particular, if kNε0 ±∂xΘ0εkHk(R) ≤ Cεα, with α > 0, then the approximation is valid on a time interval t ∈ [0, Tε] with Tε = o(ε3|log(ε)|). Moreover, if kNε0 ±∂xΘ0εkHk(R) is of order O(ε2), then the approximation error remains of order O(ε2) on a time interval t∈[0, Tε′′] with Tε′′=O(ε3).

The main difference between Theorems 1 and 2 is that the second one involves the functions Nε± and ∂xΘ±ε instead of Uε±. The error term in (11) involves the quantity kNε0±∂xΘ0εkHk(R), which is small if the wave travelling in the other direction is small. Let us also emphasize that, in contrast with the assumptions of Theorem 1, the assumptions of Theorem 2 do not involve any assumption on the M-norm.

Finally, it is worthwhile to notice that similar issues have been addressed and solved in the case of the long-wave limit of the water wave system. As a matter of fact, system (8) bears some resemblance with a Boussinesq system. In a seminal work [10], Craig proved the first rigorous convergence result towards the Korteweg-de Vries equation, under assumptions which are similar in spirit to the ones in Theorem 2, focusing on a wave travelling in a single direction. Schneider and Wayne [15] completed the analysis and were able to handle both left and right-going waves at the same time, a result similar in spirit to Theorem 1. Bona, Colin and Lannes provided a sharp error estimate in [7] (see also [17]). The asymptotics were fully justified in [1], as well as in the higher dimensional case. In order to control the interactions between the two waves, sometimes called secular growth, these authors introduce additional assumptions on the initial data, of different nature but in the spirit, similar to our introduction of the M-norm which is more natural in our context.

We also emphasize that in the regime under study, the solutions of the Gross-Pitaevskii equation have their modulus close to one, so that by the Madelung transform, one is reduced to analyze a dispersive perturbation of a hyperbolic system. In this direction, Ben Youssef and Colin considered in [2] similar limits for general hyperbolic systems perturbed by linear dispersive terms.

3As well as the functionsU±, the functionsN±andW±depend onε. We again hide this dependence in the notations in order to stress the fact that the equations they satisfy are independent ofε.

(7)

With respect to those works, the main difficulty regarding system (8) is related to the fact that the dispersion terms are nonlinear. This difficulty is overcome using the integrability of the Gross-Pitaevskii and Korteweg-de Vries equations which allow to derive suitable bounds in high regularity spaces.

1.2 Some elements in the proofs

The proofs are somewhat parallel with the proofs in [6], so that we will try to emphasize the new ideas and ingredients. Concerning Theorem 1, the left and right going waves Uε and Uε+ play the same role in estimate (7), so that we may focus for instance on the estimates on Uε. In order to simplify our notation, we set Nε =Nε, Θε= Θε,Uε=Uε,U =U, and

x=x=ε(x +√ 2t).

We also introduce the new notation Vε(x, τ) = 1

2

Nε(x, τ)−∂xΘε(x, τ)

=Uε+ x− 8

ε2τ, τ

, (12)

and compute the relevant equations for Uε and Vε,

τUε+∂x3Uε+UεxUε =fε−ε2rε, (13) where

fε≡∂x

1

6Vε2−∂x2Vε+1 3UεVε

≡∂xFε, (14) and

τVε+ 8

ε2xVε=gε2rε, (15) where

gε≡∂x

x2Nε+1

2Vε2−1

6Uε2−1 3UεVε

≡∂xGε. (16) The remainder term rε is given by the formula

rε=∂x Nεx2Nε

6(1−ε62Nε) + (∂xNε)2 12(1− ε62Nε)2

≡∂xRε. (17) On the left-hand side of (13), we recognize the (KdV) equation, whereas on the left-hand side of (15), we recognize the transport operator with constant speed −8ε−2, which stems from the fact that we are working in moving frames with speed ±4ε2. It remains to establish that the terms on the right-hand side of (13) behave as error terms. The first step is to establish that all quantities are uniformly bounded on finite time intervals.

Proposition 2. Let k ∈ N. Given any ε > 0 sufficiently small, assume that the initial data Ψ0(·) = Ψ(·,0) belongs toXk+1(R) and satisfies the assumption

kNε0kHk(R)+εk∂xk+1Nε0kL2(R)+k∂xΘ0εkHk(R)≤K0, (18) where K0 is some given positive constant. Then, there exists a positive constant K depending only on K0 and k, such that

kNε(·, τ)kHk(R)+εk∂xk+1Nε(·, τ)kL2(R)+k∂xΘε(·, τ)kHk(R)≤KexpK|τ|, (19) for any τ ∈R. In particular, we have

kUε(·, τ)kHk(R)+kVε(·, τ)kHk(R)≤KexpK|τ|. (20)

(8)

Remark 3. Here and in the sequel, when we writeεsufficiently small, we mean that 0< ε≤ε0, where ε0 is some constant which depends only onK0, but not on the order of differentiationk.

In the course of our proofs, the constant ε0 is determined so that, when assumption (18) holds, the energy of Ψ is sufficiently small in order that (2.2) holds.

It follows from Proposition 2 that the quantity rε remains bounded on finite time intervals, so that the error term ε2rε is of order ε2 as desired in Theorem 1. In contrast, the term fε

describes the interaction of the two waves and is therefore more delicate to handle. Since Vε

is not supposed to be small in Theorem 1 (in contrast with Theorem 2), fε is not small in general. However, due to the dynamics, the interaction turns out to be of lower order. Indeed, in view of (15), at leading order, the function Vε is shifted to the left with speed 8ε−2. Since the definition of fεstrongly depends on the functionVε, a related property also holds forfε, so that the average interaction turns out to be small. To provide a rigorous justification of this last claim, we need however to localize the functions Uε and Vε. This leads us to use the norm k · kM(R).

As in [6], our proofs rely on energy methods. We introduce the differenceZε≡Uε− U, which satisfies the equation

τZε+∂x3Zε+U∂xZε+ZεxU +ZεxZε=fε−ε2rε. (21) In order to compute theL2-norm of∂xkZε, we apply the differential operator∂xkto (21), multiply the resulting equation by∂xkZε and integrate onR to obtain

1

2∂τk∂xkZεk2L2(R) =− Z

R

xk+1 UZε

xkZε−1 2

Z

R

xk+1 Zε2

xkZε +

Z

R

xkfεxkZε−ε2 Z

R

xkrεxkZε. Setting

Zεk(τ)≡ Z τ

0

Z

R

xkZε

2

, (22)

and integrating in time, we are led to the differential equation 1

2∂τZεk(τ) =− Z τ

0

Z

R

xk+1 UZε

xkZε−1 2

Z τ 0

Z

R

k+1x Zε2

xkZε

+ Z τ

0

Z

R

xkfεxkZε−ε2 Z τ

0

Z

R

xkrεkxZε.

(23)

The proof of Theorem 1 then follows applying the Gronwall lemma to (23) provided we are first able to bound suitably all the terms on the right-hand side of (23). The first, second and fourth terms can be handled thanks to Proposition 2. Indeed, we will show in Section 4 that these terms can be bounded as follows

Z τ 0

Z

R

xk+1 UZε

kxZε ≤K

Z τ 0

Z

R

xkZε2

+

Z τ 0

expK|s| kZε(·, s)kHk(R)ds

, (24)

Z τ 0

Z

R

xk+1 Zε2

xkZε ≤K

Z τ 0

Z

R

xkZε2

+

Z τ 0

expK|s| kZε(·, s)kHk(R)ds

, (25)

Z τ 0

Z

R

xkrεxkZε ≤K

Z τ 0

expK|s| kZε(·, s)kHk(R)ds

. (26)

For the third term, we will prove

(9)

Proposition 3. Let ε >0 be given sufficiently small. Given any k∈N, assume that the initial datumΨ0(·) = Ψ(·,0)belongs toXk+6(R)and satisfies (5)for some positive constantK0. Then, there exists a positive constant K depending only on K0 and k, such that

Z τ 0

Z

R

xkfεxkZε

≤Kε2

ε2+k∂xkZε(·, τ)kL2(R)

expK|τ|+

Z τ 0

expK|s| kZε(·, s)kHk(R)ds

,

(27)

for any τ ∈R.

Combining (23) with bounds (24), (25) and (27), we will obtain

τZεk(τ)≤K

sign(τ)Zεk(τ) +ε4expK|τ|

. The proof of Theorem 1 will then follow applying the Gronwall lemma.

We next say a few words about the proof of Proposition 3. For sake of clarity, we assume here that k= 0. For givenτ ∈R, we then have

Z τ 0

Z

R

fεZε= Z τ

0

Z

R

1

6∂xVε2−∂x3Vε+1

3∂x UεVε Zε

= Z τ

0

Z

R

xVε

1

3VεZε−∂x2Zε+ 1 3UεZε

+ 1 3

Z τ 0

Z

R

xUεVεZε.

(28)

For the first integral on the right-hand side, we use transport equation (15) and write

xVε= ε2 8

−∂τVε+gε2rε .

This change makes apparent an ε2 factor, and then we continue the computation using integra- tion by parts for the time and space variables as well as the bounds provided by Proposition 2. It remains to handle the second integral on the right-hand side of (28), which does not in- volve as before the spatial derivative∂xVε. We somewhat artificially introduce such a derivative considering an antiderivative Υε of Vε defined by

Υε(x, τ) = Z x

R

Vε(y, τ)dy, (29)

where the positive number R will be suitably chosen in the course of the computations. We

obtain Z τ

0

Z

R

xUεVεZε= Z τ

0

Z

R

xUεxΥεZε. The function Υε satisfies a transport equation, namely

τΥε+ 8

ε2xΥε=Gε2Rε−Gε(−R)−ε2Rε(−R) + 8

ε2Vε(−R), (30) so that it is possible to implement a similar argument as above replacing∂xΥεby the expression provided by (30). In order to perform our computation, it turns out that we only need to obtain some control in time of the L-norm of Υε, which essentially amounts to controlling the M- norm of Vε.At initial time, this corresponds to assumption (1) onNε0 and ∂xΘ0ε. For later time, we have

(10)

Lemma 1. Let 0≤E0 < 232 and Ψ0 ∈ M(R)∩X4(R) be given such that E(Ψ0)≤E0. Then, there exists a positive constant K, depending only on E0, such that

kη(·, t)kM(R)+√

2k∂xϕ(·, t)kM(R) ≤K

0kM(R)+√

2k∂xϕ0kM(R)

+

Z t 0

k∂x2η(·, s)kL(R)+kη(·, s)k2W2,∞(R)+k∂xϕ(·, s)k2L(R)

ds

,

(31)

for any t∈R.

In the slow coordinate t, Lemma 1 provides a control on the norm kΥεkL(R), which is independent of εand grows linearly in time.

Lemma 2. Letε >0be sufficiently small. Assume that the initial datumΨ0(·) = Ψ(·,0)belongs to X4(R) and satisfies assumption (18) for k= 3 and some positive constant K0. Then, there exists a positive constant K, which does not depend on εnor τ, such that

kNε(·, τ)kM(R)+k∂xΘε(·, τ)kM(R) ≤K

kNε0kM(R)+k∂xΘ0εkM(R)+|τ|

, (32) for any τ ∈R. In particular, we have

kUε(·, τ)kM(R)+kVε(·, τ)kM(R) ≤K

kUε0kM(R)+kVε0kM(R)+|τ|

. (33)

The proof of estimate (27) follows combining the bounds of Proposition 2 and Lemma 2.

We next give some elements of the proof of Theorem 2. Notice first that the functions Nε± and ∂xΘ±ε associated to the coordinates x+ and x play the same role in Theorem 2, so that we may focus again for the estimates on Nε ≡ Nε and ∂xΘε ≡ ∂xΘε. Recall that the main differences with respect to Theorem 1 are that Theorem 2 addresses the functionsNε and ∂xΘε

instead of Uε, and does not involve any assumption on M-norms. The proof of Theorem 2 is however parallel to the one of Theorem 1. We write for the function Nε,

kNε− N kHk(R)≤ kVεkHk(R)+kUε− UkHk(R)+kU − N kHk(R), (34) since by definition,Vε=Nε−Uε. Similarly, the functions∂xΘε and W satisfy

k∂xΘε− WkHk(R)≤ kVεkHk(R)+kUε− UkHk(R)+kU − WkHk(R), (35) so that the proof of (11) reduces to bound each of the terms on the right-hand side of (34) and (35). For the Hk-norm of Vε, we invoke again energy estimates, based now on equation (15).

This yields

Proposition 4. Letε >0be sufficiently small. Given anyk∈N, assume that the initial datum Ψ0(·) = Ψ(·,0)belongs toXk+6(R)and satisfies (10)for some positive constantK0. Then, there exists a positive constant K depending only on K0 and k, such that

kVε(·, τ)kHk(R)≤K kVε0kHk(R)2

expK|τ|, (36)

for any τ ∈R.

Similarly, concerning the differences between the solutions to (KdV),U andN on one hand, and U and W on the other, we invoke a general stability result for the Korteweg-de Vries equation, which we recall for the sake of completeness. The proof follows from standard Hk- energy methods for the difference, using the conserved quantities of (KdV) in order to bound uniformly the Hk+1-norms coming from the quadratic terms.

(11)

Lemma 3. Let k∈N be given. Consider two functions F0 and G0 in Hk+2(R) and denote F andGthe solutions to(KdV)with initial dataF0, respectivelyG0. Then, there exists a constant K, depending only onk and the Hk+2-norms ofF0 and G0, such that

kF(·, τ)−G(·, τ)kHk(R)≤KkF0−G0kHk(R)expK|τ|, (37) for any τ ∈R.

In view of Lemma 3, and the fact that

U0− W0 =N0− U0=Vε0, we are led to

kU(·, τ)− N(·, τ)kHk(R)+kU(·, τ)− W(·, τ)kHk(R)≤KkVε0kHk(R)expK|τ|, (38) for any τ ∈R. Going back to (35), it remains to estimate the termkUε− UkHk(R)=kZεkHk(R). An upper bound for this term was given in Theorem 1 using bounds on theM-norm ofNε and

xΘε.Here, we use instead the estimates of Proposition 4 to bound the interaction terms.

Proposition 5. Let ε >0 be given sufficiently small. Given any k∈N, assume that the initial datum Ψ0(·) = Ψ(·,0) belongs to Xk+6(R) and satisfies (10) for some positive constant K0. Then, there exists a positive constant K depending only on K0 and k, such that

Z τ 0

Z

R

xkfεxkZε

≤Kε2

ε2+kVε0kHk(R)+k∂xkZε(·, τ)kL2(R)

expK|τ|

+K

ε2+kVε0kHk(R)

Z τ 0

expK|s| kZε(·, s)kHk(R)ds ,

(39)

for any τ ∈R.

We then adapt the arguments of the proof of (7) in order to obtain that kUε(·, τ)− U(·, τ)kHk(R) ≤K ε2+kVε0kHk(R)

expK|τ|, (40) for any τ ∈R. Combining with inequalities (34) and (35), and bounds (36) and (38), this will complete the proof of Theorem 2.

1.3 Outline of the paper

This paper is organized as follows. In the next section, we provide the proofs to Proposition 2, and Lemmas 1 and 2. In Section 3, we prove Proposition 3. The proof of Theorem 1 is completed in Section 4, whereas we derive Proposition 4, Lemma 3, Proposition 5, and finally Theorem 2 in Section 5. In a separate appendix, we extend the arguments of the proof of Lemma 1 to provide a rigorous framework for the notion of mass and establish its conservation.

2 Bounds for the rescaled functions

In this section, we establish a certain number of bounds for the rescaled functions Nε,∂xΘε,Uε and Vε, which are useful in the course of the proofs of Theorems 1 and 2. We first derive the Sobolev bounds of Proposition 2, and then we compute estimate (32) of Lemma 2.

(12)

2.1 Sobolev bounds

Two different arguments are under hand to derive the Sobolev bounds given by inequality (19).

The first one relies on the integrability properties of (GP). It is proved in [6] that the quantities E1(Ψ) ≡E(Ψ),

E2(Ψ)≡ 1 2

Z

R

|∂x2Ψ|2− 3 2

Z

R

η|∂xΨ|2+1 4

Z

R

(∂xη)2− 1 4

Z

R

η3,

E3(Ψ)≡1 2

Z

R|∂3xΨ|2+1 4

Z

R|∂x2η|2+5 4

Z

R|∂xΨ|4+5 2

Z

R

x2η|∂xΨ|2−5 2

Z

R

η|∂2xΨ|2

−5 4

Z

R

η(∂xη)2+15 4

Z

R

η2|∂xΨ|2+ 5 16

Z

R

η4, and

E4(Ψ)≡ 1 2

Z

R

|∂x4Ψ|2+1 4

Z

R

|∂x3η|2−7 4

Z

R

η(∂x2η)2−7 2

Z

R

η|∂x3Ψ|2+35 8

Z

R

η2(∂xη)2 +35

4 Z

R

η2|∂x2Ψ|2−35 4

Z

R

(∂xη)2|∂xΨ|2−7 2

Z

R

|∂xΨ|2|∂x2Ψ|2−7 Z

R

2xηh∂xΨ, ∂x3Ψi

−7 Z

R

|∂xΨ|2h∂xΨ, ∂3xΨi − 35 2

Z

R

η∂x2η|∂xΨ|2−35 4

Z

R

η3|∂xΨ|2−35 4

Z

R

η|∂xΨ|4− 7 16

Z

R

η5, are conserved along the Gross-Pitaevskii flow, provided that the initial datum Ψ0 belongs to X4(R). Notice that, for 1 ≤k ≤ 4, the quantities Ek defined above give a control on the L2- norms of the functions ∂xkΨ and ∂kx−1η. As a matter of fact, invoking the Sobolev embedding theorem, one can establish that there exists some universal constant K such that

Ek(ψ)≤K

kψkHk(R)+k1− |ψ|2kHk−1(R)

k+1

,

for any function ψ∈Xk(R). Similarly, there exists a positive constantK(E1(ψ), . . . , Ek1(ψ)), depending only on the quantities E1(ψ), . . .and Ek1(ψ), such that

kψk2Hk(R)+k1− |ψ|2k2Hk−1(R)≤K(E1(ψ), . . . , Ek1(ψ))Ek(ψ).

In particular, the conservation of the quantities Ek(Ψ) along the Gross-Pitaevskii flow provides bounds on the Sobolev norms of the functions Ψ and η, which only depends on the Xk-norms of the initial datum Ψ0. In the rescaled variables, we obtain

Proposition 2.1 ([6]). Let 0 ≤k ≤3 and ε > 0 be given sufficiently small. Assume that the initial datum Ψ0(·) = Ψ(·,0)belongs toXk+1(R) and satisfies assumption (18)for some positive constant K0. Then, there exists a positive constant K, which does not depend on εnor τ, such that

kNε(·, τ)kHk(R)+εk∂xk+1Nε(·, τ)kL2(R)+k∂xΘε(·, τ)kHk(R)≤K, (2.1) for any τ ∈R.

Inequality (2.1) presents the advantage to be uniform in time. We will invoke this property to derive Lemma 2. We believe that inequality (2.1) remains valid for higher order Sobolev spaces. However, it seems rather involved to prove this claim since this requires to compute, or at least to describe precisely, the higher order invariants of (GP) (see [6] for more details).

In order to compute higher Sobolev bounds, we rely on the energy estimates derived in [3] in the context of the wave limit of the Gross-Pitaevskii equation mentioned above in the

(13)

introduction. More precisely, given any k∈Nand any positive number E0 < 232, we consider some initial datum Ψ0 ∈Xk(R) such that E(Ψ0)≤E0. Then, there exists a positive constant m0, depending only onE0, such that

m0 ≤ |Ψ(x, t)| ≤ 1

m0, (2.2)

for any (x, t)∈R2 (see [4] for more details). Under this additional assumption, the computation of energy estimates for system (8) achieved in Proposition 1 of [3], provides the tame estimates

tΓkε(t)≤K(k, m02 1+ε2knε(·, t)kL

k∂xnε(·, t)kL+k∂xyε(·, t)kL

Γkε(t)+Γ0ε(t) 8

, (2.3) where K(k, m0) is some positive constant which does not depend on ε nor t, and where the functionyεis defined by

yε=wε+ iε

√2

xnε

mε , where mε= 1− ε2 6nε, while the notation Γkε(t) refers to the functional

Γkε(t)≡ k∂xknε(·, t)k2L2(R)+kmε(·, t)12xkyε(·, t)k2L2(R). (2.4) In view of (2.2), and since we have

mε(x, t) = Ψx

ε,t ε

,

the quantity Γkε controls theHk-norms ofnεandyε. Going back to the original setting, we have in particular,

Γ0ε(t) = 144 ε3 E

Ψ

·,t ε

, so that, by the conservation of the energy,

Γ0ε(t) = 144

ε3 E(Ψ0). (2.5)

Proof of Proposition 2. In the case k <4, Proposition 2 is a direct consequence of Proposition 2.1. In the case k ≥ 4, the proof of Proposition 2 is obtained applying the Gronwall lemma to inequality (2.3), using identity (2.5) and bounds (2.1). We conclude rescaling the derived inequality in the variables Nε and∂xΘε.

More precisely, invoking assumption (18) fork= 0, we first compute in the rescaled setting, E(Ψ0) = ε3

144 Z

R

Mε0(∂xΘ0ε)2+ (Nε0)22(∂xNε0)2 2Mε0

≤Kε3, (2.6)

where Mε0 ≡1−ε62Nε0 and K is a positive constant depending only onK0. In particular, given any εsufficiently small, assumption (2.2) holds for m0 = 12, so that in view of (2.3) and (2.5), we are led to

Γkε(t)≤Γkε(0) exp

Z t 0

Aε(s)ds

+18 ε3E(Ψ0)

exp

Z t 0

Aε(s)ds −1

, (2.7)

where

Aε(t)≡K(k, m02 1 +εknε(·, t)kL(R)

k∂xnε(·, t)kL(R)+k∂xyε(·, t)kL(R)

. (2.8)

Références

Documents relatifs

The main goal of this paper, therefore is to establish by rigorous singular perturbation theory a simple numerical condition guaranteeing the existence of periodic traveling

In the present study, we investigated the status of the L-arginine/NO pathway in 34 healthy individuals aged 2 days to 24 years by measuring ADMA, L-argi- nine, nitrate and nitrite

Dans cet article, nous avons proposé plusieurs méthodes d’intégration d’un champ de gradient, qui sont rapides et robustes aux discontinuités de profondeur, grâce à une

/ La version de cette publication peut être l’une des suivantes : la version prépublication de l’auteur, la version acceptée du manuscrit ou la version de l’éditeur.. Access

[r]

On peut s'étonner du fait que si peu de maisons mobiles non ancrées aient été détruites dans le passé, mais il ne faut pas oublier que de nombreuses maisons mobiles sont situées

In cultured mouse primary neurons, crebinostat potently induced acetylation of both histone H3 and histone H4 as well as enhanced the expression of the CREB target gene Egr1

We propose two architectures, both based on parallel leaf springs, then evaluate their isotropy defect using firstly an analytic model, secondly finite element anal- ysis and