• Aucun résultat trouvé

GLOBAL PERSISTENCE OF GEOMETRICAL STRUCTURES FOR THE BOUSSINESQ EQUATION WITH NO DIFFUSION

N/A
N/A
Protected

Academic year: 2021

Partager "GLOBAL PERSISTENCE OF GEOMETRICAL STRUCTURES FOR THE BOUSSINESQ EQUATION WITH NO DIFFUSION"

Copied!
25
0
0

Texte intégral

(1)

HAL Id: hal-01290221

https://hal.archives-ouvertes.fr/hal-01290221

Preprint submitted on 17 Mar 2016

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.

GLOBAL PERSISTENCE OF GEOMETRICAL STRUCTURES FOR THE BOUSSINESQ EQUATION

WITH NO DIFFUSION

Raphaël Danchin, Xin Zhang

To cite this version:

Raphaël Danchin, Xin Zhang. GLOBAL PERSISTENCE OF GEOMETRICAL STRUCTURES FOR

THE BOUSSINESQ EQUATION WITH NO DIFFUSION. 2016. �hal-01290221�

(2)

BOUSSINESQ EQUATION WITH NO DIFFUSION

RAPHA¨EL DANCHIN AND XIN ZHANG

Abstract. Here we investigate the so-calledtemperature patch problemfor the incompressible Boussinesq system with partial viscosity, in the whole spaceRN(N ≥2),where the initial temperature is the charac- teristic function of some simply connected domain withC1,εH¨older regularity. Although recent results in [1, 15] ensure that an initiallyC1patch persists through the evolution, whether higher regularity is preserved has remained an open question. In the present paper, we give a positive answer to that issue globally in time, in the 2-D case for large initial data and in the higher dimension case for small initial data.

1. Introduction

This paper is devoted to the temperature patch problem for the following incompressible Boussinesq system with partial viscosity:

(B

ν,N

)

 

 

 

t

θ + u · ∇θ = 0,

t

u + u · ∇u − ν∆u + ∇Π = θe

N

, div u = 0,

(θ, u)|

t=0

= (θ

0

, u

0

).

Above, e

N

= (0, · · · , 0, 1) stands for the unit vertical vector in R

N

with N ≥ 2. The unknowns are the scalar function θ (the temperature), the velocity field u = (u

1

, u

2

, ..., u

N

) and the pressure Π, depending on the time variable t ≥ 0 and on the space variable x ∈ R

N

. We assume the viscosity ν to be a positive constant.

The above Boussinesq system is a toy model for describing the convection phenomenon in viscous incom- pressible flows, and arises in simplified models for geophysics (see e.g. [23]). A number of works are dedicated to the global well-posedness issue of (B

ν,2

) (see e.g. [1, 4, 15, 19, 21]). In particular, R. Danchin and M. Paicu proved in [15] (see also [18]) that (B

ν,2

) has a unique global solution (θ, u) such that θ ∈ C R

+

; L

2

(R

2

)

and

1

(1.1) u ∈

C R

+

; L

2

( R

2

)

∩ L

2loc

R

+

; H

1

( R

2

)

∩ L e

1loc

R

+

; H

2

( R

2

)

2

, whenever the initial data (θ

0

, u

0

) are in L

2

(R

2

)

3

and satisfy div u

0

= 0. Additionally, the following energy equality is fulfilled for all t ≥ 0:

ku(t)k

2L2(R2)

+ 2ν Z

t

0

k∇uk

2L2(R2)

dt

= ku

0

k

2L2(R2)

+ 2 Z

t

0

Z

R2

θu

2

dx dt

.

Global well-posedness results are also available in dimension N ≥ 3, but exactly as for the standard Navier- Stokes equations, the initial data have to satisfy a suitable smallness condition, see for instance [14, 15].

To better explain the main motivation of our work, which is the temperature patch problem, let us assume that N = 2 for a while. Based on the aforementioned well-posedness result, one may consider for θ

0

the characteristic function of some simply connected bounded domain D

0

of the plane. Given that θ is just advected by the velocity field u, we expect to have θ(t, x) = 1

Dt

(x) for all t ≥ 0, where D

t

:= ψ

u

(t, D

0

) and ψ

u

stands for the flow associated to u, that is to say the solution to the following (integrated) ordinary differential equation:

(1.2) ψ

u

(t, x) := x +

Z

t 0

u(t

, ψ

u

(t

, x)) dt

.

Key words and phrases. Boussinesq equations; Incompressible flows; Striated regularity, Para-vector field; H¨older spaces;

Temperature patch problem.

1The notationLe1loc(R+;H2) designates a (close) superspace ofL1loc(R+;H2), see (2.2).

1

(3)

If the regularity of u is given by (1.1), then it has been proved by J.-Y. Chemin and N. Lerner in [9] that (1.2) has a unique solution, which is in C( R

+

; C

0,1−η

)

2

for any η > 0. Now, if we add a bit more regularity on the initial data, for instance

2

u

0

∈ (B

02,1

)

2

and θ

0

∈ B

2,10

then, according to [1], all the entries of ∇u are in L

1loc

(R

+

; C

b

), and the flow ψ

u

(t, ·) is thus C

1

. Consequently, the C

1

regularity of the temperature patch is preserved for all time.

Then a natural question arises: what if we start with a C

1,ε

H¨ older domain D

0

with ε ∈]0, 1[ ? Our concern has some similarity with the celebrated vortex patch problem for the 2-D incompressible Euler equations. In that case, it has been proved (see e.g. [6, 7] and the references therein) that the C

1,ε

regularity of the patch of the vorticity persists for all time. Proving that in our framework, too, the H¨ older regularity of D

t

is conserved is the main purpose of the present paper. Just like for the vortex patch problem for Euler equations, our result will come up as a consequence of a much more general property of global-in-time persistence of striated regularity, a definition that originates from the work of J.-Y. Chemin in [5].

Before stating our main results, we need to introduce some notation, and to clarify what striated regularity is. Assume that

3

X = X

k

(x)∂

k

is some vector field acting on functions in C

1

(R

N

; R). As usual, vector fields are identified with vector valued functions from R

N

to R

N

, and ∂

X

f stands for the directional derivative of f ∈ C

1

(R

N

; R) along the vector field X, namely

X

f := X

k

k

f = X · ∇f.

The evolution X

t

(x) := X (t, x) of any continuous initial vector field X

0

along the flow of u is defined by:

X (t, x) := (∂

X0

ψ

u

) ψ

u−1

(t, x) .

In the C

1

case, combining the chain rule and the definition of the flow in (1.2) implies that X satisfies the transport equation

4

(1.3)

t

X + u · ∇X = ∂

X

u, X|

t=0

= X

0

.

Applying operator div to (1.3) and remembering that div u = 0, we obtain in addition (1.4)

t

div X + u · ∇div X = 0, div X|

t=0

= div X

0

.

This implies that the divergence-free property is conserved through the evolution.

As we will see in Section 5, the temperature patch problem is closely related to the conservation of H¨ older regularity C

0,ε

for X. According to the classical theory of transport equations, if u is Lipschitz with respect to the space variable (a condition that will be ensured if the data of (B

ν,N

) have critical Besov regularity), then conservation of C

0,ε

regularity for X is equivalent to the fact that all the components of ∂

X

u have the regularity C

0,ε

with respect to the space variable.

In the 2-D case, it is natural to recast the regularity of u along the vector field X in terms of the vorticity ω := ∂

1

u

2

− ∂

2

u

1

as the simple transport-diffusion equation is fulfilled:

(1.5)

t

ω + u · ∇ω − ν∆ω = ∂

1

θ, ω|

t=0

= ω

0

,

and as it is known (see e.g. [2], Chap. 7) that

5

:

(1.6) k∂

X

uk

Cε

. k∇uk

L

kX k

Cε

+ kdiv (Xω)k

Cε−1

, where for any real number s, we denote

6

C

s

≡ C

s

(R

N

) := B

∞,∞s

(R

N

).

2See the definition of Besov spaces in the next section.

3We adopt Einstein summation convention in the whole text: summation is taken with respect to the repeated indices, whenever they occur both as a subscript and a superscript.

4Omitting the indextinXtfor notational simplicity.

5In all the paper, we agree thatA.BmeansA≤CB for some harmless constantC.

6Recall thatCk+εcoincides with the standard H¨older spaceCk,εwheneverk∈Nandε∈]0,1[.

(4)

Now, applying operators ∂

X

and div (X·) to the temperature and vorticity equations, respectively, we get the following system for ∂

X

θ, div (Xω)big):

(1.7)

( ∂

t

X

θ + u · ∇∂

X

θ = 0,

t

div (Xω) + u · ∇div (Xω) − ν∆div (Xω) = f, with f := νdiv X ∆ω − ∆(Xω)

+ div (X∂

1

θ).

Let us recap. Roughly speaking, to propagate the C

ε

regularity of X, it is sufficient to control ∂

X

u in L

1loc

( R

+

; C

ε

)

2

, and this may be achieved, thanks to (1.6), if bounding the distribution div (Xω) in L

1loc

( R

+

; C

ε−1

). Then by taking advantage of smoothing properties of the heat flow, this latter informa- tion may be obtained through a bound of f in the very negative space L

1loc

(R

+

; C

ε−3

), if assuming that div (X

0

ω

0

) ∈ C

ε−3

. Staring at the expression of f, we thus need to bound ∂

X

θ in L

loc

(R

+

; C

ε−2

). As no gain of regularity may be expected from the transport equation satisfied by ∂

X

θ, we have to assume initially that

X0

θ

0

∈ C

ε−2

. This motivates the following statement which is our main result of propagation of striated regularity in the 2-D case

7

.

Theorem 1.1. Suppose that (ε, q) ∈]0, 1[×]1,

2−ε2

[. Let θ

0

be in B

2 q−1

q,1

(R

2

) and u

0

be a divergence-free vector field in L

2

(R

2

)

2

, with vorticity ω

0

:= ∂

1

u

20

− ∂

2

u

10

in B

2 q−2

q,1

(R

2

). Then there exists a unique global solution (θ, u) of System (B

ν,2

), such that

(1.8) (θ, u, ω) ∈ C(R

+

; B

2 q−1

q,1

) × C(R

+

; L

2

)

2

× C(R

+

; B

2 q−2

q,1

) ∩ L

1loc

(R

+

; B

2 q

q,1

) . Furthermore, if we consider some X

0

in C

ε

(R

2

)

2

satisfying ∂

X0

θ

0

∈ C

ε−2

(R

2

) and div (X

0

ω

0

) ∈ C

ε−3

(R

2

), then there exists a unique global solution

8

X ∈ C

w

(R

+

; C

ε

) to (1.3) and we have

X

θ, div (Xω)

∈ C

w

(R

+

; C

ε−2

) × C

w

(R

+

; C

ε−3

) ∩ L e

1loc

(R

+

; C

ε−1

) .

Additionally, there is a constant C

0,ν

depending only on the initial data and viscosity constant such that for any t ≥ 0,

kX k

Lt (Cε)

≤ C

0,ν

exp

exp exp(C

0,ν

t

4

) . A few comments are in order:

• The functional space B

2 q−1

q,1

( R

2

) for θ

0

is large enough to contain the characteristic function of any bounded C

1

domain. This will be needed to investigate the temperature patch problem later (see Corollary 1.2 and Section 5 below).

• It may be seen by means of elementary paradifferential calculus that if (ε, q) ∈]0, 1[×]1,

2−ε2

[ then div (X

0

ω

0

) and ∂

X0

θ

0

are distributions of C

−3

and C

−2

, respectively, and that the following (sharp) estimates are fulfilled:

kdiv (X

0

ω

0

)k

C−3

. kX

0

k

Cε

0

k

B

2 q−2 q,1

and k∂

X0

θ

0

k

C−2

. kX

0

k

Cε

0

k

B

2 q−1 q,1

.

Therefore, our striated regularity assumption on the initial data is indeed additional information.

• The required level of regularity is much lower than in the (inviscid) 2-D vortex patch problem where div (X

0

ω

0

) ∈ C

ε−1

(R

2

) is needed. This is because the smoothing effect given by the heat flow enables us to gain two derivatives with respect to the initial data. Another difference is that to tackle the temperature patch problem, it is not necessary to consider a family of vector fields that does not degenerate on the whole R

2

: just one suitably chosen vector-field that does not vanish in the neighborhood of the boundary of the patch is enough, as we shall see just below.

7To fully benefit from the smoothing properties of the heat flow in the endpoint case, one has to work in (close) superspaces ofL1loc(R+;Cs) denoted byLe1loc(R+;Cs) and defined in (2.2).

8IfEis a Banach space with predualEthenCw(R+;E) stands for the set of measurable functionsh:R+→Esuch that for allφ∈E,the functiont7→ hh(t), φiE×E is continuous onR+.

(5)

Let us now go to the temperature patch problem in the 2-D case. More precisely, consider a C

1,ε

simply connected bounded domain D

0

of R

2

(in other words ∂D

0

is a C

1,ε

Jordan curve on R

2

). Let D

0

be any bounded domain of R

2

such that

9

(1.9) D

0

∩ D

0

= ∅.

Then the following result holds true.

Corollary 1.2. Let (M

1

, M

2

) be in R

2

. Assume that θ

0

= M

1

1

D0

and that the vorticity ω

0

of u

0

may be decomposed into

(1.10) ω

0

= M

2

1

D0

− ω e

0

for some ω e

0

∈ L

r

(R

2

) with r > 1, supported in D

0

and such that (1.11)

Z

R2

ω e

0

(x) dx = M

2

|D

0

|.

Then there exists a unique solution (θ, u) to System (B

ν,2

), satisfying the properties listed in Theorem 1.1.

Furthermore, we have θ(t, ·) = M

1

1

Dt

where D(t) := ψ

u

(t, D

0

) and ∂D(t) remains a C

1,ε

Jordan curve of R

2

for all t ≥ 0.

Let us make some comments on that corollary.

• Hypothesis (1.11) ensures that the initial vorticity is mean free, which is necessary to have u

0

∈ L

2

(R

2

)

2

. As a matter of fact, the more natural assumption ω

0

= M

2

1

D0

would require our extending Theorem 1.1 to infinite energy velocity fields, which introduces additional technicalities.

• Assumption (1.10) on the vorticity may seem somewhat artificial as it has no persistency whatsoever through the time evolution, even in the asymptotics ν → 0 (in contrast with the slightly viscous vortex patch problem, see [10]). This is just to have a concrete example of initial velocity for which one can give a positive answer to the temperature patch problem.

• Corollary 1.2 may be generalized to the case where the initial velocity u

0

∈ L

2

( R

2

)

2

is such that ω

0

∈ B

2 q−2

q,1

(R

2

) for some 1 < q <

2−ε2

and satisfies div (X

0

ω

0

) ∈ C

ε−3

(R

2

) for some vector field X

0

∈ C

ε

( R

2

) that does not vanish on ∂D

0

and is tangent to ∂D

0

. One just has to follow the proof that is proposed in Section 5 to get this more general result.

In space dimension N ≥ 3, the vorticity equation has an additional stretching term, and it is thus less natural to measure the striated regularity by means of div (X Ω) with Ω denoting the matrix of curl u (even though we suspect that our 2-D approach is adaptable to the high-dimensional case, like in [17]). We shall thus concentrate on the regularity of ∂

X

θ and ∂

X

u. An additional (related) difficulty is that one cannot expect to prove global existence for general large initial data, since (B

ν,N

) contains the standard incompressible Navier-Stokes equations as a particular case. Therefore we shall prescribe some smallness condition on the data (the same one as in [15]) to achieve a global statement. This leads to the following theorem:

Theorem 1.3. Suppose that N ≥ 3 and that (ε, p) ∈]0, 1[×]N,

1−εN

[. Assume that θ

0

is in B

0N,1

(R

N

) ∩ L

N3

(R

N

) and that the components of the divergence-free vector field u

0

are in B

N p−1

p,1

(R

N

) and in the weak Lebesgue space L

N,∞

(R

N

). If there exists a (small) positive constant c independent of p such that

ku

0

k

LN,∞

+ ν

−1

0

k

LN3

≤ cν, then Boussinesq system (B

ν,N

) has a unique global solution

(θ, u, ∇Π) ∈ C( R

+

; B

0N,1

) × C( R

+

; B

N p−1

p,1

) ∩ L

1loc

( R

+

; B

N p+1 p,1

)

N

× L

1loc

( R

+

; B

N p−1 p,1

)

N

. Moreover, suppose that the vector field X

0

is in the space C e

ε

(R

N

)

N

defined by e

C

ε

(R

N

)

N

:= {Y ∈ C

ε

(R

N

)

N

: div Y ∈ C

ε

(R

N

)},

9We denote byAthe closure of the subsetAinRN, N≥2.

(6)

and that the components of (∂

X0

θ

0

, ∂

X0

u

0

) are in C

ε−2

( R

N

). Then the System (1.3) has a unique solution X ∈ C

w

( R

+

; C e

ε

), that satisfies for all t ≥ 0,

kX k

L

t (Ceε)

≤ C

0,ν

exp exp(C

0,ν

t) with some constant C

0,ν

depending only on the initial data and on ν.

Furthermore, the triplet (∂

X

θ, ∂

X

u, ∂

X

∇Π) belongs to

C

w

(R

+

; C

ε−2

) × C

w

(R

+

; C

ε−2

) ∩ L e

1loc

(R

+

; C

ε

)

N

× L e

1loc

(R

+

; C

ε−2

)

N

.

As in the 2-D case, the above result will enable us to solve the temperature patch problem. Before giving the exact statement, let us recall what a C

1,ε

domain is in dimension N ≥ 2.

Definition. A simply connected bounded domain D ⊂ R

N

is of class C

1,ε

if its boundary ∂D is some compact hypersurface of class C

1,ε

.

Fix some domain D

0

of class C

1,ε

and further consider another C

1

simple bounded domain J

0

such that D

0

⊂ J

0

. Then we have the following statement

10

:

Corollary 1.4. Let N ≥ 3 and (m

1

, m

2

) be a pair of sufficiently small constants. Assume that θ

0

= m

1

1

D0

and that the initial vorticity Ω

0

:= curl u

0

, i.e. (Ω

0

)

ij

:= ∂

j

u

i0

− ∂

i

u

j0

for any i, j = 1, ..., N, satisfies Ω

0

:= m

2

1

J0

A

0

where A

0

stands for the anti-symmetric matrix defined by ( A

0

)

ij

= 1 for i < j.

Then θ(t, ·) = m

1

1

Dt

where D(t) := ψ

u

(t, D

0

) and D(t) remains a simply connected domain of class C

1,ε

, for any t ≥ 0.

The rest of the paper unfolds as follows. In the next section, we shortly introduce Besov spaces and present some linear or nonlinear estimates, which will be needed to achieve our results. Then the proofs of main theorems for the propagation of striated regularity will be revealed in Section 3 (for 2-D case) and Section 4 (for N-D case). Section 5 is devoted to the temperature patch problems. Some technical commutator bounds are proved in the Appendix.

2. Basic notations and linear estimates

We here introduce definitions and notations that are used throughout the text, and recall some properties of Besov spaces and transport or transport-diffusion equations.

Let us begin with the definition of the nonhomogeneous Littlewood-Paley decomposition (for more details see [2], Chap. 2). Set B := {ξ ∈ R

N

: |ξ| ≤ 4/3} and C := {ξ ∈ R

N

: 3/4 ≤ |ξ| ≤ 8/3}. We fix two smooth radial functions χ and ϕ, supported in B and C, respectively, and such that

χ(ξ) + X

j≥0

ϕ(2

−j

ξ) = 1, ∀ ξ ∈ R

N

.

We then introduce the Fourier multipliers ∆

−1

:= χ(D) and ∆

j

:= ϕ(2

−j

D) with j ≥ 0 (the so-called nonhomogeneous dyadic blocks ) and the low frequency cut-off operator

S

j

:= X

j≤j−1

j

.

With those notations, the nonhomogeneous Besov space B

p,rs

(R

N

) may be defined by B

p,rs

(R

N

) := {u ∈ S

(R

N

) : kuk

Bsp,r

:= 2

js

k∆

j

uk

Lp

r(N∪{−1})

< ∞}, for (s, p, r) ∈ R × [1, ∞]

2

.

Let us next introduce the following paraproduct and remainder operators:

T

u

v := X

j≥−1

S

j−N0

u∆

j

v and R(u, v) ≡ X

j≥−1

j

u ∆ e

j

v := X

j≥−1

|j−k|≤N0

j

u∆

k

v,

where N

0

stands for some large enough (fixed) integer.

10Like in the 2-D case, it goes without saying that much more general initial velocities may be considered

(7)

The following decomposition, first introduced by J.-M. Bony in [3]:

(2.1) uv = T

u

v + T

v

u + R(u, v),

holds true whenever the product of the two tempered distribution u and v is defined. It will play a funda- mental role in our study.

Bilinear operators R and T possess continuity properties in a number of functional spaces (see e.g. Chap.

2 in [2]). We shall recall a few of them throughout the text, when needed.

When investigating evolutionary equations in Besov spaces and, in particular, parabolic type equations, it is natural to use the following tilde Besov spaces first introduced by J.-Y. Chemin in [8]: for any t ∈]0, ∞]

and (s, p, r, ρ) ∈ R × [1, ∞]

3

, we set L e

ρt

B

sp,r

(R

N

)

:= n

u ∈ S

(]0, t[×R

N

) : kuk

Leρ

t(Bp,rs )

:= 2

js

k∆

j

uk

Lρ(]0,t[;Lp)

r

< ∞ o

·

In the particular case where p = r = 2 (resp. p = r = ∞), B

p,rs

coincides with the Sobolev space H

s

(resp.

the generalized H¨ older space C

s

), and we shall alternately denote (2.2) L e

ρt

H

s

(R

N

)

:= L e

ρt

B

s2,2

(R

N

)

and L e

ρt

C

s

(R

N

)

:= L e

ρt

B

∞,∞s

(R

N

) .

Let us next state some a priori estimates for the transport and transport-diffusion equations in (nonhomo- geneous) Besov spaces.

Proposition 2.1. Assume that v is a divergence free vector field. Let (p, p

1

, r, ρ, ρ

1

) ∈ [1, ∞]

5

and s ∈ R satisfy

−1 − min N

p

1

, N p

< s < 1 + min N

p , N p

1

and ρ

1

≤ ρ.

Let f be a smooth solution of the following transport-diffusion equation with diffusion parameter ν ≥ 0:

(T D

ν

)

t

f + div (f v) − ν∆f = g, f |

t=0

= f

0

.

Then there exists a constant C depending on N, p, p

1

and s such that for all t ≥ 0, (2.3) ν

1ρ

kf k

Leρt(Bs+ 2p,rρ)

≤ Ce

C(1+νt)

1 ρVp1(t)

(1 + νt)

1ρ

kf

0

k

Bp,rs

+ (1 + νt)

1+1ρρ11

ν

ρ11−1

kgk

Leρt1(Bs−2+ 2p,r ρ1)

, where

V

p1

(t) :=

Z

t 0

k∇vk

B

N p1 p1,∞∩L

dt

. In the limit case s = −1 − min

Np1

,

Np

, one just need to refine k∇vk

B

N p1 p1,∞∩L

by k∇vk

B

N p1 p1,1

in the definition of V

p1

.

Proof. The above statement has been proved in [12] in the case p ≤ p

1

. The generalization to the case p > p

1

is straightforward: in fact that restriction came from the following commutator estimate

11

: 2

js

k[v · ∇, ∆

j

]f k

Lp

r

. k∇vk

B

N p1 p,1

kf k

Bp,rs

, that has been proved only in the case p ≤ p

1

.

To handle the case p > p

1

, we proceed exactly as in [12], decomposing R

j

:= [v·∇, ∆

j

]f into R

j

= P

6 i=1

R

ij

with

R

1j

:= [T

˜vk

, ∆

j

]∂

k

f, R

j2

:= T

kjf

˜ v

k

, R

3j

:= −∆

j

T

kf

˜ v

k

, R

j4

:= ∂

k

R(˜ v

k

, ∆

j

f ), R

5j

:= −∂

k

j

R(˜ v

k

, f), R

j6

:= [S

0

v

k

, ∆

j

]∂

k

f, where ˜ v := v − S

0

v.

11We here adopt the usual notation [A, B] for the commutatorAB−BA.

(8)

In [12], Condition p ≤ p

1

is used only when bounding R

3j

. Now, if p > p

1

and s < 1 +

Np

then combining standard continuity results for the paraproduct with the embedding B

p,rs−1

( R

N

) ֒ → B

s−1−

N

∞,r p

( R

N

) implies that 2

j(s+p1NNp)

kR

3j

k

Lp

r

. k∇vk

B

N p1 p1,∞

k∇f k

Bs−1

p,r

,

whence the desired inequality.

Remark 2.2. We shall often use the above proposition in the particular case ν = 0 and (p, r, ρ, ρ

1

) = (∞, ∞, ∞, 1). Then Inequality (2.3) reduces to

kf k

Lt (Cs)

≤ Ce

CVp1(t)

kf

0

k

Cs

+ kgk

Le1 t(Cs)

if − 1 − N p

1

< s < 1.

We finally recall a refinement of Vishik’s estimates for the transport equation [24] obtained by T. Hmidi and S. Keraani in [20], which is the key to the study of long-time behavior of the solution in critical spaces for Boussinesq system (B

ν,N

) (see [1, 15]).

Proposition 2.3. Assume that v is divergence-free and that f satisfies the transport equation (T D

0

). There exists a constant C such that for all (p, r) ∈ [1, ∞]

2

and t > 0,

kf k

Le

t (Bp,r0 )

≤ C kf

0

k

Bp,r0

+ kgk

Le1 t(Bp,r0 )

1 +

Z

t 0

k∇vk

L

·

3. Propagation of striated regularity in the 2D-case

This section is devoted to the proof of Theorem 1.1. To simplify the computations, we shall first make the change of unknowns

(3.1) e θ(t, x) = ν

2

θ(νt, x), e u(t, x) = νu(νt, x), P e (t, x) = ν

2

P (νt, x) so as to reduce the study to the case ν = 1.

Throughout this section, we always assume that

(3.2) 0 < ε < 1, q > 1 and ε

2 + 1 q > 1, in accordance with the hypotheses of Theorem 1.1.

3.1. A priori estimates for the Lipschitz norm of the velocity field. Those estimates will be based on the following global existence theorem (see [1]).

Theorem 3.1. Let u

0

be a divergence-free vector field belonging to the space L

2

(R

2

) ∩ B

∞,1−1

(R

2

)

2

and let θ

0

be in B

2,10

(R

2

). Then there exists a unique global solution (u, θ, ∇Π) for System (B

1,2

) such that

(3.3) u ∈ C(R

+

; L

2

∩ B

∞,1−1

) ∩ L

2loc

(R

+

; H

1

) ∩ L

1loc

(R

+

; B

∞,11

)

2

, θ ∈ C

b

(R

+

; B

2,10

) and ∇Π ∈ L

1loc

(R

+

; B

2,10

)

2

.

Moreover, for any t > 0, there exists a constant C

0

depending only on the initial data such that kuk

L1t(B∞,11 )

≤ C

0

e

C0t4

.

We claim that the data (θ

0

, u

0

) of Theorem 1.1 fulfill the assumptions of the above theorem. Indeed, decompose u

0

= ∆

−1

u

0

+ (Id − ∆

−1

)u

0

and apply the following Biot-Savart law

∇u = −∇(−∆)

−1

ω with ∇

:= (−∂

2

, ∂

1

).

Then thanks to the obvious embedding B

2 q−2

q,1

(R

2

) ֒ → B

−12,1

(R

2

), we obtain that ku

0

k

B−1

∞,1

. ku

0

k

B02,1

. k∆

−1

u

0

k

L2

+ k(Id − ∆

−1

)∇u

0

k

B−1

2,1

. ku

0

k

L2

+ kω

0

k

B

2 q−2 q,1

.

(3.4)

(9)

Besides, the obvious embedding θ

0

∈ B

2 q−1

q,1

(R

2

) ֒ → B

02,1

(R

2

) holds. Hence one may apply the above theorem to our data. The corresponding global solution (θ, u) fulfills (3.3) and the following inequality for all t ≥ 0, (3.5) k∇uk

L1t(L)

. kuk

L1t(B1∞,1)

≤ C

0

e

C0t4

.

3.2. A priori estimates for θ and ω. We now want to prove that (θ, u) fulfills the additional property (1.8). To this end, the first observation is that θ satisfies a free transport equation. Hence, from the standard theory of transport equations (apply Proposition 2.1 with ν = 0) and the bound (3.5), we deduce that

(3.6) kθk

Let (B

2 q−1

q,1 )

≤ e

Ck∇ukL1t(L)

0

k

B

2 q−1 q,1

≤ kθ

0

k

B

2 q−1 q,1

exp C

0

exp(C

0

t

4

) .

In order to bound ω, one may apply Proposition 2.1 to the vorticity equation, which yields for all t ≥ 0, kωk

Let (B

2 q−2 q,1 )

+ kωk

L1t(B

2 q

q,1)

≤ C(1 + t)e

C(1+t)k∇ukL1t(L)

0

k

B

2 q−2 q,1

+ k∇θk

L1t(B

2 q−2 q,1 )

·

Hence, taking advantage of (3.5) and (3.6), we get

(3.7) kωk

Lt (B

2 q−2 q,1 )

+ kωk

L1t(B

2 q

q,1)

≤ C exp(exp(C

0

t

4

)) kω

0

k

B

2 q−2 q,1

+ kθ

0

k

B

2 q−1 q,1

.

Then Biot-Savart law allows to improve the regularity of the velocity u as follows:

(3.8) U

q

(t) := k∇uk

L1t(B

2 q q,1)

. kωk

L1t(B

2 q q,1)

≤ C

0

exp exp(C

0

t

4

) , where C

0

depends only on the Lebesgue and Besov norms of the data in Theorem 1.1.

3.3. A priori estimates for the striated regularity. We shall need the following lemma which is a straightforward generalization of Inequality (1.6).

Lemma 3.2. For any ε ∈]0, 1[, there exists a constant C such that the following estimate holds true:

(3.9) k∂

X

uk

Le1

t(Cε)

. Z

t

0

k∇uk

L

kXk

Cε

dt

+ kdiv (Xω)k

Le1 t(Cε−1)

.

As we explained in the introduction, the above lemma implies that it suffices to bound div (Xω) in L e

1t

( C

ε−1

) for all t ≥ 0 in order to propagate the H¨ older regularity C

ε

of X. This will be based on Proposition 2.1, remembering that div (Xω) fulfills the transport-diffusion equation:

t

div (Xω) + u · ∇div (Xω) − ∆div (Xω) = f with

f = div F + div (X∂

1

θ) and F := X ∆ω − ∆(Xω).

Thanks to Proposition 2.1, we have

(3.10) kdiv (Xω)k

Lt (Cε−3)

+ kdiv (Xω)k

Le1

t(Cε−1)

. (1 + t)

kdiv (X

0

ω

0

)k

Cε−3

+ Z

t

0

U

q

kdiv (Xω)k

Cε−3

dt

+ kf k

Le1 t(Cε−3)

, where U

q

has been defined in (3.8).

Let us now bound the source term f in L e

1t

( C

ε−3

) of (3.10). It is not hard to estimate F via the following decomposition:

F = [T

X

, ∆]ω + T

∆ω

X + R(X, ∆ω) − ∆T

ω

X − ∆R(ω, X ).

Using the commutator estimates of Lemma A.1 and standard results of continuity for the paraproduct and remainder operators (see e.g. [2], Chap. 2), we get if Condition (3.2) is fulfilled,

(3.11) kF k

Cε−2

. kωk

B

2 q q,1

kXk

Cε

. The last part of f may be decomposed into

div (X∂

1

θ) = ∂

1

(div (Xθ)) − div (θ∂

1

X ).

(10)

By Bony’s decomposition (2.1),

θ∂

1

X = T

θ

1

X + T

1X

θ + R(θ, ∂

1

X ).

As Condition (3.2) is fulfilled, we thus have

(3.12) kθ∂

1

X k

Cε−2

. kθk

B

2 q−1 q,1

k∂

1

X k

Cε−1

.

Next, we see that

div (Xθ) = ∂

X

θ + θ div X.

The last term θ div X may be bounded as in (3.12). Therefore, combining with Inequalities (3.11) and (3.12), integrating with respect to time, and using also the obvious embedding L

1t

( C

ε−3

) ֒ → L e

1t

( C

ε−3

), we end up with

(3.13) kf k

Le1

t(Cε−3)

. Z

t

0

(kωk

B

2 q q,1

+ kθk

B

2 q−1 q,1

)kX k

Cε

dt

+ k∂

X

θk

L1

t(Cε−2)

.

Next, resuming to the transport equation (1.3) satisfied by X, combining with the last item of Proposition 2.1, and using also (3.9) and (3.10), we get

kXk

Lt (Cε)

≤ kX

0

k

Cε

+ C Z

t

0

k∇uk

L

kXk

Cε

dt

+ Ck∂

X

uk

Le1 t(Cε)

≤ kX

0

k

Cε

+ C Z

t

0

k∇uk

L

kXk

Cε

dt

+ Ckdiv (Xω)k

Le1 t(Cε−1)

≤ kX

0

k

Cε

+ C(1 + t)kdiv (X

0

ω

0

)k

Cε−3

+ C(1 + t) Z

t

0

U

q

kXk

Cε

+ kdiv (Xω)k

Cε−3

dt

+ kf k

Le1 t(Cε−3)

. (3.14)

Set

Z (t) := kX k

Lt (Cε)

+ kdiv (Xω)k

Lt (Cε−3)

and

W (t) := U

q

(t) + Z

t

0

(kωk

B

2 q q,1

+ kθk

B

2 q−1 q,1

) dt

. Observe that the bounds (3.6), (3.7) and (3.8) imply that

(3.15) W (t) ≤ C

0

exp exp(C

0

t

4

) kω

0

k

B

2 q−2 q,1

+ kθ

0

k

B

2 q−1 q,1

.

Then putting together (3.10), (3.13) and (3.14) yields Z (t) . (1 + t)

Z (0) + k∂

X

θk

L1

t(Cε−2)

+ Z

t

0

W

(t

)Z(t

) dt

· Hence, by virtue of Gronwall lemma,

(3.16) Z(t) ≤ C(1 + t)(Z(0) + k∂

X

θk

L1t(Cε−2)

)e

C(1+t)W(t)

.

Now, Proposition 2.1 and the fact that ∂

X

θ satisfies a free transport equation imply that (3.17) k∂

X

θk

Lt (Cε−2)

≤ e

CUq(t)

k∂

X0

θ

0

k

Cε−2

.

Adding (3.15) and (3.17) to (3.16), we thus obtain

(3.18) Z (t) ≤ C

0

exp

exp exp(C

0

t

4

)

· Resuming to (3.10), one can eventually conclude that

(3.19) kdiv (Xω)k

Le1

t(Cε−1)

+ kdiv (Xω)k

Lt (Cε−3)

≤ C

0

exp

exp exp(C

0

t

4

)

·

(11)

3.4. Completing the proof of Theorem 1.1. In order to make the previous computations rigorous, we have to work on smooth solutions. To this end, we solve System (B

1,2

) with smoothed out initial data

0n

, u

n0

) := (S

n

θ

0

, S

n

u

0

).

It is clear by embedding and (3.4) that the components of (θ

0

, u

0

) belong to all Sobolev spaces H

s

( R

2

).

Thanks to the result in [4], we thus get a unique global smooth solution (θ

n

, u

n

, ∇Π

n

) having Sobolev regularity of any order. Furthermore, as θ

0n

belongs to all spaces B

sq,1

and satisfies a linear transport equation with a smooth velocity field, we are guaranteed that θ

n

∈ C(R

+

; B

q,1s

) for all s ∈ R. Then as ω

n

is the solution of the transport-diffusion equation with ω

n0

∈ B

q,1s

(R

2

) and source term ∂

1

θ

n

in C R

+

; B

q,1s

(R

2

) for all s ∈ R, we conclude that (θ

n

, ω

n

) belongs to all sets E

qs

defined by

E

qs

:= n

(ϑ, σ) : ϑ ∈ L

loc

(R

+

; B

2 q−1+s

q,1

), σ ∈ L e

loc

(R

+

; B

2 q−2+s

q,1

) ∩ L

1loc

(R

+

; B

2 q+s q,1

) o

·

Finally, regularizing X

0

into X

0n

:= S

n

X

0

and setting X

n

(t, x) := (∂

X0n

ψ

un

)(ψ

u−1n

(t, x)), we see that X

n

belongs to H¨ older spaces of any order, and satisfies (1.3) with velocity field u

n

. The estimates that we proved so far are thus valid for (θ

n

, u

n

, X

n

). In particular, Inequalities (3.6), (3.15) and (3.18) are satisfied for all n ∈ N, with their r.h.s. depending on n through the (regularized) initial data respectively.

Deducing from (3.6) and (3.15) that the sequence (θ

n

, ω

n

)

n∈N

is bounded in E

q0

is obvious as S

n

maps L

p

to itself for all n ∈ N, with a norm independent of n. This guarantees that

(3.20) kθ

n0

k

B

2 q−1 q,1

. kθ

0

k

B

2 q−1 q,1

, ku

n0

k

L2

. ku

0

k

L2

and kω

0n

k

B

2 q−2 q,1

. kω

0

k

B

2 q−2 q,1

. To justify the uniform boundedness of (X

n

, ∂

Xn

θ

n

, div (X

n

ω

n

)) in

L

loc

(R

+

; C

ε

)

2

× L

loc

(R

+

; C

ε−2

) × L

loc

(R

+

; C

ε−3

) ∩ L e

1loc

(R

+

; C

ε−1

) ,

it is only a matter of checking that the ‘constant’ C

0n

that appears in the r.h.s. of (3.17) and (3.18) could be bounded independently of n. In fact, besides the norms appearing in (3.20), C

0n

depends only (continuously) on

kX

0n

k

Cε

, k∂

X0n

θ

n0

k

Cε−2

and kdiv (X

0n

ω

n0

)k

Cε−3

.

Arguing as in (3.20), we see that kX

0n

k

Cε

can be uniformly controlled by kX

0

k

Cε

. Furthermore, combining Lemma A.3 and Lemma 2.97 in [2], we get

k∂

X0n

θ

0n

k

Cε−2

. k∂

X0

θ

0

k

Cε−2

+ kθ

0

k

B

2 q−1 q,1

kX

0

k

Cε

. Finally, we claim that

kdiv (X

0n

ω

0n

)k

Cε−3

. kdiv (X

0

ω

0

)k

Cε−3

+ kω

0

k

B

2 q−2 q,1

kX

0

k

Cε

. This is a consequence of the decomposition

div (Y f ) − T

Y

f = div (T

f

Y + R(f, Y )) + [∂

k

, T

Yk

]f, applied to (Y, f ) = (X

0n

, ω

n0

) or (X

0

, ω

0

), and of the fact that

T

X0n

ω

n0

= S

n

T

X0

ω

0

+ [T

(Xn0)k

, S

n

]∂

k

ω

0

+ T

(X0n−X0)

ω

n0

, where the last term T

(Xn0−X0)

ω

n0

vanishes if N

0

in (2.1) is taken larger than 1.

Let us now establish that (θ

n

, u

n

, ∇Π

n

) converges (strongly) to some solution (θ, u, ∇Π) of (B

1,2

) belonging to the space F

20

(R

2

) where, for all s ∈ R, p ∈ [1, ∞] and N ≥ 2, we set

F

ps

(R

N

) := n

(ϑ, v, ∇P ) : ϑ ∈ L e

loc

(R

+

; B

N p−1+s

p,1

(R

N

)), ∇P ∈

L

1loc

R

+

; B

N p−1+s

p,1

(R

N

)

N

and v ∈

L e

loc

R

+

; B

N p−1+s p,1

( R

N

)

∩ L

1loc

R

+

; B

N p+1+s

p,1

( R

N

)

N

o

· To this end, we shall first prove that (θ

n

, u

n

, ∇Π

n

)

n∈N

is a Cauchy sequence in the space F

2−1

(R

2

). Indeed, if setting

(δθ

mn

, δu

mn

, δΠ

mn

) := (θ

m

− θ

n

, u

m

− u

n

, Π

m

− Π

n

),

(12)

then we get from (B

1,N

) that

(B

1,Nm,n

)

 

 

t

δθ

mn

+ div (u

m

δθ

mn

) = −div (δu

mn

θ

n

),

t

δu

mn

+ div (u

m

⊗ δu

mn

) − ∆δu

mn

+ ∇δΠ

mn

= δθ

nm

e

N

− div (δu

mn

⊗ u

n

), div δu

mn

= 0.

In dimension N = 2, we infer from Bony decomposition (2.1) and continuity results for the paraproduct and remainder, that

kdiv (δu

mn

θ

n

)k

B−1

2,1

. kδu

mn

k

B1

2,1

n

k

B0

2,1

and kdiv (δu

mn

⊗ u

n

)k

B−1

2,1

. kδu

mn

k

B−1

2,1

ku

n

k

B2

2,1

.

Using the above two inequalities and the estimates for the transport and transport-diffusion equations stated in Proposition 2.1, we end up with

kδθ

nm

k

Le

t (B2,1−1)

≤ e

Ck∇u

mkL1 t(B1

2,1 )

kδθ

mn

(0)k

B−1

2,1

+ Z

t

0

kδu

mn

k

B2,11

n

k

B2,10

dt

,

E

nm

(t) . (1 + t

2

)e

C(1+t)k∇u

mkL1 t(B1

2,1)

kδu

mn

(0)k

B−1

2,1

+ kδθ

mn

(0)k

B−1

2,1

+ Z

t

0

ku

n

k

B12,1

+ kθ

n

k

B02,1

E

m

n

(t

) dt

, where E

m

n

(t) := kδu

mn

k

Le

t (B−12,1)

+ kδu

mn

k

L1

t(B12,1)

.

Combining Gronwall lemma with the uniform bounds that we proved for (θ

n

, u

n

), we deduce that (θ

n

, u

n

, ∇Π

n

)

n∈N

strongly converges in the norm of k ·k

F−1

2 (R2)

. Then interpolating with the uniform bounds, we see that strong convergence holds true in F

2−η

( R

2

), too, for all η > 0, which allows to justify that (θ, u, ∇Π) satisfies (B

1,2

). Taking advantage of Fatou property of Besov spaces, we gather that the limit (θ, u, ∇Π) be- longs to F

20

( R

2

). In addition, as the sequence (θ

n

, ω

n

) is bounded in E

q0

, Fatou property also implies that (θ, ω) belongs to E

q0

. Finally, using (complex) interpolation, we obtain that for any 0 < η, δ < 1, sequence (θ

n

, ω

n

)

n∈N

converges to (θ, ω) in the space E

q−δηδ

with q

δ

:=

δ2

+

1−δq

.

To complete the proof of Theorem 1.1, we have to check that the announced proprieties of striated regularity are fulfilled. In fact, taking advantage of the (uniform) Lipschitz-continuity of u

n

, we may obtain that for all η > 0 (see [7]),

(3.21) X = lim

n→∞

X

n

in L

loc

(R

+

; C

ε−η

),

which allows to justify that X satisfies (1.3), and also that for all η

> 0, Sequence ∂

Xn

θ

n

, div (X

n

ω

n

)

n∈N

converges in the space

L

loc

( R

+

; C

ε−2−η

) × L

1loc

( R

+

; C

ε−1−η

).

As we know in addition that (3.17) and (3.19) are fulfilled by ∂

Xn

θ

n

and div (X

n

ω

n

) for all n ∈ N, Fatou property enables us to conclude that we have

X

θ ∈ L

loc

(R

+

; C

ε−2

) and div (Xω) ∈ L

loc

(R

+

; C

ε−3

) × L e

1loc

(R

+

; C

ε−1

), and that both (3.17) and (3.19) hold true.

4. Propagation of striated regularity in the general case N ≥ 3

This section is dedicated to the proof of Theorem 1.3. The first (easy) step is to extend the global existence result of R. Danchin and M. Paicu in [15], to nonhomogeneous Besov spaces. The second step is to propagate striated regularity. As pointed out in the introduction, it suffices to bound ∂

X

u in the space L e

1t

( C

ε

), which requires our using the smoothing properties of the heat flow. By applying the directional derivative ∂

X

to the velocity equation, we discover that ∂

X

u may be seen as the solution to some evolutionary Stokes system with a nonhomogeneous divergence condition, namely (using the fact that div u = 0),

div (∂

X

u) = div ∂

u

X − u div X

.

(13)

In order to reduce our study to that of a solution to the standard Stokes system, it is thus natural to decompose ∂

X

u into

X

u = z + ∂

u

X − u div X.

Now, z is indeed divergence free. Unfortunately, the above decomposition introduces additional source terms in the Stokes system satisfied by z. One of those terms is ∆∂

u

X, that we do not know how to handle without losing regularity. Moreover, this also leads to the failure of global estimates technically.

To overcome this difficulty, we here propose to replace the directional derivative ∂

X

by the para-vector field T

X

that is defined by

T

X

u := T

Xk

k

u.

The reason why this change of viewpoint is appropriate is that, as we shall see below, the aforementioned loss of regularity does not occur anymore when estimating T

X

u, and that T

X

u may be seen of the leading order part of ∂

X

u, as shown by the following two inequalities (that hold true whenever

Np

+ ε − 1 ≥ 0, see Lemma A.3):

k∂

X

u − T

X

uk

Lt (Cε−2)

. kX k

L t (Ceε)

kuk

Lt (B

N p−1 p,1 )

, (4.1)

k∂

X

u − T

X

uk

L1t(Cε)

. kX k

Lt (Cε)

kuk

L1t(B

N p+1 p,1 )

. (4.2)

Thus the second step of the proof of Theorem 1.3 will be mainly devoted to proving a priori estimates for T

X

u in L

t

( C

ε−2

) ∩ L e

1t

( C

ε

)

N

.

Finally, in the last step of the proof, we smooth out the data (as in the 2-D case) so as to justify that the a priori estimates of the first two steps are indeed satisfied.

4.1. Global existence in nonhomogeneous Besov spaces. The following result is an obvious modifica- tion of the work by the first author and M. Paicu in [15].

Theorem 4.1. Let N ≥ 3 and p ∈]N, ∞[. Assume that θ

0

∈ B

N,10

( R

N

) ∩ L

N3

( R

N

) and that the initial divergence-free velocity field u

0

is in (B

N p−1

p,1

( R

N

) ∩ L

N,∞

( R

N

))

N

. There exists a (small) positive constant c independent of p such that if

(4.3) ku

0

k

LN,∞

+ ν

−1

0

k

LN3

≤ cν, then the Boussinesq system (B

ν,N

) has a unique global solution

(θ, u, ∇Π) ∈ C(R

+

; B

N,10

) ∩ C(R

+

; B

N p−1

p,1

) ∩ L

1loc

(R

+

; B

N p+1 p,1

)

N

∩ L

1loc

(R

+

; B

N p−1 p,1

)

N

. Moreover, there exists some positive constant C independent on N such that for any t ≥ 0, we have

(4.4) kθ(t)k

LN3

= kθ

0

k

LN3

, ku(t)k

LN,∞

≤ C(ku

0

k

LN,∞

+ ν

−1

0

k

LN3

),

(4.5) U

p

(t) ≤ A(t),

(4.6) kθk

Le

t (BN,10 )

≤ Ckθ

0

k

B0

N,1

(1 + ν

−1

A(t)),

(4.7) k(∂

t

u, ∇Π)k

L1t(B

N p−1

p,1 )

≤ Cν

−1

A

2

(t) + Ctkθ

0

k

BN,10

1 + A(t) , where

U

p

(t) := kuk

Let (B

N p−1

p,1 )

+ νkuk

L1t(B

N p+1 p,1 )

and A(t) := C(ku

0

k

B

N p−1 p,1

+ν)e

−1tkθ0kB0

N,1

+ Cν

2

0

k

B0

N,1

+ ν

2

0

k

2B0

N,1

e

−1tkθ0kB0 N,1

− 1

·

Références

Documents relatifs

On the other hand, if κ = ν = 0 then constructing global unique solutions for some nonconstant θ 0 is a challenging open problem (even in the two-dimensional case) which has

In this paper, we would like to establish rigorous stabilization results for the Boussinesq system, around a stationary solution, by feedback controls of finite dimension,

(An analogous result for the BBM-BBM system was proved in [BC].) Having in hand such a well-posedness result, enables one to consider this type of boundary conditions, realized,

Another interesting case of global existence for (2) corresponding to large initial data but with special structure is the more classical case of axisymmetric solutions without

In this paper we deal with the local exact controllability to the trajectories of the Boussinesq system with N + 1 distributed scalar controls supported in small sets.. In a first

In order to prove the main Theorem 2.1, we recall the following results about the local existence and uniqueness of H k -solution of the inviscid Boussinesq equations (1.1) which is

Schonbek [11] concern- ing the problem of existence of global entropic weak solutions for the classical Boussinesq system, as well as the study of the regularity of these solutions

They showed that the problem is NP -complete in the strong sense even if each product appears in at most three shops and each shop sells exactly three products, as well as in the