• Aucun résultat trouvé

Incompressible, inviscid limit of the compressible Navier-Stokes system

N/A
N/A
Protected

Academic year: 2022

Partager "Incompressible, inviscid limit of the compressible Navier-Stokes system"

Copied!
26
0
0

Texte intégral

(1)

2001 Éditions scientifiques et médicales Elsevier SAS. All rights reserved S0294-1449(00)00123-2/FLA

INCOMPRESSIBLE, INVISCID LIMIT OF THE COMPRESSIBLE NAVIER–STOKES SYSTEM

Nader MASMOUDI1

Ceremade, URA CNRS 749, Université Paris IX Dauphine, Place du Maréchal de Lattre de Tassigny, 75775 Paris Cedex 16, France

Received in 2 February 2000

ABSTRACT. – We prove some asymptotic results concerning global (weak) solutions of compressible isentropic Navier–Stokes equations. More precisely, we establish the convergence towards solutions of incompressible Euler equations, as the density becomes constant, the Mach number goes to 0 and the Reynolds number goes to infinity. 2001 Éditions scientifiques et médicales Elsevier SAS

RÉSUMÉ. – Nous prouvons quelques résultats asymptotiques concernant des solutions (faibles) globales des équations de Navier–Stokes (isentropique) compressible. Plus précisément, nous établissons la convergence vers une solution des équations d’Euler incompressible, lorsque la densité devient constante, le nombre de Mach tend vers 0 et le nombre de Reynolds tend vers l’infini.2001 Éditions scientifiques et médicales Elsevier SAS

1. Introduction

From a physical point of view, one can formally derive incompressible models from compressible ones, when the Mach number goes to zero and the density becomes almost constant. In Lions and the author [12], this problem is investigated starting form the global solutions of the compressible Navier–Stokes equations constructed by Lions [11].

We have shown the convergence towards the incompressible Navier–Stokes equations as well as the convergence towards the incompressible Euler equations (if the viscosity coefficients go to zero and if the initial data are “well prepared”). These results have been precised and extended in different works (see [13,2,1,14]).

In this paper, we extend the result shown in [12] concerning the convergence to the Euler system to the case of more general initial data. In fact if the viscosity goes to zero too, we loose spatial compactness properties. In order to circumvent this difficulty, we use energy arguments. Hence, we have to describe (precisely) the oscillations that take place and include them in the energy estimates. Ideas of this type were introduced by

E-mail address: masmoudi@cims.nyu.edu (N. Masmoudi).

1Current adress: Courant Institute, New York University 251 Mercer St, New York, NY 10012, USA.

(2)

Schochet [17], (see also [4]) and extended to the case of vanishing viscosity coefficients in [16]. Let us now precise the scalings, we are going to use, which are the same as those used in [12]. The unknowns (ρ, v)˜ are respectively the density and the velocity of the fluid (gaz). We scaleρ˜ andv(and thusp) in the following way

˜

ρ=ρ(x, εt), v=εu(x, εt), (1)

and we assume that the viscosity coefficientsµ,˜ ξ˜ (ξ˜ − ˜µ,µ˜ are called the volumic and dynamical viscosity coefficients) are also small and scale like

˜

µ=εµε, ξ˜=εξε, (2)

where ε(0,1) is a “small parameter” and the normalized coefficient µε, ξε satisfy µε>0, ξε+µε>0. Moreover, we assume that

µε→0 asεgoes to 0+. (3)

With the preceding scalings, the compressible Navier–Stokes system reads

∂ρ

∂t +div(ρu)=0, ρ0,

∂ρu

∂t +div(ρu⊗u)µεuξε∇divu+ a

ε2ργ =0.

(4)

One can always assume thata=1/γ by replacingεby√aγ ε. All throughout this paper the domainwill be the the whole space RN or the torusTN (in this last case, all the functions are defined on RN and assumed to be periodic with period 2π ai in the ith variable). We recall here that the inviscid limit, namely the convergence of the Navier–

Stokes equations to Euler equations in the case of a domain with boundary is an open problem even in the incompressible case, which seems to be easier (see [15] for a partial result).

1.1. Statement of the results 1.1.1. The whole space case

Let us begin with the case of the whole space. We consider a sequence of global weak solutionsε, uε)of the compressible Navier–Stokes equations (4) and we assume thatρε−1∈L(0,∞;Lγ2)C([0,∞), Lp2)for all 1p < γ ,whereLp2 = {fL1loc,

|f|1|f|1Lp, |f|1|f|1L2}, uεL2(0, T; H1) for all T(0,∞) (with a norm which can explode when ε goes to 0),ρε|uε|2L(0,∞; L1)and ρεuεC([0,∞);

L2γ /(γ+1)w) i.e. is continuous with respect tot0 with values inL2γ /(γ+1)endowed with its weak topology. We require (4) to hold in the sense of distributions and we impose the following conditions at infinity

ρε→1 as|x| → +∞, uε→0 as|x| → +∞. (5)

(3)

Finally, we prescribe initial conditions

ρε|t=0=ρε0, ρεuε|t=0=m0ε, (6) whereρε00,ρε0−1∈Lγ,m0εL2γ /(γ+1),m0ε=0 a.e. on{ρε0=0}andρε0|u0ε|2L1, denoting by u0ε =m0εε0 on {ρ0ε >0}, u0ε =0 on {ρε0=0}. The initial conditions also satisfy the following uniform bounds

ρε0u0ε

2+ 1

ε2 ρε0

γ −1−γρε0−1C, (7)

where, here and below, C denotes various positive constants independent ofε. Let us notice that (7) implies in particular that, roughly speaking, ρε0is of order 1+O(ε).In the sequel, we will use the following notation ρε=1+εϕε. Notice that ifγ <2, we cannot deduce any bound forϕε inL(0, T;L2). This is why we prefer to work with the following approximation

Φε=1 ε

2a γ −1

ρεγ−1−γ (ρε−1).

Furthermore, we assume that

ρε0u0ε converges strongly in L2 to someu0. Then, we denote byu0=Pu˜0, whereP is the projection on divergence-free vector fields, we also defineQ(the projection on gradient vector fields), henceu˜0=Pu˜0+Qu˜0. Moreover, we assume thatΦε0converges strongly inL2to someϕ0. This also implies thatϕε0converges toϕ0inLγ2.

Our last requirement onε, uε)concerns the total energy: we assume that we have Eε(t)+

t 0

Dε(s)dsEε0 a.e.t, dEε

dt +Dε0 inD(0,), (8)

where

Eε(t)=

1

2ρε|uε|2(t)+ a ε2 −1)

ε)γ −1−γ (ρε−1)(t),

Dε(t)=

µε|Duε|2(t)+ξε(divuε)2(t) and

Eε0=

1

2ρε0u0ε2+ a ε2 −1)

ρε0γ −1−γρ0ε−1.

We now wish to emphasize the fact that we assume the existence of a sequence of solutions with the above properties, and we shall also assume that γ > N/2. We recall

(4)

the results of Lions [11] which yield the existence of such solutions precisely when γ > N/2 ifN 4,γ 9/5 ifN =3 andγ 3/2 ifN =2. We also refer to [12] for the proof of the uniform bounds.

Whenεgoes to zero andµε goes to 0, we expect thatuεconverges tov, the solution of the Euler system

tv+div(v⊗v)+ ∇π=0,

divv=0 v|t=0=u0, (9)

inC([0, T);Hs). We show the following theorem

THEOREM 1.1 (The whole space case). – We assume thatµεε0 (such thatµε+ ξε>0 for allε) and thatPu˜0Hs for somes > N/2+1, thenP (

ρεuε)converges to vinL(0, T;L2)for allT < T, wherevis the unique solution of the Euler system in Lloc([0, T);Hs)andTis the existence time of(9). In additionρεuε converges tov inLp(0, T;L2loc)for all 1p <+∞and allT < T.

1.1.2. The periodic case

Now, we take=TN and consider a sequence of solutionsε, uε)of (4), satisfying the same conditions as in the whole space case (the functions are now periodic in space and all the integration are performed over TN). Of course, the conditions at infinity are removed and the spaces Lp2 can be replaced byLp. Here, we have to impose more conditions on the oscillating part (acoustic waves), namely we have to assume thatQu˜0 is more regular than L2. In fact, in the periodic case, we do not have a dispersion phenomenon as in the case of the whole space and the acoustic waves will not go to infinity, but they are going to interact with each other. This is way, we have to include them in the energy estimates to show our convergence result.

For the next theorem, we assume that Qu˜0, ϕ0Hs1 and that there exists a nonnegative constantν such thatµε+ξε2ν >0 for allε. For simplicity, we assume thatµε+ξεconverges to 2ν.

THEOREM 1.2 (The periodic case). – We assume thatµεε0 (such thatµε+ξε→ 2ν >0) and thatPu˜0Hsfor somes > N/2+1, andQu˜0, ϕ0Hs1thenP (

ρεuε) converges to v in L(0, T;L2) for all T < T, wherev is the unique solution of the Euler system inLloc(0, T;Hs)andT is the existence time of(9). In addition

ρεuε

converges weakly tovinL(0, T;L2).

In the above theorem, one can remove the condition 2ν >0. In that case, we still have the result of Theorem 1.2 but only on an interval of time(0, T∗∗)which is the existence interval for the equation governing the oscillating part, which will be given later on.

THEOREM 1.3 (The periodic case, ν =0). – We assume that µεε0 (such that µε +ξε0) and thatu˜0Hs for some s > N/2+1, and ϕ0Hs then P (ρεuε) converges tovinL(0, T;L2)for allT < inf(T, T∗∗), wherevis the unique solution of the Euler system in Lloc(0, T;Hs), T is the existence time of (9) and T∗∗ the existence time of (32), with ν = 0. In addition

ρεuε converges weakly to v in L(0, T;L2).

(5)

2. The whole space case

We recall that in the case of the whole space, we do not assume any extra condition on the viscosityξε, neither do we assume any regularity (more thanL2) for the gradient part of the initial data. The proof relies on the dispersion property of the wave equation [18,1] and the notion of non dissipative solutions for the Euler system [10,12].

First, using the energy bounds, we deduce thatρε−1 converges to 0 inL(0, T;Lγ2) and that there exists some uL(0, T;L2) and a subsequence √ρεuε converging weakly to u. Hence, we also deduce thatρεuε converges weakly to u in L2γ /(γ+1). We next introduce the following group(L(τ ), τ ∈R)defined by eτ LwhereLis the operator defined onD×(D)N, by

L ϕ

v

= − divv

ϕ

. (10)

It is easy to check that eτ L is an isometry on eachHs×(Hs)N for alls∈Rand for all τ. This will allow

eτ L ϕ

v

= ϕ(τ )

v(τ )

solves

∂ϕ

∂τ = −divv, ∂v

∂τ = −∇ϕ and thus ∂τ2ϕ2ϕ=0.

Letε, mε= ∇qε)be the solution of the following system

∂ψε

∂t = −1

εdivmε, ψε(t=0)=Φε0,

∂mε

∂t = −1

ε∇ψε, mε(t=0)=Q

ρ0εu0ε.

(11)

We recall that for all vHs, we define Qv = ∇1divv and P v=vQv. We also introduce the following regularizations ψε,δ=ψεχδ, ∇qε,δ= ∇qεχδ, where χC0(RN) such that χ =1 and χδ(x)=(1/δN)χ (x/δ). Since (11) is linear it is easy to see that ε,δ,qε,δ) is a solution of the same system with regularized initial data. Using (as in [1]) Strichartz type inequality, we get

ψε,δ

qε,δ

Lp(R;Ws,q(RN)))

1/p

Φε0 Q(

ρε0u0ε)

χδ

Hs+σ

(12) for allp, q >2 andσ >0 such that

2

q =(N−1) 1

2− 1 p

, σ=N+1 N−1.

(6)

This yields that for all fixedδand for alls∈R, we have asεgoes to 0

ε,δ,∇qε,δ)→0 inLpR;Ws,qRN. (13) Now, we turn to the energy estimates. Let us rewrite the energy inequality for almost allt

1 2

ρε|uε|2(t)+Φε2(t)+ t

0

µε|Duε|2+ξε(divuε)21 2

ρε0u0ε2+Φε2(0). (14) Then, the conservation of energy forvreads

1

2|v|2(t)=

1

2|u0|2, (15)

and using the fact thatLis an isometry onL2, we obtain for allt

1

2ψε,δ2 (t)+1

2|∇qε,δ|2(t)=

1

2ψε,δ2 (0)+1

2|∇qε,δ|2(0). (16) Next, the weak formulation of the conservation of mass yields for almost allt

ψε,δϕε(t)+1 ε

t 0

div(ρεuεε,δ+div(∇qε,δε=

ψε,δϕε(0), (17) while the weak formulation of (4) implies that we have for almost allt

ρεuε.v(t)+ t

0

ρεuε.(v.v+ ∇p)

t 0

ρεuεuε.v+µε

t 0

uε.v=

ρε0u0ε.u0, (18)

ρεuε.qε,δ(t)+ t 0

ρεuε. 1

εψε,δ

t 0

ρεuεuε.mε,δ

+ t 0

µε∇uε.∇mε,δ+ξεdiv(uε)div(mε,δ)

t 0

1

εϕε+γ −1 2 Φε2

div(∇qε,δ)=

ρε0u0ε.m0ε,δ. (19) Summing up (14), (15), (16) and subtracting (17), (18), (19), we deduce from straightforward computations the following inequality

(7)

1 2

ρεuεvmε,δ2(t)+εψε,δ)2(t)+ t 0

µε|∇uε|2+ξεdiv(uε)2

εϕεε,δ(t)+

εϕεε,δ(0) +

ρε−1

ρεuε.(v+mε,δ)(t)

ρε0−1ρε0u0ε.v0+m0ε,δt

0

µε|∇uε|2+ξεdiv(uε)2

+ t 0

µεuε.(v+mε,δ)+ξεdiv(uε)div(mε,δ)

+ t 0

ρεuε.(v.v+ ∇p)ρεuε(uε.v) ()

t 0

ρεuε(uε.∇mε,δ)

γ −1 2 Φε2

div(∇qε,δ) +1

2

ρε0u0εv0m0ε,δ2+εψε,δ)2(0). (20) Only the term marked by () in the second hand side of (20) needs some special treatment, we are going to compute it below. In the sequel, we denote by wε,δ =

ρεuεvmε,δ. Then, we have easily t

0

ρεuε.(v.v+ ∇p)ρεuε(uε.v)

= − t 0

wε,δ.vwε,δ+ t 0

ρε−√ ρε

uε.(v.v)+ρεuε.p

ρεuεv.∇v2 2 −

t 0

mε,δ.∇vwε,δ+

ρεuεv.∇v.mε,δ. (21) Finally, we can see that (20) may be rewritten as

wε,δ(t)2L2+Φεψε,δ(t)2L2

wε,δ(0)2L2+Φεψε,δ(0)2L2+2Aδε+2 t 0

wε,δ(s)2L2v(s)Lds, (22)

where

(8)

Aδε= −

εϕεε,δ(t)+

εϕεε,δ(0) +

ρε−1

ρεuε.(v+mε,δ)(t)ρε0−1ρε0u0ε.v0+m0ε,δ

+ t

0

µεuε.(v+mε,δ)+ξεdiv(uε)div(mε,δ)

t

0

ρεuε(uε.mε,δ)

γ −1 2 Φε2

div(∇qε,δ)

+ t

0

ρε−√ ρε

uε.(v.v)+ρεuε.p

ρεuεv.v2 2

t

0

mε,δ.vwε,δ+

ρεuεv.v.mε,δ. (23) For all fixedδ, it is easy to see thatAδε(t)converges to 0 for almost allt, uniformly in t whenε goes to 0. Then, by Grönwall’s inequality, we deduce that we have for almost allt(0, T )

wε,δ(t)2L2+Φεψε,δ(t)2L2

wε,δ(0)2L2 +Φεψε,δ(0)2L2+ sup

0st

Aδε(s)eC t

0v(s)2L∞. (24) Then, lettingεgo to 0, we obtain

uv2L(0,T;L2(Ω))lim

ε

wε,δ2L(0,T;L2(Ω))+ Φεψε,δ2L(0,T;L2(Ω))

C0u˜0u0Qu˜0χδ

L2(Ω)+ϕ0ϕ0χδ

L2(Ω)

, where

C0=eC T

0 v(s)2L∞ <+∞.

Then, letting δ go to 0, we deduce that u =v and we obtain also the uniform convergence intofP (

ρεuε)tovinL2, since limε

Pρεuε

vL(0,T;L2)

C0lim

δ u˜0u0Qu˜0χδ

L2(Ω)+ϕ0ϕ0χδ

L2(Ω)

=0.

Moreover, we see that√ρεuεmεconverges uniformly int, strongly inL2tov. In fact

ρεuεmεvL2

ρεuεmε,δvL2+ mεmε,δL2+ Φεψε,δL2,

(9)

and since ε0, m0ε =Q

ρε0u0ε) converges strongly to 0, m0) and since L is an isometry inL2, we deduce that we have

mεmε,δL2+ Φεψε,δL2 →0 whenδ→0 uniformly intandε.

Finally, we can also deduce the local strong convergence of √ρεuε to v in Lp(0, T;L2(B)). Indeed let us denote by B a bounded domain ofRN. Then, we have for allt

ρεuεvL2(B)

ρεuεmε,δvL2(B)+ mε,δL2(B)

ρεuεmε,δvL2(B)+ mε,δLq(B)

for anyq >2. Then, using the fact that for allδ,mε,δLp(0,T;Lq(B))converges to 0 asε goes to 0, we conclude easily by taking the limit inεand then inδas above.

3. The periodic case

As in the case of the whole space, we can deduce, using the energy bounds, that ρε−1 converges strongly to 0 in L(0, T;Lγ). Then, using the bound on

ρεuε in L(0, T;L2), we may extract a subsequence which converges weakly to some u. To pass to the limit in the equation, we need to describe the oscillations in time and show that they will not affect the limit equation.

We next introduce, as in the whole space case, the following group (L(τ ), τ ∈R) defined by eτ L where L is the operator defined on D0 ×(D)N, where D0 = {ϕD,ϕ=0}, by

L ϕ

v

= − divv

ϕ

. (25)

In the sequel, we will use the following notations

Uε=ϕε, Q(ρεuε) and Vε=L(t/ε)ϕε, Q(ρεuε),

and for some technical reasons related to the L2 integrability, we will also use the following approximations

U¯ε=Φε, Q(

ρεuε) and V¯ε=L(t/ε)Φε, Q(ρεuε), which satisfy

Uε− ¯UεL(L2γ /(γ+1))→0 whenε→0.

Let us project Eq. (4) on “gradient vector-fields”

∂tQ(ρεuε)+Qdiv(ρεuεuε)ε+ξε)∇divuε

(26) + a

ε2ρεγγρε+ −1)+ 1

ε2ε−1)=0.

(10)

Combining (26) with the conservation of mass, we obtain ε∂ϕε

∂t +divQ(ρεuε)=0, ε

∂tQ(ρεuε)+ ∇ϕε=εFε, (27) where

Fε=ε+ξε)∇divuεQdiv(ρεuεuε)a1

ε2

ρεγγρε+ −1).

Eq. (27) yields that tUε = 1εLUε +(0, Fε), which can be rewritten as tVε = L(−t/ε)(0, Fε). It is easy to check that Fε is bounded in L2(Hs) (for some s ∈R), hence Vε is compact in time (the oscillations have been cancelled). If we had enough compactness in space we could pass to the limit in this equation and recover the following limit system for the oscillating part

tV¯ +Q1(u,V )¯ +Q2(V ,¯ V )¯ −νV¯ =0, (28) where Q1 and Q2 are respectively a linear and a bilinear forms in V¯ (which will be defined and computed later on) and the term−νV¯ explained below. In fact, as in [17]

(see also [16]), we consider L(−t/ε)(0, Fε)as an almost periodic function in τ =t/ε and compute its mean value, which yields (28).

DEFINITION 3.1. – For all divergence-free vector field uL2(Ω)N and all V = (ψ,q)L2(Ω)N+1, we define the following linear and bilinear symmetric forms in V

Q1(u, V )= lim

τ→∞

1 τ τ

0

L(s)

0

div(u⊗L2(s)V +L2(s)Vu)

ds, (29) and

Q2(V , V )= lim

τ→∞

1 τ

τ 0

L(s)

0

div(L2(s)VL2(s)V )+γ21(L1(s)V )2

ds. (30)

The convergences stated above takes place in W1,1 and can be shown by using almost-periodic functions (see [16] and the references therein). Indeed the functions inside the integral in (29) and (30) are almost periodic inW1,1andQ1(u, V ),Q2(V , V ) are their mean value. We will come back to this issue in the next section. We also remark that in Eq. (28) the viscosity term was divided by 2 and that it applies for both component of the vector V¯. This is due to the following proposition, which can be proved easily using almost periodic functions (see also [5] and [3]).

PROPOSITION 3.2. – Under the same hypothesis onV, we have

νV = lim

τ→∞

1 τ

τ 0

−L(s)

0 2νL2(s)V

ds. (31)

(11)

Nevertheless, the fact that the viscosity applies now for both components ofV is not sufficient to yield compactness in space for Vε. To recover compactness in space, we will use the regularity of the limit system and extend the method used in [12] to the case of general initial data as was done in [16]. LetV0be the solution of the following system

tV0+Q1(v, V0)+Q2(V0, V0)νV0=0,

V|0t=0=0, Qu˜0), (32)

wherev is, as in the case of the whole space, the solution of the incompressible Euler equations with initial datau0. The existence of global strong solutions for the system (32) (and local solutions if the viscosity term is removed) as well as the exact computations of the two formsQ1andQ2will be detailed in the next section. We only need the following two propositions.

PROPOSITION 3.3. – For all u, V, V1 and V2 (regular enough to define all the products), we have

Q1(u, V )V =0 and

Q2(V , V )V =0, (33)

Q1(u, V1)V2+Q1(u, V2)V1=0, (34)

Q2(V1, V1)V2+2Q2(V1, V2)V1=0. (35) The proof of (33) will be given in the next section, (34) can be shown by applying the first part of (33) toV1+V2and toV1V2. Finally (35) can be shown by applying the second part of (33) toV1+X V2and identifying the term of degree 1.

The next proposition is a very simple consequence of the theory of almost-periodic functions (see for instance Lemma 2.3 of [16]).

PROPOSITION 3.4. – For all uLp(0, T;L2) and VLq(0, T;L2), we have the following weak convergences (pandqare such that the product are well defined )

w- lim

ε L

t ε

0 divuL2

t

ε

V +L2

t

ε

Vu

=Q1(u, V ) (36) and

w- lim

ε L

t ε

0 divL2

t

ε

VL2

t

ε

V+γ21L1

t

ε

V2

=Q2(V , V ). (37) Using the symmetry ofQ2, we deduce easily the following proposition.

PROPOSITION 3.5. – Eq.(37)of Proposition 3.4 can be extended to the case where we takeV1andV2using the symmetry ofQ2, namely

(12)

w- lim

ε L

t ε

0 divL2

t

ε

V1L2

t

ε

V2+L2

t

ε

V2L2

t

ε

V1

+L

t ε

0

γ1 2L1

t

ε

V1L1

t

ε

V2

=Q2(V1, V2). (38) Moreover, the above identity holds for V1Lq(0, T;Hs)andV2Lp(0, T;Hs)with s∈Rand 1/p+1/q=1. It is also possible to extend it to the case where we replace V2 in the left hand side by a sequence V2ε such that V2ε converges strongly to V2 in Lp(0, T;Hs).

In order to show the convergence in Theorem 1.2, we will try to estimate

ρεuεvL2

t ε

V0

2

L2

+ΦεL1

t ε

V0

2

L2

.

We also introduce the following notations wε = √ρεuεvL2(t/ε)V0 and βε = ΦεL1(t/ε)V0. In the sequel, we also note ε, mε)=L(t/ε)V0. The proof follows the same lines as the proof in the whole space case apart from the fact that the equation satisfied by the gradient part is not trivial and that we have to use the precise equation satisfied by the oscillating terms. We recall the energy inequality

1 2

ρε|uε|2(t)+Φε2(t)+ t 0

µε|Duε|2+ξε(divuε)21 2

ρε0u0ε2+Φε2(0) (39) as well as the conservation of energy forv

1

2|v|2(t)=

1

2|u0|2. (40)

Using that

Q1

u, V0V0=0,

Q2

V0, V0V0=0, we deduce from (32) the following energy identity

1

2V02(t)+ν t

0

V02(s)ds=

1

2V0(t=0)2. (41) Next, using the weak formulation of the conservation of mass, we obtain for almost allt

ψεϕε(t)+1 ε t

0

div(ρεuεε+div(∇qεε

Références

Documents relatifs

Starting from isentropic compressible Navier-Stokes equations with growth term in the con- tinuity equation, we rigorously justify that performing an incompressible limit one arrives

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

[17] A. Kazhikhov: Solvability of the initial-boundary value problem for the equations of the motion of an inhomogeneous viscous incompressible fluid, Dokl. Kotschote:

Here we prove the existence of global in time regular solutions to the two- dimensional compressible Navier-Stokes equations supplemented with arbitrary large initial velocity v 0

The mild formulation together with the local Leray energy inequality has been as well a key tool for extending Leray’s theory of weak solutions in L 2 to the setting of weak

Marius Mitrea, Sylvie Monniaux. The regularity of the Stokes operator and the Fujita–Kato approach to the Navier–Stokes initial value problem in Lipschitz domains.. approach to

This decomposition is used to prove that the eigenvalues of the Navier–Stokes operator in the inviscid limit converge precisely to the eigenvalues of the Euler operator beyond

Under a localization assumption for the data, we show that if (8) is satisfied, and if this condition remains true in a small interval of time, then the corresponding solution