• Aucun résultat trouvé

Sharp ill-posedness and well-posedness results for the KdV-Burgers equation: the real line case

N/A
N/A
Protected

Academic year: 2021

Partager "Sharp ill-posedness and well-posedness results for the KdV-Burgers equation: the real line case"

Copied!
34
0
0

Texte intégral

(1)

HAL Id: hal-00436652

https://hal.archives-ouvertes.fr/hal-00436652v2

Submitted on 31 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

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

Sharp ill-posedness and well-posedness results for the KdV-Burgers equation: the real line case

Luc Molinet, Stéphane Vento

To cite this version:

Luc Molinet, Stéphane Vento. Sharp ill-posedness and well-posedness results for the KdV-Burgers

equation: the real line case. Annali della Scuola Normale Superiore di Pisa, Classe di Scienze, Scuola

Normale Superiore 2011, 10 (3), pp.531-560. �hal-00436652v2�

(2)

Sharp ill-posedness and well-posedness results for the KdV-Burgers equation: the real line case

Luc Molinet and St´ ephane Vento

Abstract. We complete the known results on the Cauchy problem in Sobolev spaces for the KdV-Burgers equation by proving that this equation is well-posed in H

−1

( R ) with a solution-map that is analytic from H

−1

( R ) to C([0, T ]; H

−1

( R )) whereas it is ill-posed in H

s

( R ), as soon as s < −1, in the sense that the flow-map u

0

7→ u(t) cannot be continuous from H

s

( R ) to even D

( R ) at any fixed t > 0 small enough. As far as we know, this is the first result of this type for a dispersive-dissipative equation. The framework we develop here should be useful to prove similar results for other dispersive- dissipative models.

1 Introduction and main results

The aim of this paper is to establish positive and negative optimal results on the local Cauchy problem in Sobolev spaces for the Korteweg-de Vries- Burgers (KdV-B) equation posed on the real line :

u

t

+ u

xxx

− u

xx

+ uu

x

= 0 (1.1) where u = u(t, x) is a real valued function.

This equation has been derived as an asymptotic model for the propagation of weakly nonlinear dispersive long waves in some physical contexts when dissipative effects occur (see [17]). It thus seems natural to compare the well-posedness results on the Cauchy problem for the KdV-B equation with the ones for the Korteweg-de-Vries (KdV) equation

u

t

+ u

xxx

+ uu

x

= 0 (1.2)

that correspond to the case when dissipative effects are negligible and for the dissipative Burgers (dB) equation

u

t

− u

xx

+ uu

x

= 0 (1.3)

that corresponds to the case when dissipative effect are dominant.

(3)

To make this comparison more transparent it is convenient to define different notions of well-posedness (and consequently ill-posedness) related to the smoothness of the flow-map (see in the same spirit [13], [8]).

Throughout this paper we shall say that a Cauchy problem is (locally) C

0

-well-posed in some normed function space X if, for any initial data u

0

∈ X, there exist a radius R > 0, a time T > 0 and a unique solu- tion u, belonging to some space-time function space continuously embedded in C([0, T ]; X), such that for any t ∈ [0, T ] the map u

0

7→ u(t) is continuous from the ball of X centered at u

0

with radius R into X. If the map u

0

7→ u(t) is of class C

k

, k ∈ N ∪ {∞}, (resp. analytic) we will say that the Cauchy is C

k

-well-posed (resp. analytically well-posed). Finally a Cauchy problem will be said to be C

k

-ill-posed, k ∈ N ∪ {∞}, if it is not C

k

-well-posed.

For the KdV equation on the line the situation is as follows: it is an- alytically well-posed in H

−3/4

( R ) (cf. [14] and [11] for the limit case) and C

3

-ill-posed below this index

1

(cf. [4]). On the other hand the results for the dissipative Burgers equation are much clear. Indeed this equation is known to be analytically well-posed in H

s

( R ) for s ≥ −1/2 (cf [7] and [1] for the limit case) and C

0

ill-posed in H

s

for s < −1/2 (cf. [7] ). At this stage it is interesting to notice that the critical Sobolev exponents obtained by scaling considerations are respectively −3/2 for the KdV equation and −1/2 for the dissipative Burgers equation. Hence for the KdV equation there is an im- portant gap between this critical exponent and the best exponent obtained for well-posedness.

Now, concerning the KdV-Burgers equation, Molinet and Ribaud [16]

proved that this equation is analytically well-posed in H

s

( R ) as soon as s > −1. They also established that the index −1 is critical for the C

2

- well-posedness. The surprising part of this result was that, according to the above results, the C

critical index s

c

(KdV B) = −1 was lower that the one of the KdV equation s

c

(KdV ) = −3/4 and also lower than the C

index s

c

(dB) = −1/2 of the dissipative Burgers equation.

In this paper we want in some sense to complete this study by proving that the KdV-Burgers equation is analytically well-posed in H

−1

( R ) and C

0

-ill-posed in H

s

( R ) for s < −1 in the sense that the flow-map defined on H

−1

( R ) is not continuous for the topology inducted by H

s

, s < −1, with values even in D

( R ). It is worth emphasizing that the critical index s

0c

= −1 is still far away from the critical index s

c

= −3/2 given by the scaling symmetry of the KdV equation. We believe that this result strongly

1See also [5] where it is proven that the solution-map is even not uniformly continuous on bounded sets below this index

(4)

suggest that the KdV equation should also be C

0

-ill-posed in H

s

( R ) for s < −1.

To reach the critical Sobolev space H

−1

( R ) we adapt the refinement of Bourgain’s spaces that appeared in [20] and [19] to the framework de- veloped in [16]. One of the main difficulty is related to the choice of the extension for negative times of the Duhamel operator (see the discussion in the beginning of Section 4). The approach we develop here to overcome this difficulty should be useful to prove optimal results for other dispersive- dissipative models. The ill-posedness result is due to a high to low frequency cascade phenomena that was first observed in [2] for a quadratic Schr¨ odinger equation..

At this stage it is worth noticing that, using the integrability theory, it was recently proved in [13] that the flow-map of KdV equation can be uniquely continuously extended in H

−1

( T ). Therefore, on the torus, KdV is C

0

-well-posed in H

−1

if one takes as uniqueness class, the class of strong limit in C([0, T ]; H

−1

( T )) of smooth solutions. In the present work we use in a crucial way the global Kato smoothing effect that does not hold on the torus. However, in a forthcoming paper ([15]) we will show how one can modify the approach developed here to prove that the same results hold on the torus, i.e. analytic well-posedness in H

−1

( T ) and C

0

-ill-posedness in H

s

( T ) for s < −1. In view of the result of Kappeler and Topalov for KdV it thus appears that, at least on the torus, even if the dissipation part of the KdV-Burgers equation (it is important to notice that the dissipative term −u

xx

is of lower order than the dispersive one u

xxx

) allows to lower the C

critical index with respect to the KdV equation, it does not permit to improve the C

0

critical index .

Our results can be summarized as follows:

Theorem 1.1. The Cauchy problem associated to (1.1) is locally analyti- cally well-posed in H

−1

( R ). Moreover, at every point u

0

∈ H

−1

( R ) there exist T = T(u

0

) > 0 and R = R(u

0

) > 0 such that the solution-map u

0

7→ u is analytic from the ball centered at u

0

with radius R of H

−1

( R ) into C([0, T ]; H

−1

( R )). Finally, the solution u can be extended for all posi- tive times and belongs to C( R

+

; H

( R )).

Theorem 1.2. The Cauchy problem associated to (1.1) is ill-posed in H

s

( R ) for s < −1 in the following sense: there exists T > 0 such that for any 0 <

t < T , the flow-map u

0

7→ u(t) constructed in Theorem 1.1 is discontinuous

at the origin from H

−1

( R ) endowed with the topology inducted by H

s

( R )

into D

( R ).

(5)

Acknowlegements: L.M. was partially supported by the ANR project

”Equa-Disp”.

2 Ill-posedness

The ill-posedness result can be viewed as an application of a general result proved in [2]. Roughly speaking this general ill-posedness result requires the two following ingredients:

1. The equation is analytically well-posed until some index s

c

with a solution-map that is also analytic.

2. Below this index one iteration of the Picard scheme is not continuous.

The discontinuity should be driven by high frequency interactions that blow up in frequencies of order least or equal to one.

The first ingredient is given by Theorem 1.1 whereas the second one has been derived in [16] where the discontinuity of the second iteration of the Picard scheme in H

s

( R ) and H

s

( T ) for s < −1 is established.

However, due to the nature of the equation, our result is a little better than the one given by the general theory developed in [2]. Indeed, we will be able to prove the discontinuity of the flow-map u

0

7→ u(t) for any fixed t > 0 less than some T > 0 and not only of the solution-map u

0

7→ u. Therefore for sake of completeness we will prove the result with hand here.

Let us first recall the counter-example constructed in [16] that we renor- malize here in H

−1

( R ). We define the sequence of initial data {φ

N

}

N≥1

by

φ ˆ

N

= N

−1

χ

IN

(ξ) + χ

IN

(−ξ)

, (2.1)

where I

N

= [N, N + 2] and ˆ φ

N

denotes the space Fourier transform of φ

N

. Note that kφ

N

k

H−1(R)

∼ 1 and φ

N

→ 0 in H

s

( R ) for s < −1. This sequence yields a counter-example to the continuity of the second iteration of the Picard Scheme in H

s

( R ), s < −1, that is given by

A

2

(t, h, h) = Z

t

0

S(t − t

)∂

x

[S(t

)h]

2

dt

(6)

where S is the semi-group associated to the linear part of (1.1) (see (3.2)).

Indeed, computing the space Fourier transform we get F

x

(A

2

(t, φ

N

, φ

N

))(ξ) =

Z

R

e

−tξ2

e

itξ3

φ ˆ

N

1

) ˆ φ

N

(ξ − ξ

1

) (iξ)

Z

t

0

e

−(ξ21+(ξ−ξ1)2−ξ2)t

e

i(ξ31+(ξ−ξ1)3−ξ3)t

dt

1

= (iξ) e

itξ3

e

−tξ2

Z

R

φ ˆ

N

1

) ˆ φ

N

(ξ − ξ

1

)

e

−(ξ12+(ξ−ξ1)2−ξ2)t

e

i3ξξ1(ξ−ξ1)t

− 1

−2ξ

1

(ξ − ξ

1

) + i3ξξ

1

(ξ − ξ

1

) dξ

1

, so that

kA

2

(t, φ

N

, φ

N

)k

2Hs

≥ Z

1/2

−1/2

(1 + |ξ|

2

)

s

|F

x

(A

2

(t, φ

N

, φ

N

))(ξ)|

2

= N

4

Z

1/2

−1/2

(1 + |ξ|

2

)

s

|ξ|

2

Z

Kξ

e

−(ξ21+(ξ−ξ1)2)t

e

i3ξξ1(ξ−ξ1)t

− e

−ξ2t

−2ξ

1

(ξ − ξ

1

) + i3ξξ

1

(ξ − ξ

1

) dξ

1

2

dξ ,

where

K

ξ

= {ξ

1

/ ξ − ξ

1

∈ I

N

, ξ

1

∈ −I

N

} ∪ {ξ

1

/ ξ

1

∈ I

N

, ξ − ξ

1

∈ −I

N

} . Note that for any ξ ∈ [−1/2, 1/2], one has mes(K

ξ

) ≥ 1 and

3ξξ

1

(ξ − ξ

1

) ∼ N

2

1

(ξ − ξ

1

) ∼ N

2

, ∀ξ

1

∈ K

ξ

. Therefore, fixing 0 < t < 1 we have

Re (e

−(ξ21+(ξ−ξ1)2)t

e

i3ξξ1(ξ−ξ1)t

− e

−ξ2t

) ≤ −e

−t/4

+ e

−2(N+2)2t

, which leads for N = N (t) > 0 large enough to

Z

Kξ

e

−(ξ21+(ξ−ξ1)2)t

e

i3ξξ1(ξ−ξ1)t

− e

−ξ2t

−2ξ

1

(ξ − ξ

1

) + i3ξξ

1

(ξ − ξ

1

) dξ

1

≥ C e

−t/4

N

2

and thus

kA

2

(t, φ

N

, φ

N

)k

2Hs

≥ Ce

−t/4

≥ C

0

(2.2)

(7)

for some positive constant C

0

> 0. Since φ

N

→ 0 in H

s

( R ), for s < −1, this ensures that, for any fixed t > 0, the map u

0

7→ A

2

(t, u

0

, u

0

) is not continuous at the origin from H

s

( R ) into D

( R ).

Now, we will use that A

2

(t, φ

N

, φ

N

) is of order at least one in H

s

( R ) to prove that somehow A

2

(t, εφ

N

, εφ

N

) is the main contribution to u(t, εφ

N

) in H

s

( R ) as soon as s < −1, ε > 0 is small and N is large enough. The discontinuity of u

0

7→ u(t) will then follow from the one of u

0

7→ A

2

(t, u

0

, u

0

).

According to Theorem 1.1 there exist T > 0 and ε

0

> 0 such that for any

|ε| ≤ ε

0

, any khk

H−1(R)

≤ 1 and 0 ≤ t ≤ T , u(t, εh) = εS(t)h +

+∞

X

k=2

ε

k

A

k

(t, h

k

)

where h

k

:= (h, . . . , h), h

k

7→ A

k

(t, h

k

) is a k-linear continuous map from

H

−1

( R )

k

into C([0, T ]; H

−1

( R )) and the series converges absolutely in C([0, T ]; H

−1

( R )).

In particular,

u(t, εφ

N

) − ε

2

A

2

(t, φ

N

, φ

N

) = εS(t)φ

N

+

+∞

X

k=3

ε

k

A

k

(t, φ

kN

) . On the other hand, kS(t)φ

N

k

Hs(R)

≤ kφ

N

k

Hs(R)

∼ N

1+s

and

X

∞ k=3

ε

k

A

k

(t, φ

kN

)

H−1

≤ ε ε

0

3

X

∞ k=3

ε

k0

kA

k

(t, φ

N

)k

H−1

≤ Cε

3

. Hence, for s < −1,

sup

t∈[0,T]

u(t, εφ

N

) − ε

2

A

2

(t, φ

N

, φ

N

)

Hs(R)

≤ Cε

3

+ O(N

1+s

) .

In view of (2.2) this ensures that, fixing 0 < t < 1 and taking ε small enough and N large enough, ε

2

A

2

(t, φ

N

, φ

N

) is a “good” approximation of u(t, εφ

N

). In particular, taking ε ≤ C

0

C

−1

/4 we get

ku(t, εφ

N

)k

Hs(R)

≥ C

0

ε

2

/2 + O(N

1+s

) .

Since u(t, 0) ≡ 0 and φ

N

→ 0 in H

s

( R ) for s < −1 this leads to the

discontinuity of the flow-map at the origin by letting N tend to infinity. It

is worth noticing that since φ

N

⇀ 0 in H

−1

( R ) we also get that u

0

7→ u(t, u

0

)

is discontinuous from H

−1

( R ) equipped with its weak topology with values

even in D

( R ).

(8)

3 Resolution space

In this section we introduce a few notation and we define our functional framework.

For A, B > 0, A . B means that there exists c > 0 such that A ≤ cB.

When c is a small constant we use A ≪ B. We write A ∼ B to denote the statement that A . B . A. For u = u(t, x) ∈ S

( R

2

), we denote by u b (or F

x

u) its Fourier transform in space, and u e (or F u) the space-time Fourier transform of u. We consider the usual Lebesgue spaces L

p

, L

px

L

qt

and abbreviate L

px

L

pt

as L

p

. Let us define the Japanese bracket hxi = (1+|x|

2

)

1/2

so that the standard non-homogeneous Sobolev spaces are endowed with the norm kfk

Hs

= kh∇i

s

f k

L2

.

We also need a Littlewood-Paley analysis. Let η ∈ C

0

( R ) be such that η ≥ 0, supp η ⊂ [−2, 2], η ≡ 1 on [−1, 1]. We define next ϕ(ξ) = η(ξ) −η(2ξ).

Any summations over capitalized variables such as N, L are presumed to be dyadic, i.e. these variables range over numbers of the form 2

, ℓ ∈ Z . We set ϕ

N

(ξ) = ϕ(ξ/N ) and define the operator P

N

by F(P

N

u) = ϕ

N

b u. We introduce ψ

L

(τ, ξ) = ϕ

L

(τ − ξ

3

) and for any u ∈ S

( R

2

),

F

x

(P

N

u(t))(ξ) = ϕ

N

(ξ)ˆ u(t, ξ), F(Q

L

u)(τ, ξ) = ψ

L

(τ, ξ)˜ u(τ, ξ).

Roughly speaking, the operator P

N

localizes in the annulus {|ξ| ∼ N } whereas Q

L

localizes in the region {|τ − ξ

3

| ∼ L}.

Furthermore we define more general projection P

.N

= P

N1.N

P

N1

, Q

≫L

= P

L1≫L

Q

L1

etc.

Let e

−t∂xxx

be the propagator associated to the Airy equation and define the two parameters linear operator W by

F

x

(W (t, t

)φ)(ξ) = exp(itξ

3

− |t

2

) ˆ φ(ξ), t ∈ R . (3.1) The operator W : t 7→ W (t, t) is clearly an extension to R of the linear semi-group S(·) associated with (1.1) that is given by

F

x

(S(t)φ)(ξ) = exp(itξ

3

− tξ

2

) ˆ φ(ξ), t ∈ R

+

. (3.2) We will mainly work on the integral formulation of (1.1):

u(t) = S(t)u

0

− 1 2

Z

t 0

S(t − t

)∂

x

u

2

(t

)dt

, t ∈ R

+

. (3.3)

Actually, to prove the local existence result, we will apply a fixed point argu-

ment to the following extension of (3.3) (See Section 4 for some explanations

(9)

on this choice).

u(t) = η(t) h

W (t)u

0

− 1 2 χ

R+

(t)

Z

t 0

W (t − t

, t − t

)∂

x

u

2

(t

)dt

− 1 2 χ

R

(t)

Z

t 0

W (t − t

, t + t

)∂

x

u

2

(t

)dt

i

. (3.4) If u solves (3.4) then u is a solution of (3.3) on [0, T ], T < 1.

In [16], the authors performed the iteration process in the space X

s,b

equipped with the norm

kuk

Xs,b

= khi(τ − ξ

3

) + ξ

2

i

b

hξi

s

uk e

L2

which take advantage of the mixed dispersive-dissipative part of the equa- tion. In order to handle the endpoint index s = −1 without encountering logarithmic divergence, we will rather work in its Besov version X

s,b,q

(with q = 1) defined as the weak closure of the test functions that are uniformly bounded by the norm

kuk

Xs,b,q

= X

N

h X

L

hN i

sq

hL + N

2

i

bq

kP

N

Q

L

uk

qL2 xt

i

2/q

1/2

.

This Besov refinement, which usually provides suitable controls for nonlinear terms, is not sufficient here to get the desired bound especially in the high- high regime, where the nonlinearity interacts two components of the solution u with the same high frequency. To handle these divergences, inspired by [19], we introduce, for b ∈ {

12

, −

12

}, the space Y

s,b

endowed with the norm

kuk

Ys,b

= X

N

[hN i

s

kF

−1

[(i(τ − ξ

3

) + ξ

2

+ 1)

b+1/2

ϕ

N

e u]k

L1

tL2x

]

2

1/2

,

so that

kuk

Y−1,12

∼ X

N

[hN i

−1

k(∂

t

+ ∂

xxx

− ∂

xx

+ I)P

N

uk

L1

tL2x

]

2

1/2

.

Next we form the resolution space S

s

= X

s,12,1

+Y

s,12

, and the ”nonlinear space” N

s

= X

s,−12,1

+ Y

s,−12

in the usual way:

kuk

X+Y

= inf{ku

1

k

X

+ ku

2

k

Y

: u

1

∈ X, u

2

∈ Y, u = u

1

+ u

2

}.

In the rest of this section, we study some basic properties of the function

space S

−1

.

(10)

Lemma 3.1. For any φ ∈ L

2

, X

L

[L

1/2

kQ

L

(e

−t∂xxx

φ)k

L2

]

2

1/2

. kφk

L2

. Proof. From Plancherel theorem, we have

X

L

[L

1/2

kQ

L

(e

−t∂xxx

φ)k

L2

]

2

1/2

∼ k|τ − ξ

3

|

1/2

F(e

−t∂xxx

φ)k

L2

.

Moreover if we set η

T

(t) = η(t/T ) for T > 0, then

F (η

T

(t)e

−t∂xxx

φ)(τ, ξ) = η c

T

(τ − ξ

3

) φ(ξ). b

Thus we obtain with the changes of variables τ − ξ

3

→ τ

and T τ

→ σ that k|τ − ξ

3

|

1/2

F(η

T

(t)e

−t∂xxx

φ)k

L2

. kφk

L2

k|τ

|

1/2

T η(T τ b

)k

L2

τ

. kφk

L2

. Taking the limit T → ∞, this completes the proof.

Lemma 3.2. 1. For each dyadic N , we have k(∂

t

+ ∂

xxx

)P

N

uk

L1

tL2x

. kP

N

uk

Y0,12

. (3.5) 2. For all u ∈ S

−1

,

kuk

L2

xt

. kuk

S−1

. (3.6)

3. For all u ∈ S

0

,

X

L

[L

1/2

kQ

L

uk

L2

]

2

1/2

. kuk

S0

. (3.7) Proof. 1. From the definition of Y

0,12

, the right-hand side of (3.5) can be

rewritten as kP

N

uk

Y0,12

= k(∂

t

+ ∂

xxx

− ∂

xx

+ I )P

N

uk

L1

tL2x

.

Thus, by the triangle inequality, we reduce to show (3.5) with ∂

t

+∂

xxx

replaced by I − ∂

xx

. Using Plancherel theorem as well as Young and H¨older inequalities, we get

k(I − ∂

xx

)P

N

uk

L1

tL2x

. F

t−1

ξ

2

+ 1

i(τ − ξ

3

) + ξ

2

+ 1 (i(τ − ξ

3

) + ξ

2

+ 1)ϕ

N

u e

L1tL2ξ

.

(11)

In the sequel, it will be convenient to write ϕ

N

for ϕ

N/2

+ ϕ

N

+ ϕ

2N

. With this slight abuse of notation, we obtain

k(I − ∂

xx

)P

N

uk

L1tL2x

.

F

t−1

ϕ

N

(ξ)(ξ

2

+ 1) i(τ − ξ

3

) + ξ

2

+ 1

L1tLξ

k(∂

t

+ ∂

xxx

− ∂

xx

+ I )P

N

uk

L1

tL2x

. On the other hand, a direct computation yields

F

t−1

ϕ

N

(ξ)(ξ

2

+ 1) i(τ − ξ

3

) + ξ

2

+ 1

= Cϕ

N

(ξ)(1 + ξ

2

)e

−t(1+ξ2)

χ

R+

(t)

so that

F

t−1

ϕ

N

(ξ)(ξ

2

+ 1) i(τ − ξ

3

) + ξ

2

+ 1

L1tLξ

. khNi

2

e

−thNi2

χ

R+

(t)k

L1

t

. 1, and the claim follows.

2. We show that for any fixed dyadic N , we have

kP

N

uk

L2

. kP

N

uk

S−1

. (3.8) Estimate (3.6) then follows after square-summing. Observe that (3.8) follows immediately from the estimate hN i

−1

hL + N

2

i

1/2

& 1 if the right-hand side is replaced by kP

N

uk

X−1,12,1

, so it suffices to prove (3.8) with kP

N

uk

Y−1,12

in the right-hand side. But applying again Young and H¨older’s inequalities, this is easily verified:

kP

N

uk

L2

= F

t−1

1

i(τ − ξ

3

) + ξ

2

+ 1 (i(τ − ξ

3

) + ξ

2

+ 1)ϕ

N

e u

L2

.

F

t−1

ϕ

N

(ξ) i(τ − ξ

3

) + ξ

2

+ 1

L2tLx

kP

N

uk

Y0,12

. ke

−thNi2

χ

R+

(t)k

L2

t

kP

N

uk

Y0,12

. hNi

−1

kP

N

uk

Y0,12

. kP

N

uk

Y−1,12

. 3. First it is clear from definitions that P

L

[L

1/2

kQ

L

uk

L2

]

2

1/2

. kuk

X0,12,1

. Setting now v = (∂

t

+ ∂

xxx

)u, we see that u can be rewritten as

u(t) = e

−t∂xxx

u(0) + Z

t

0

e

−(t−t)∂xxx

v(t

)dt

.

(12)

By virtue of Lemma 3.1, we have X

L

[L

1/2

kQ

L

e

−t∂xxx

u(0)k

L2

]

2

1/2

. ku(0)k

L2

. kuk

Lt L2x

.

Moreover, we get as previously kuk

Lt L2x

.

F

t−1

1

i(τ − ξ

3

) + ξ

2

+ 1

L

kuk

Y0,12

. kuk

Y0,12

. (3.9) Thanks to estimate (3.5), it remains to show that

X

L

h L

1/2

Q

L

Z

t

0

e

−(t−t)∂xxx

v(t

)dt

L2

i

2

1/2

. kvk

L1tL2x

. (3.10) In order to prove this, we split the integral R

t

0

= R

t

−∞

− R

0

−∞

. By Lemma 3.1, the contribution with integrand on (−∞, 0) is bounded by

.

Z

0

−∞

e

txxx

v(t

)dt

L2x

. kvk

L1

tL2x

. For the last term, we reduce by Minkowski to show that

X

L

[L

1/2

kQ

L

t>t

e

−(t−t)∂xxx

v(t

))k

L2

tx

]

2

1/2

. kv(t

)k

L2x

. This can be proved by a time-restriction argument. Indeed, for any T > 0, we have

X

L

[L

1/2

kQ

L

T

(t)χ

t>t

e

−(t−t)∂xxx

v(t

))k

L2

]

2

1/2

. k|τ |

1/2

b v(t

)F

t

T

(t)χ

t>t

)(τ )k

L2

. kv(t

)k

L2

k|τ |

1/2

F

t

(η(t)χ

tT >t

)k

L2

. kv(t

)k

L2

.

We conclude by passing to the limit T → ∞.

Now we state a general and classical result which ensures that our resolu- tion space is well compatible with dispersive properties of the Airy equation.

Actually, it is a direct consequence of Lemma 4.1 in [19] together with the

fact that the resolution space S

0

used by Tao to solve 4-KdV contains our

space S

0

thanks to estimate (3.5)

(13)

Lemma 3.3 (Extension lemma). Let Z be a Banach space of functions on R × R with the property that

kg(t)u(t, x)k

Z

. kgk

Lt

ku(t, x)k

Z

holds for any u ∈ Z and g ∈ L

t

( R ). Let T be a spacial linear operator for which one has the estimate

kT (e

−t∂xxx

P

N

φ)k

Z

. kP

N

φk

L2

for some dyadic N and for all φ. Then one has the embedding kT (P

N

u)k

Z

. kP

N

uk

S0

.

Combined with the unitary of the Airy group in L

2

and the sharp Kato smoothing effect

k∂

x

e

−t∂xxx

φk

LxL2t

. kφk

L2

, ∀φ ∈ L

2

, (3.11) we deduce the following result.

Corollary 3.1. For any u, we have

1

kuk

L

t H−1x

. kuk

S−1

, (3.12) kP

N

uk

Lx L2

t

. N

−1

kP

N

uk

S0

, (3.13) provided the right-hand side is finite. In particular, S

−1

֒ → L

t

H

−1

.

4 Linear estimates

In this section we prove linear estimates related to the operator W as well as to the extension of the Duhamel operator introduced in (3.4).

At this this stage let us give some explanations on our choice of this extension. Let us keep in mind that this extension has to be compatible with linear estimates in both norms X

s,1/2,1

and Y

s,1/2

. First, since X

s,1/2,1

is a Besov in time space we are not allowed to simply multiply the Duhamel term by χ

R+

(t). Second, in order to prove the desired linear estimate in Y

s,1/2

the strategy is to use that the Duhamel term satisfies a forced KdV-Burgers equation. Unfortunately, it turns out that the extension introduced in [16],

1Note that (3.12) can also be deduced from estimate (3.9).

(14)

that makes the calculus simple, does not satisfy such PDE for negative time. The new extension that we introduce in this work has the properties to satisfy some forced PDE related to KdV-Burgers for negative times (see (4.14)) and to be compatible with linear estimates in X

s,1/2,1

. However the proof is now a little more complicated even if it follows the same lines than the one of Propositions 2.3 in [16], see also Proposition 4.4, [12].

The following lemma is a dyadic version of Proposition 2.1 in [16].

Proposition 4.1. For all φ ∈ H

−1

( R ), we have

kη(t)W (t)φk

S−1

. kφk

H−1

. (4.1) Proof. We bound the left-hand side in (4.1) by the X

−1,12,1

-norm of η(t)W (t)φ.

After square-summing in N , we may reduce to prove X

L

hL + N

2

i

1/2

kP

N

Q

L

(η(t)W (t)φ)k

L2

xt

. kP

N

φk

L2

(4.2) for each dyadic N . Using Plancherel, we obtain

X

L

hL + N

2

i

1/2

kP

N

Q

L

(η(t)W (t)φ)k

L2

xt

. X

L

hL + N

2

i

1/2

N

(ξ)ϕ

L

(τ ) φ(ξ)F b

t

(η(t)e

−|t|ξ2

)(τ )k

L2

τ ξ

. kP

N

φk

L2

X

L

hL + N

2

i

1/2

N

(ξ)P

L

(η(t)e

−|t|ξ2

)k

L

ξ L2t

. Hence it remains to show that

X

L

hL + N

2

i

1/2

N

(ξ)P

L

(η(t)e

−|t|ξ2

)k

L

ξ L2t

. 1. (4.3) We split the summand into L ≤ hN i

2

and L ≥ hN i

2

. In the former case, we get by Bernstein

X

L≤hNi2

hL + N

2

i

1/2

N

(ξ)P

L

(η(t)e

−|t|ξ2

)k

L

ξ L2t

. X

L≤hNi2

hN iL

1/2

sup

|ξ|∼N

kη(t)e

−|t|ξ2

k

L1

t

Also, one can bound kη(t)e

−|t|ξ2

k

L1

either by kηk

L1

or by ke

−|t|ξ2

k

L1

t

∼ |ξ|

−2

. It follows that

X

L≤hNi2

hL + N

2

i

1/2

N

(ξ)P

L

(η(t)e

−|t|ξ2

)k

L

ξ L2t

. hN i

2

min(1, N

−2

) . 1.

(15)

Now we deal with the case L ≥ hN i

2

. A standard paraproduct rearrange- ment allows us to write

P

L

(η(t)e

−|t|ξ2

) = P

L

X

M&L

(P

M

η(t)P

.M

e

−|t|ξ2

+ P

.M

η(t)P

M

e

−|t|ξ2

= P

L

(I) + P

L

(II).

Using the Schur’s test, the term P

L

(I ) is directly bounded by X

L≥hNi2

hL + N

2

i

1/2

N

P

L

(I)k

L

ξ L2t

. X

L

L

1/2

X

M&L

N

P

M

η(t)k

L

ξ L2t

N

P

.M

e

−|t|ξ2

k

Lξt

. X

M

M

1/2

kP

M

ηk

L2

t

. 1.

Similarly for P

L

(II), we have X

L≥hNi2

hL + N

2

i

1/2

N

P

L

(II)k

L

ξ L2t

. X

L

L

1/2

X

M&L

N

P

.M

η(t)k

Lξt

N

P

M

e

−|t|ξ2

k

L

ξ L2t

. X

M

M

1/2

N

P

M

e

−|t|ξ2

k

L

ξ L2t

.

Moreover, it is not too hard to check that if |ξ| ∼ N , then kP

M

e

−|t|ξ2

k

L2

t

.

kP

M

e

−|t|N2

k

L2

t

, thus X

L≥hNi2

hL + N

2

i

1/2

N

P

L

(II)k

L

ξ L2t

. X

M

M

1/2

kP

M

e

−|t|N2

k

L2t

. 1, where we used the fact that the Besov space ˙ B

2,11/2

has a scaling invariance and e

−|t|

∈ B ˙

2,11/2

.

Lemma 4.1. For w ∈ S ( R

2

), consider k

ξ

defined on R by k

ξ

(t) = η(t)ϕ

N

(ξ)

Z

R

e

itτ

e

(t−|t|)ξ2

− e

−|t|ξ2

iτ + ξ

2

w(τ e )dτ.

Then, for all ξ ∈ R , it holds X

L

hL + N

2

i

1/2

kP

L

k

ξ

k

L2t

. X

L

hL + N

2

i

−1/2

L

(τ )ϕ

N

(ξ) wk e

L2τ

.

(16)

Proof. Following [16], we rewrite k

ξ

as k

ξ

(t) = η(t)e

(t−|t|)ξ2

Z

|τ|≤1

e

itτ

− 1

iτ + ξ

2

g w

N

(τ )dτ + η(t) Z

|τ|≤1

e

(t−|t|)ξ2

− e

−|t|ξ2

iτ + ξ

2

w g

N

(τ )dτ + η(t)e

(t−|t|)ξ2

Z

|τ|≥1

e

itτ

iτ + ξ

2

g w

N

(τ )dτ − η(t) Z

|τ|≥1

e

−|t|ξ2

iτ + ξ

2

g w

N

(τ )dτ

= I + II + III − IV

where w

N

is defined by F

x

(w

N

)(ξ) = ϕ

N

(ξ)F

x

(w)(ξ).

Contribution of IV . Clearly we have kP

L

(IV )k

L2

t

. kP

L

(η(t)e

−|t|ξ2

)k

L2

t

Z

|τ|≥1

| g w

N

(τ )|

hiτ + ξ

2

i dτ.

On the other hand, by Cauchy-Schwarz in τ , Z

|τ|≥1

| g w

N

(τ )|

hiτ + ξ

2

i dτ . X

L

hL+N

2

i

−1

L

g w

N

k

L1τ

. X

L

hL+N

2

i

−1/2

L

g w

N

k

L2τ

,

which combined with (4.3) yields the desired bound.

Contribution of II. By Cauchy-Schwarz inequality, kP

L

(II)k

L2

t

. kP

L

(η(t)(e

(t−|t|)ξ2

− e

−|t|ξ|2

))k

L2

t

×

Z | g w

N

(τ )|

2

hiτ + ξ

2

i dτ

1/2

Z

|τ|≤1

hiτ + ξ

2

i

|iτ + ξ

2

|

2

!

1/2

. kP

L

(η(t)(e

(t−|t|)ξ2

− e

−|t|ξ|2

))k

L2

t

× N

−2

hN i X

L

hL + N

2

i

−1/2

L

g w

N

k

L2τ

. (4.4) Hence we need to estimate

X

L

hL + N

2

i

1/2

kP

L

(η(t)(e

(t−|t|)ξ2

− e

−|t|ξ|2

))k

L2

t

. X

L

hL + N

2

i

1/2

(kP

L

(η(t)e

(t−|t|)ξ2

)k

L2

t

+ kP

L

(η(t)e

−|t|ξ2

)k

L2

t

).

The second term in the right-hand side is bounded by 1 thanks to esti-

mate(4.3). Denote θ(t) = η(t)e

(t−|t|)ξ2

. It is not too hard to check that one

(17)

integration by parts yields | θ(τ ˆ )| .

|τ|1

whereas two integrations by parts give us | θ(τ ˆ )| .

hξi|τ|22

. We thus infer that

X

L

hL + N

2

i

1/2

L

θk ˆ

L2τ

. X

L≤1

hN iL

1/2

kθk

L1t

+ X

1≤L≤hNi2

hN i

L

1/2

+ X

L≥hNi2

hLi

1/2

hN i

2

L

3/2

. hN i. (4.5) This provides the result for N ≥ 1. In the case N ≤ 1, we use a Taylor expansion and obtain

kP

L

(η(t)(e

(t−|t|)ξ2

− 1 + 1 − e

−|t|ξ2

)k

L2t

. X

n≥1

|ξ|

2n

n!

kP

L

(|t|

n

η(t))k

L2

t

+ 2

n

kP

L

(t

n

η(t)χ

R

(t))k

L2

t

.

According to the Sobolev embedding H

1

֒ → B

2,11/2

as well as the estimate kχ

R

f k

H1

. kf k

H1

provided f (0) = 0, we deduce

X

L

hL + N

2

i

1/2

kP

L

(η(t)(e

(t−|t|)ξ2

− e

−|t|ξ2

))k

L2

t

. ξ

2

X

n≥1

1

n! (k|t|

n

η(t)k

B1/2

2,1

+ 2

n

kt

n

η(t)χ

R

(t)k

B1/2 2,1

) . N

2

X

n≥1

2

n

n! k|t|

n

η(t)k

Ht1

. N

2

. Gathering this and (4.4) we conclude that

X

L

hL + N

2

i

1/2

kP

L

(II)k

L2t

. X

L

hL + N

2

i

−1/2

L

g w

N

k

L2τ

.

Contribution of I. Since I can be rewritten as I = η(t)e

(t−|t|)ξ2

Z

|τ|≤1

X

n≥1

(itτ )

n

n!

g w

N

(τ ) iτ + ξ

2

dτ, we have

kP

L

(I)k

L2

t

. X

n≥1

1

n! kP

L

(t

n

θ(t))k

L2

t

Z

|τ|≤1

|τ |

n

|iτ + ξ

2

| | g w

N

(τ )|dτ.

(18)

Using Cauchy-Schwarz we get, for n ≥ 1, Z

|τ|≤1

|τ |

n

|iτ + ξ

2

| | w g

N

(τ )dτ .

Z | g w

N

(τ )|

2

hiτ + ξ

2

i dτ

1/2

Z

|τ|≤1

|τ |

2

hiτ + ξ

2

i

|iτ + ξ

2

|

2

!

1/2

. hN i

−1

X

L

hL + N

2

i

−1/2

L

g w

N

k

L2τ

.

Thus we see that it suffices to show that (see above the contribution of II for the definition of θ)

X

L

hL + N

2

i

1/2

X

n≥1

1

n! kP

L

(t

n

θ(t))k

L2t

. hN i.

But again we have |F

t

(t

n

θ(t))| . 2

n

min(

|τ|1

,

hξiτ22

) and arguing as in (4.5), we get

X

L

hL + N

2

i

1/2

X

n≥1

1

n! kP

L

(t

n

θ(t))k

L2t

. X

n≥1

hN i 2

n

n! . hN i.

Contribution of III. Setting ˆ g(τ ) =

gwN(τ)2

χ

|τ|≥1

, we have to prove X

L

hL + N

2

i

1/2

kP

L

(θg)k

L2

t

. X

L

hL + N

2

i

1/2

kP

L

gk

L2

t

. (4.6) Using the paraproduct decomposition, we have

P

L

(θg) = P

L

X

M&L

(P

.M

θP

∼M

g + P

∼M

θP

.M

g)

= P

L

(III

1

) + P

L

(III

2

)

and we estimate the contributions of these two terms separately.

Contribution of III

1

. The sum over L ≥ hN i

2

is estimated in the following way:

X

L≥hNi2

hL + N

2

i

1/2

kP

L

(III

1

)k

L2

t

. X

L≥hNi2

hLi

1/2

X

M&L

kP

.M

θk

Lt

kP

M

gk

L2

t

. X

M

hM i

1/2

kP

M

gk

L2

t

.

(19)

Now we deal with the case where L . hN i

2

. If ˆ θ is localized in an annulus {|τ | ∼ M }, we get from Bernstein inequality that

X

L≤hNi2

hL + N

2

i

1/2

X

M&L

kP

L

(P

M

θP

M

g)k

L2

t

. X

M

hN i X

L.M

L

1/2

kP

M

θP

M

gk

L1

t

. X

M

hN iM

1/2

kP

M

θk

L2

t

kP

M

gk

L2

. X

M

hN ikP

M

gk

L2

t

, (4.7)

where we used the estimate kP

M

θk

L2t

. k

ϕMτ(τ)

k

L2τ

. M

−1/2

. If ˆ θ is localized in a ball {|τ | ≪ M }, then we must have M ∼ L and thus

X

L≤hNi2

hL + N

2

i

1/2

X

M∼L

kP

L

(P

≪M

θP

M

g)k

L2t

. X

L

hN ikP

≪L

θk

Lt

kP

L

gk

L2t

,

which is acceptable.

Contribution of III

2

. Consider the case L ≥ hN i

2

. Since | θ| ˆ .

hξiτ22

, we have kP

L

(P

M

θP

.M

g)k

L2

t

. kϕ

M

θk ˆ

L1τ

kP

.M

gk

L2

t

. hN i

2

M kgk

L2

t

. It follows that

X

L≥hN2i

hL + N

2

i

1/2

kP

L

(III

2

)k

L2

t

. X

M&hNi2

M

1/2

hN i

2

M kgk

L2

t

. hN ikgk

L2

t

. It remains to establish the bound in the case L ≤ hN i

2

. We may assume that ˆ g is supported in a ball {|τ | ≪ M} since the other case has already been treated (cf. estimate (4.7)). Therefore, M ∼ L and

X

L≤hNi2

hL + N

2

i

1/2

X

M&L

kP

L

(P

M

θP

≪M

g)k

L2

t

. X

L

hN ikP

L

θP

≪L

gk

L2

t

. X

L

hN ikP

L

θk

L2

t

X

M≪L

kP

M

gk

Lt

. X

L

hN iL

−1/2

X

M≪L

M

1/2

kP

M

gk

L2

t

. X

M

hN ikP

M

gk

L2

t

.

The proof of Lemma 4.1 is complete.

(20)

Proposition 4.2. Let L : f → Lf denote the linear operator Lf (t, x) = η(t)

χ

R+

(t) Z

t

0

W (t − t

, t − t

)f (t

)dt

R

(t)

Z

t

0

W (t − t

, t + t

)f (t

)dt

. (4.8)

If f ∈ N

−1

, then

kLf k

S−1

. kf k

N−1

. (4.9) Proof. It suffices to show that

kLf k

X−1,12,1

. kf k

X−1,12,−1

(4.10) and

kLfk

Y−1,12

. kf k

Y−1,12

. (4.11)

Taking the x-Fourier transform, we get Lf (t, x) = U (t)

"

χ

R+

(t) η(t) Z

R

e

ixξ

Z

t

0

e

−|t−t2

F

x

(U (−t

)f (t

))(ξ) dt

dξ + χ

R−

(t) η(t)

Z

R

e

ixξ

Z

t

0

e

−|t+t2

F

x

(U (−t

)f(t

))(ξ) dt

#

= U (t)

"

η(t) Z

R

e

ixξ

Z

t

0

e

−|t|ξ2

e

tξ2

F

x

(U (−t

)f (t

))(ξ) dt

# . Setting w(t

) = U (−t

)f (t

), and using the time Fourier transform, we infer that

Lf(t, x) = U (t)

"

η(t) Z

R2

e

ixξ

e

itτ

e

(t−|t|)ξ2

− e

−|t|ξ2

iτ + ξ

2

w(τ, ξ)dτ dξ ˜

# . Estimate (4.10) follows then easily from Lemma 4.1.

Now we turn to estimate (4.11). After square summing, it suffices to prove that for any dyadic N ,

k(∂

t

+ ∂

xxx

− ∂

xx

+ I )P

N

Lf k

L1

tL2x

. kP

N

f k

L1

tL2x

. (4.12) In view of the expression of L it suffices to prove (4.12) separately for χ

R+

Lf and χ

R

Lf First, a straightforward calculation leads to

(∂

t

+ ∂

xxx

− ∂

xx

+ I)(χ

R+

Lf (t))

= η(t)χ

R+

(t)f(t) + (η

(t) + η(t))χ

R+

(t) Z

t

0

W (t − t

, t − t

)f (t

)dt

.

Références

Documents relatifs

Sanz-Serna and Christi di- vis d a modified Petrov-Galerkin finite elem nt method, and compared their method to the leading methods that had been developed to date:

Well-posedness results for the generalized Benjamin-Ono equation with arbitrary large initial data. Well-posedness results for the generalized Benjamin-Ono equation with small

[1] Gregory R. Baker, Xiao Li, and Anne C. Analytic structure of two 1D-transport equations with nonlocal fluxes. Bass, Zhen-Qing Chen, and Moritz Kassmann. Non-local Dirichlet

Then to cope with the loss of regularity of the perturbation with respect to the background state due to the degeneracy of the equation, we apply the Nash-Moser-H¨ ormander iter-

By introducing a Bourgain space associated to the usual KPI equation (re- lated only to the dispersive part of the linear symbol in the KPBI equation), Molinet-Ribaud [14] had

In Section 4 we show some L 2 -bilinear estimates which are used to prove the main short time bilinear estimates in Section 5 as well as the energy estimates in Section 6.. Theorem

In the three-dimensional case, as far as we know, the only available result concerning the local well-posedness of (ZK ) in the usual Sobolev spaces goes back to Linares and Saut

In view of the result of Kappeler and Topalov for KdV it thus appears that even if the dissipation part of the KdV-Burgers equation allows to lower the C ∞ critical index with