• Aucun résultat trouvé

On the small time asymptotics of the two-dimensional stochastic Navier-Stokes equations

N/A
N/A
Protected

Academic year: 2022

Partager "On the small time asymptotics of the two-dimensional stochastic Navier-Stokes equations"

Copied!
18
0
0

Texte intégral

(1)

www.imstat.org/aihp 2009, Vol. 45, No. 4, 1002–1019

DOI: 10.1214/08-AIHP192

© Association des Publications de l’Institut Henri Poincaré, 2009

On the small time asymptotics of the two-dimensional stochastic Navier–Stokes equations

Tiange Xu and Tusheng Zhang

Department of Mathematics, University of Manchester, Oxford Road, Manchester M13 9PL, England, UK. E-mail: tzhang@maths.man.ac.uk Received 1 February 2008; revised 27 August 2008; accepted 10 September 2008

Abstract. In this paper, we establish a small time large deviation principle (small time asymptotics) for the two-dimensional stochastic Navier–Stokes equations driven by multiplicative noise, which not only involves the study of the small noise, but also the investigation of the effect of the small, but highly nonlinear, unbounded drifts.

Résumé. Dans cet article, nous établissons un principe de grandes déviations en temps petit pour l’équation de Navier–Stokes bi-dimensionnelle stochastique conduite par un bruit multiplicatif. Celui-ci nécessite non seulement l’étude d’un bruit faible, mais aussi la compréhension des effets de dérives petites mais non bornées et non linéaires.

MSC:Primary 60H15; secondary 60F10; 35Q30

Keywords:stochastic Navier–Stokes equation; small time asymptotics; large deviation principle

1. Introduction

It is well known that the two-dimensional stochastic Navier–Stokes equation with Dirichlet boundary condition de- scribes the time evolution of an incompressible fluid and is given by

⎧⎪

⎪⎩

du−νudt+(u· ∇)udt+ ∇pdt=gdt+σ (t, u)dW (t ),

(divu)(t, x)=0 for xD, t >0,

u(t, x)=0 for x∂D, t >0,

u(0, x)=u0(x) for xD,

whereDis a bounded domain inR2with smooth boundary∂D,u(t, x)∈R2denotes the velocity field at timet and positionx,p(t, x)denotes the pressure field,ν >0 is the viscosity andW is a Brownian motion on a Hilbert space.

Moreover,σ (t, u)dW is the random force field acting on the fluid andgis the deterministic part of the force.

To formulate the stochastic Navier–Stokes equations, we introduce the following standard spaces:

V = vH01

D;R2

:∇ ·v=0 a.e., inD , with the norm

vV :=

D

|∇v|2dx 1/2

= v

(2)

and denote by((·,·))the inner product ofV.H is the closure ofV in theL2-norm

|v|H:=

D

|v|2dx 1/2

= |v|.

The inner product onHwill be denoted by(·,·).

Define the operatorA(Stokes operator) inHby the formula Au= −νPHu,uH2

D;R2

V ,

where the linear operatorPH(Helmhotz–Hodge projection) is the projection operator fromL2toH, and the nonlinear operatorB

B(u, v)=PH

(u· ∇)v ,

with the notationB(u)=B(u, u). Obviously the domain ofBrequires that(u· ∇v)belongs to the spaceL2. By applying the operatorPH to each term of the above stochastic Navier–Stokes equation (SNSE), we can rewrite the SNSE in the following abstract form:

du(t )+Au(t )dt+B(u(t ))dt=f (t )dt+σ t, u(t )

dW (t ) inL2(0, T;V) (1.1)

with the initial condition

u(0)=u0 inH, (1.2)

whereW (t )is a Brownian motion taking values in a Hilbert spaceU, andVis the dual ofV.

There exists a great amount of literature on the stochastic Navier–Stokes equation. Let us mention some of them.

A good reference for stochastic Navier–Stokes equations driven by additive noise is the book [4] and the references therein. The existence and uniqueness of solutions for stochastic 2-D Navier–Stokes equations with multiplicative noise were obtained in [8], [13] and [14]. The ergodic properties and invariant measures of the stochastic 2-D Navier–

Stokes equations were studied in [9] and [11]. The small noise large deviation of the stochastic 2-D Navier–Stokes equations was established in [14] and the large deviation of occupation measures was considered in [10].

The purpose of this paper is to study the small time asymptotics (large deviations) of the two-dimensional sto- chastic Navier–Stokes equations driven by multiplicative noise onC([0,1];H ). After the existence, uniqueness and continuity of the solution of SNSE are established, it is still generally difficult to quantify the behavior of the solution at a position x and a positive timet. It might be relatively simple to estimate the limiting behavior of the solution in time interval[0, t]ast goes to zero, which describes the behavior of the velocity of the fluid at a given point in space when the time is very small. This is one of the motivations for studying the small time asymptotics. Another motivation will be to get the following Varadhan identity through the small time asymptotics:

lim

t02tlogP

u(0)B, u(t )C

= −d2(B, C), (1.3)

whered is an appropriate Riemannian distance associated with the diffusion generated by the solution of the SNSE.

This will be a topic of future study.

Apart from the above motivations, the small time asymptotics is also theoretically interesting, since the study involves the investigation of the small noise and the effect of the small, but highly nonlinear drift. We like to mention that the study of the small time asymptotics (large deviations) of finite dimensional diffusion processes was initiated by Varadhan in the influential work [16]. The small time asymptotics of infinite dimensional diffusion processes were studied in [1,2,7,12] and [17].

To establish the small time large deviation for the stochastic Navier–Stokes equation, as one expects, the main difficulty lies in dealing with the nonlinear term B(u)=PH((u· ∇)u) and the unbounded term Au= −νPHu, which barely belong toV. To controlB(u), our idea is to show that the probability that the solution stays outside an energy ball is exponentially small so that we can restrict the solution in a sufficiently large energy ball. We hope

(3)

that this method used for treating the non-linear term could be useful in some other study of stochastic Navier–Stokes equations. Another key step of obtaining the small time large deviation is to prove that the law of the solution of the Navier–Stokes equationu(εt )is exponentially equivalent to the law of the solutions of the equation:

vε(t )=x+√ ε

t

0

σ

εs, vε(s)

dW (s). (1.4)

This is done through several approximations. A remarkable martingale inequality proved by Barlow, Davis and Yor in [5] and [3], plays an important role throughout the paper. We point out that this martingale inequality is more precise than most of the B–D–G’s inequalities stated in the literature because the constant on the right isp1/2(see (3.11)), but notp.

The rest of the paper is organized as follows. In Section 2, we collect some preliminaries which are frequently used in the sequel. Section 3 is the main part of the paper, where we proved the small time large deviations for stochastic Navier Stokes equation. In Section 4, we further relax the conditions on the diffusion coefficientσ (·).

2. Preliminaries

IdentifyingHwith its dualH, we consider Eq. (1.1) in the framework of Gelfrand triple:

VH∼=HV.

In this way, we may considerAas a bounded operator fromV intoV. Moreover, we also denote by·,·, the duality betweenV andV. Hence, foru=(ui)V,w=(wi)V, we have

Au, w =ν

i,j

D

iujiwjdx=ν (u, w)

. (2.1)

Introduce a trilinear form onH×H×Hby setting b(u, v, w)=

2 i,j

D

uiivjwjdx, (2.2)

whenever the integral in (2.2) makes sense. In particular, ifu, v, wV, then B(u, v), w

=

(u· ∇v), w

= 2

i,j

D

uiivjwjdx=b(u, v, w).

By the integration by parts,

b(u, v, w)= −b(u, w, v), (2.3)

therefore

b(u, v, v)=0, ∀u, vV . (2.4)

There are some well-known estimates forb(see [15], for example), which will be required in the rest of this paper and we list them here. Throughout the paper, we denote various generic positive constants by the same letterc. We have

b(u, v, w)cu · v · w|, (2.5)

b(u, v, w)c|u| · v · |Aw|, (2.6)

b(u, v, w)cu · |v| · |Aw|, (2.7)

b(u, v, w)≤2u1/2· |u|1/2· w1/2· |w|1/2· v (2.8)

(4)

for suitableu, v, w.Moreover, combining (2.3) and (2.8), we obtain a useful estimate as follows:

|Bu|V= sup

v1

b(u, u, v)= sup

v1

b(u, v, u)≤2u · |u|. (2.9)

3. Small time asymptotics

First, we introduce the precise assumptions on σ. Let(Ω,F, P ) be a complete probability space with a filtration Ft, t≥0 that satisfies the usual conditions. LetW (·)be aH-valued Brownian motion on(Ω,F, P )with the covari- ance operatorQ, which is a positive, symmetric, trace class operator onH. LetLQdenote the class of linear operators T such thatT Q1/2is a Hilbert–Schmidt operator fromH toH.LQis endowed with the norm|T|2LQ=Tr(T QT), where Tr(T QT)denotes the trace of operatorT QT. LetLVQdenote the class of linear operatorsT˜ such thatT Q˜ 1/2 is a Hilbert–Schmidt operator fromH toV, endowed with the norm| ˜T|2LV

Q=Tr(T Q˜ T˜). Introduce:

(A.1) E|u0|4<+∞, fL4(0, T;V).

(A.2) There exists a constantLsuch that|σ (t, u)|2LQL(1+ |u|2),for allt(0, T ),and alluH. (A.3) There exists a constantLˆ such that for|σ (t, u)|2LV

Q

≤ ˆL(1+ u2),for allt(0, T ),and alluV. (A.4) There exists a constantKsuch that|σ (t, u)σ (t, v)|2LQK|uv|2,for allt(0, T ),and allu, vH. (A.5) There exists a constantKˆ such that|σ (t, u)σ (t, v)|2LV

Q

≤ ˆKuv2,for allt(0, T )and allu, vV. It is known that (see, for example, [14]), under the assumptions (A.1), (A.2) and (A.4), the SNSE

du(t )+Au(t )dt+B u(t )

dt=f (t )dt+σ t, u(t )

dW (t ),

u(0)=xH (3.1)

has a unique strong solutionuL2;C([0, T];H ))L2× [0, T];V ), i.e., u(t )=x

t

0

Au(s)ds− t

0

B u(s)

ds+ t

0

f (s)ds+ t

0

σ s, u(s)

dW (s). (3.2)

Let ε >0, by the scaling property of the Brownian motion, it is easy to see thatu(εt ) coincides in law with the solution of the following equation:

uε(t )=xε t

0

Auε(s)ds−ε t

0

B uε(s)

ds+ε t

0

f (εs)ds+√ ε

t

0

σ

εs, uε(s)

dW (s). (3.3)

Letμεxbe the law ofuε(·)onC([0,1];H ).Define a functionalI (g)onC([0,1];H )by I (g)= inf

hΓg

1 2

1

0

h(t )˙ 2

H0dt

,

whereH0:=Q1/2Hendowed with the norm|h|2H0= |Q1/2h|2(hH0), and Γg=

hC

[0, T];H

: h(·)is absolutely continuous and such that g(t )=x+

t

0

σ

s, g(s)h(s)˙ ds,0≤t≤1

.

(5)

Theorem 3.1. μεxsatisfies a large deviation principle with the rate functionI (·),that is, (i) For any closed subsetFC([0,1];H ),

lim sup

ε0,xnx

εlogμεx

n(F )≤ −inf

gF

I (g) . (ii) For any open subsetGC([0,1];H ),

lim inf

ε0,xnxεlogμεx

n(G)≥ −inf

gG

I (g) .

Proof. Letvε(·)be the solution of the stochastic equation vε(t )=x+√

ε t

0

σ

εs, vε(s)

dW (s), (3.4)

andνε be the law ofvε(·)on the C([0,1];H ). Then by [4], we know that νε satisfies a large deviation principle with the rate functionI (·). Our main task is to show that two families of the probability measuresμε andνε are exponentially equivalent, that is, for anyδ >0,

εlim0εlogP

sup

0t1

uε(t )vε(t )2> δ

= −∞. (3.5)

Then Theorem 3.1 follows from (3.5) and Theorem 4.2.13 in [6] forxn=x.Slight modifications of the proof yields

the general case.

Because of the non-linear formB(·,·,·), and the unbounded operatorA, the proof of (3.5) is quite involved. We split it into several lemmas. The following result is an estimate of the probability that the solution of (3.3) leaves an energy ball.

Lemma 3.1.

Mlim→∞ sup

0<ε1

εlogPuεH

V(1)2

> M

= −∞, (3.6)

where(|uε|HV(1))2=sup0t1|uε(t )|2+εν1

0uε(t )2dt. Proof. SinceB(uε(s)), uε(s) =0, applying Itô’s formula , we get

uε(t )2= |x|2−2ε t

0

uε(s), Auε(s)ds+2ε t

0

uε(s), f (εs) ds +2√

ε t

0

uε(s), σ

εs, uε(s) dW (s)

+ε t

0

σ

εs, uε(s)2

LQds, that is,

uε(t )2+2εν t

0

uε(s)2ds= |x|2+2ε t

0

uε(s), f (εs)

ds+2√ ε

t

0

uε(s), σ

εs, uε(s) dW (s) +ε

t

0

σ

εs, uε(s)2

LQds

= |x|2+l1+l2+l3. (3.7)

Forl1, we have

|l1| ≤2ε t

0

uε(s)·f (εs)

Vds≤εν t

0

uε(s)2ds+1

ν|f|2L2(0,εt;V). (3.8)

(6)

In view of (A.2),

|l3| ≤ε·L t

0

1+uε(s)2

ds. (3.9)

Putting(3.7), (3.8)and(3.9)together, it follows that uε(t )2+εν

t

0

uε2ds≤

|x|2+εLt+1

ν|f|2L2(0, εt;V)

+εL

t

0

uε(s)2ds +2√

ε t

0

uε(s), σ

εs, uε(s)dW (s). Therefore,

uεH

V(T )2

≤2

|x|2+εLT+1

ν|f|2L2(0,εT;V)

+2εL

T

0

uε(s)H

V(s)2

ds +4√

ε sup

0tT

t

0

uε(s), σ

εs, uε(s)dW (s). Hence, forp≥2, we have,

EuεH

V(T )2p1/p

≤2

|x|2+εLT+1

ν|f|2L2(0,εT;V)

+2εL

E

T

0

uεH

V(s)2

ds p1/p

+4√ ε

E

sup

0tT

t

0

(uε(s), σ

εs, uε(s)

dW (s)) p1/p

. (3.10)

To estimate the stochastic integral term, we will use the following remarkable result from [3] and [5] that there exists a universal constantcsuch that, for anyp≥2 and for any continuous martingale(Mt)withM0=0, one has

Mt

pcp1/2M1/2t

p, (3.11)

whereMt=sup0st|Ms|and · pstands for theLP-norm.

Using this result, we have 4√

ε

E

sup

0tT

t

0

uε(s), σ

εs, uε(s)

dW (s)p1/p

≤4c√

E

T

0

uε(s)2·σ

εs, uε(s)2

LQds

p/21/p

≤4c√

E

T

0

uε(s)2

1+uε(s)2 ds

p/21/p

≤4c√

E

T 0

1+uε(s)22

ds

p/22/p1/2

≤4c√

E

T

0

1+uε(s)4 ds

p/22/p1/2

≤4c√

T

0

1+

Euε(s)2p2/p

ds 1/2

, (3.12)

(7)

wherecis a constant which may change from line to line. On the other hand, 2εL

E

T

0

uεH

V(s)2

ds p1/p

≤2εL T

0

EuεH

V(s)2p1/p

ds. (3.13)

Combining(3.10), (3.12)and(3.13), we arrive at EuεH

V(T )2p2/p

≤8

|x|2+εLT +1

ν|f|2L2(0, εT;V)

2

+8ε2L2T T

0

EuεH

V(s)2p2/p

ds +32c2pεT +32c2

T

0

EuεH

V(s)2p2/p

ds. (3.14)

Applying the Gronwall’s inequality, we obtain EuεH

V(1)2p2/p

8

|x|2+εL+1

ν|f|2L2(0,ε;V)

2

+32c2

·exp

2L2+32c2

. (3.15)

SinceP ((|uε|HV(1))2> M)MpE(|uε|HV(1))2p, letp=1ε in(3.15)to get εlogPuεH

V(1)2

> M

≤ −logM+logEuεH

V(1)2p1/p

≤ −logM+log

8

|x|2+εL+1

ν|f|2L2(0, ε;V)

2

+32c2

+4ε2L2+16c2. Therefore,

sup

0<ε1

εlogPuεH

V(1)2

> M

≤ −logM+log

8

|x|2+L+1

ν|f|2L2(0,1;V)

2

+32c2

+16c2+4L2.

LettingM→ ∞on both side of the above inequality, we finish the proof.

SinceV is dense inH, there exists a sequence{xn}n=1V such that

n→+∞lim |xnx| =0.

Letuεn(·)be the solution of (3.3) with the initial valuexn. From the proof of Lemma 3.1, it follows that

M→+∞lim sup

n

sup

0<ε1

εlogPuεnH

V(1)2

> M

= −∞. (3.16)

Letvεn(·)be the solution of(3.4)with the initial valuexn, and we have the following result.

Lemma 3.2. For anyn∈Z+,

Mlim→∞ sup

0<ε1

εlogP

sup

0t1

vnε(t )2> M

= −∞.

(8)

Proof. Applying Itô’s formula tovnε(t )2, one obtains vnε(t )2= xn2+2√

ε t

0

vεn(s), σ

εs, vnε(s)

dW (s) +ε

t

0

σ

εs, vεn(s)2

LVQds.

By (A.3) and inequality(3.11), we have

E

sup

0tr

vnε(t )2p2/p

≤2xn4+8cεp

E r

0

vεn(s)2σ

εs, vεn(s)2

LVQds

p/22/p

+4ε2Lˆ2r

r+ r

0

E

sup

0ls

vεn(l)2p2/p

ds

≤2xn4+16cεpLˆ

r+ r

0

E

sup

0ls

vεn(l)2p2/p

ds

+4ε2Lˆ2r

r+ r

0

E

sup

0ls

vεn(l)2p2/p

ds

. (3.17)

By Gronwall’s inequality,

E

sup

0t1

vεn(t )2p2/p

2xn4+16cεpLˆ+4ε2Lˆ2

e16cεpLˆ+2Lˆ2. (3.18)

The rest of the proof is the same as that of Lemma 3.1.

Lemma 3.3. For anyδ >0,

n→+∞lim sup

0<ε≤1

εlogP

sup

0≤t≤1

uε(t )uεn(t )2> δ

= −∞. (3.19)

Proof. ForM >0, define a stopping time τε,M=inf

t: εν

t

0

uε(r)2dr > M, oruε(t )2> M

. Clearly,

P

sup

0t1

uε(t )uεn(t )2> δ,uεH

V(1)2

MP

sup

0t1

uε(t )uεn(t2> δ, τε,M≥1

P

sup

0t1τε,M

uε(t )uεn(t )2> δ

. (3.20)

Letkbe a positive constant. Applying Itô’s formula to e

t∧τε,M

0 uε(s)2ds|uε(tτε,M)uεn(tτε,M)|2, we get e

t∧τε,M

0 uε(s)2dsuε(tτε,M)uεn(tτε,M)2+2εν tτε,M

0

e

s

0uε(r)2druε(s)uεn(s)2ds

= |xxn|2 tτε,M

0

e0suε(r)2druε(s)2uε(s)uεn(s)2ds

−2ε t∧τε,M

0

e

s

0uε(r)2dr b

uε(s), uε(s), uε(s)uεn(s)

b

uεn(s), uεn(s), uε(s)uεn(s) ds

(9)

+ε tτε,M

0

e0suε(r)2drσ

εs, uε(s)

σ

εs, uεn(s)2

LQds +2√

ε tτε,M

0

e

s

0uε(r)2dr

uε(s)uεn(s), σ

εs, uε(s)

σ

εs, uεn(s) dW (s)

. (3.21)

Notice that by (2.4), we have b

uεn(t ), uεn(t ), uε(t )uεn(t )

=b

uεn(t ), uε(t ), uε(t )uεn(t ) , hence by (2.9),

b

uε(t ), uε(t ), uε(t )uεn(t )

b

uεn(t ), uεn(t ), uε(t )uεn(t )

=b

uε(t )uεn(t ), uε(t ), uε(t )uεn(t )

≤2uε(t )uεn(t )·uε(t )·uε(t )uεn(t ). Therefore, by(3.21),

e

tτε,M

0 uε(s)2dsuε(tτε,M)uεn(tτε,M)2+2εν tτε,M

0

e

s

0uε(r)2druε(s)uεn(s)2ds

≤ |xxn|2 tτε,M

0

e

s

0uε(r)2druε(s)2uε(s)uεn(s)2ds +4ε

tτε,M

0

e

s

0uε(r)2druε(s)uεn(s)·uε(s)·uε(s)uεn(s)ds +εK

tτε,M

0

e0suε(r)2druε(s)uεn(s)2ds +2√

ε tτε,M

0

e

s

0uε(r)2dr

uε(s)uεn(s), σ

εs, uε(s)

σ

εs, uεn(s)dW (s)

≤ |xxn|2 t∧τε,M

0

e

s

0uε(r)2druε(s)2uε(s)uεn(s)2ds +εν

tτε,M

0

e

s

0uε(r)2druε(s)uεn(s)2ds +4ε

ν

tτε,M

0

e0suε(r)2druε(s)uεn(s)2·uε(s)2ds +εK

t∧τε,M

0

e

s

0uε(r)2druε(s)uεn(s)2ds +2√

ε tτε,M

0

e

s

0uε(r)2dr

uε(s)uεn(s), σ

εs, uε(s)

σ

εs, uεn(s)dW (s). (3.22) Choosingk >4ν and using(3.11), we have,

E

sup

0stτε,M

ek0suε(r)2druε(s)uεn(s)2p2/p

≤2|xxn|4+2ε2K2 t

0

E

sup

0rsτε,M

e0ruε(l)2dluε(r)uεn(r)2p2/p

ds +8cεpK2

t 0

E

sup

0rsτε,M

e2kε

r

0uε(l)2dluε(r)uεn(r)4p/22/p

ds

(10)

≤2|xxn|4+2ε2K2 t

0

E

sup

0rsτε,M

e0ruε(l)2dluε(r)uεn(r)2p2/p

ds +8cεpK2

t

0

E

sup

0rsτε,M

e0ruε(l)2dluε(r)uεn(r)2p2/p

ds. (3.23)

Applying Gronwall’s inequality, one obtains,

E

sup

0t1τε,M

e

t

0uε(s)2dsuε(t )uεn(t )2p2/p

≤2|xxn|4e2K2+8cεpK2. (3.24)

Hence,

E

sup

0t1τε,M

uε(t )uεn(t )2p2/p

E

sup

0t1τε,M

e0tuε(s)2dsuε(t )uεn(t )2p

ekpε

1∧τε,M

0 uε(s)2ds2/p

≤e2kM

E

sup

0t1τε,M

e

t

0uε(s)2dsuε(t )uεn(t )2p2/p

≤2e2kM|xxn|4e2K2+8cεpK2. (3.25)

FixM, and takep=2/εto get sup

0<ε1

εlogP

sup

0t1τε,M

uε(t )uεn(t )2> δ

≤ sup

0ε1

εlogE[sup0t1τε,M|uε(t )uεn(t )|2p] δp

≤2kM+2K2+16cK2−2 logδ+log 2|xxn|4

→ −∞, asn→ +∞. (3.26)

By Lemma 3.1, for anyR >0, there exists a constantMsuch that for anyε(0,1], the following inequality holds, PuεH

V(1)2

> M

≤eR/ε. (3.27)

For such aM, by(3.20), (3.26), there exits a positive integerN, such that for anynN, sup

0<ε1

εlogP

sup

0t1

uε(t )uεn(t )2> δ,uεH

V(1)2

M

≤ −R. (3.28)

Putting(3.27)and(3.28)together, one sees that there exists a positive integerN, such that for anynN, ε(0,1] P

sup

0t1

uε(t )uεn(t )2> δ

≤2eR/ε. (3.29)

SinceRis arbitrary, the lemma follows.

We have the following similar result forvε(·)andvnε(·).

Références

Documents relatifs

The proofs of Theorems 2.2 and 2.5 are based on the representations of the density and its derivative obtained by using Malliavin calculus and on the study of the asymptotic behavior

However, a uniform bound on the vorticity does not allow to control the L ∞ norm of the velocity field, hence a priori estimates are not sufficient to prove that solutions of (1.1)

Combining the weighted energy and enstrophy estimates of Section 4 with the results of Section 3, we prove in Section 5 the first two assertions of Theorem 1.2, namely the uniform

Unfortunately, the results of [5] do not apply directly to the Navier-Stokes equations (1.1), (1.2), because we do not know a priori if the energy density (1.9) stays uniformly

In Section 2, we prove the existence of global solutions of the vorticity equation (9) in a neighborhood of the origin, and we estimate their decay rate as t → ∞.. The results we

In this paper, in order to approximate the solution of discretized problem (10, 11), we use the Galerkin method for the momentum and continuity equations which allows to circumvent

Then one could use the same type of arguments as in [8] to prove a smoothing property but another difnculty appears. The introduced exponential factor is very large and we are not

« infinité order », they have become two of the most efficient numerical methods for solving the nonlinear partial differential équations arising in fluid dynamics (see [1-7])..