• Aucun résultat trouvé

Convergence of a low order non-local Navier-Stokes-Korteweg system: the order-parameter model

N/A
N/A
Protected

Academic year: 2021

Partager "Convergence of a low order non-local Navier-Stokes-Korteweg system: the order-parameter model"

Copied!
43
0
0

Texte intégral

(1)

HAL Id: hal-00787268

https://hal.archives-ouvertes.fr/hal-00787268

Submitted on 11 Feb 2013

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

Convergence of a low order non-local

Navier-Stokes-Korteweg system: the order-parameter model

Frédéric Charve

To cite this version:

Frédéric Charve. Convergence of a low order non-local Navier-Stokes-Korteweg system: the order- parameter model. Asymptotic Analysis, IOS Press, 2016, 100 (3-4), pp.153-191. �hal-00787268�

(2)

Convergence of a low order non-local

Navier-Stokes-Korteweg system: the order-parameter model

Fr´ed´eric Charve

Abstract

In the present article we consider a capillary compressible system introduced by C. Rohde after works of Bandon, Lin and Rogers, called the order-parameter model, and whose aim is to reduce the numerical difficulties that one encounters in the case of the classical local Korteweg system (involving derivatives of order three) or the non-local system (also introduced by Rohde after works of Van der Waals, and which involves a convolution operator). We prove that this system has a unique global solution for initial data close to an equilibrium and we precisely study the convergence of this solution towards the local Korteweg model.

1 Introduction

1.1 Presentation of the systems

In the mathematical study of liquid-vapour mixture, Gibbs first modelled phase transi- tions thanks to the minimization of an energy functional with a nonconvex energy density (see [17]). The phases are separated by an hypersurface and there are mainly two ways to describe it: either we consider that the interface behaves like a discontinuity for the fluid parameters (this is the Sharp Interface model), either we consider that between the phases lies a thin region of continuous transition (this is the Diffuse Interface approach, where the phase changes are seen through the variations of the density and which is much simpler numerically). Unfortunately the basic models provide an infinite number of solu- tions (few of them being physically relevant) and this is why authors tried to penalize the high variations of the density (with capillary terms related to surface tension) in order to select the physically correct solutions.

In the present paper, we are interested in the local and non-local Korteweg systems (in the diffuse interface model). These systems are based upon the compressible Navier- Stokes system with a Van der Waals state law for ideal fluids, and endowed with a capillary tensor.

Let us recall that the local model was introduced by Korteweg and renewed by Dunn and Serrin (see [16]) and the non-local model was introduced by Van der Waals and renewed by F. Coquel, D. Diehl, C. Merkle and C. Rohde. For an in-depth presentation of the capillary models, we refer to [30] and [11]).

Universit´e Paris-Est Cr´eteil, Laboratoire d’Analyse et de Math´ematiques Appliqu´ees (UMR 8050), 61 Avenue du G´en´eral de Gaulle, 94 010 Cr´eteil Cedex (France). E-mail: frederic.charve@u-pec.fr

(3)

Letρ and u denote the density and the velocity of a compressible viscous fluid (ρ is a non-negative function and u is a vector-valued function defined onRd). We denote by A the following diffusion operator

Au=µ∆u+ (λ+µ)∇divu, with µ >0 and ν =λ+ 2µ >0.

The Navier-Stokes equations for compressible fluids endowed with internal capillarity

read: (

tρ+ div (ρu) = 0,

t(ρu) + div (ρu⊗u)− Au+∇(P(ρ)) =κρ∇D[ρ].

The capillary coefficient κ may depend on ρ but in this article it is chosen constant. In the local Korteweg system (N SK), the capillary termD[ρ] is given by (see [16]):

D[ρ] = ∆ρ,

and, in the non-local Korteweg system (N SRW) (see [29], [11], and [34]), if φ is an interaction potential which satisfies the following conditions

(|.|+|.|2)φ(.)∈L1(Rd), Z

Rd

φ(x)dx= 1, φ even, andφ≥0, (1.1) then D[ρ] is the non-local term given by:

D[ρ] =φ∗ρ−ρ.

Comparing the Fourier transform of the capillary terms, we have (φ(ξ)b −1)ρ(ξ) in theb non-local model, and −|ξ|2ρ(ξ) in the local model so that a natural question is to studyb the closedness of the solutions of these models when φ(ξ) is formally ”close” to 1b − |ξ|2. For this, we introduced in [7] a specific interaction potential and considered the following non-local system:

(N SRWε)

(∂tρε+ div (ρεuε) = 0,

tεuε) + div (ρεu⊗uε)− Auε+∇(P(ρε)) =ρεκ

ε2∇(φε∗ρε−ρε), with

φε= 1 εdφ(x

ε) with φ(x) = 1 (2π)de|x|

2 4

For a fixedξthe Fourier transform ofφεisφbε(ξ) =eε2|ξ|2, and whenεis small, φbε(ξ)−1 ε2 is close to −|ξ|2.

Using energy methods, we proved in [7] that this system has a unique global strong solution for initial data close to an equilibrium state. The functional setting are classical and hybrid Besov spaces (taylored to the capillary term). We also obtained that when the small parameter ε goes to zero, the solution tends to the corresponding solution of the local Korteweg system and we obtained a rate of convergence in terms of ε. In [9], we provided by Lagrangian methods more precise a priori estimates giving a better understanding of the convergence and the hybrid Besov setting in terms of the linear Fourier structures.

(4)

Though, these models are not completely satisfying. On one hand, recalling the re- sults from [12], [2] (compressible Navier-Stokes system), [14], [8] (local Korteweg model), and [19], [7] (non-local Korteweg model), we observe that the density in the local cap- illary model is far more regular than in the non-local model where it shares the same frequency structure as in the compressible Navier-Stokes model (heat regularization in low frequencies and only a damping in high frequencies).

On the other hand, from a numerical point of view the local model is difficult to handle because the capillary term contains third-order derivatives. The non-local model also presents difficulties in numerical studies: even if the capillary term only contains derivatives of order one, it involves a convolution operator, whose numerical difficulty is comparable.

For this reason, C.Rohde presented in [31] a new model, called the order-parameter model, and inspired by the work of D. Brandon, T. Lin and R. C. Rogers in [5].

This new system consists in introducing in the capillary term α2∇(c −ρ) a new variableccalled the ”order parameter”, which is coupled to the density via the following relation related to the Euler-Lagrange equation from the variational approach (αcontrols the coupling between ρ and c):

ε2∆c+α2(ρ−c) = 0 so that the new system he considers is the following:

(N SOPα)





tρα+ div (ραuα) = 0,

tαuα) + div (ραuα⊗uα)− Auα+∇(P(ρα)) =κα2ρα∇(cα−ρα), ε2∆cα2α−cα) = 0.

As emphasized by C. Rohde, from a numerical point of view this system is much more interesting because now we only have one derivative in the capillary tensor (which is local), and the additionnal equation for the order parameter is a simple linear elliptic equation that can be easily and numerically fast solved at least when the mesh is fixed.

Moreover as we will see later, as for the previous non-local capillary model, this system has the same frequency structure as the classical Navier-Stokes model.

In [31], C. Rohde proves (forε=λ=µ= 1 in the two-dimensionnal case) that the system has a unique local classical solution:

Theorem 1 ([31]) Assume that the initial data (ρ0, u0) is independant of α >0 with u0 ∈H4(R2), ρ0 > 0 and ρ0−ρ¯∈ H4(R2) for some constant ρ. Let¯ c0 the solution of the elliptic problem −∆c02c02ρ0. There exists a constant T >0 such that the initial-value problem(N SOPα) has a unique solution defined on [0, T[satisfying:

ρα−ρ, u¯ α∈L(0, T;H4(R2)), ρα>0, cα−ρ¯∈L(0, T;H5(R2)).

Moreover, for allt∈[0, T[we have

αlim→∞α(t, .)−cα(t, .)kL2(R2)= 0.

In this paper, C. Rohde also conjectures that when the coupling constant α goes to infinity, these solutions converge to the solution of the local Korteweg model.

(5)

1.2 Statement of the results

In the present article, following what we did in the whole space for the non-local system (see [7]) and using lagrangian methods from [9], we will prove that under smallness conditions, and with less regular initial data, the system has global strong solutions in the following critical spaces (we refer to the appendix for more details on Besov spaces and hybrid spaces). We also prove the above conjectured convergence and give an explicit rate of convergence with respect toα.

Definition 1 The spaceFαs is the set of functions(q, c, u) in

Cb(R+,B˙2,1s1∩B˙2,1s )∩L1(R+,B˙αs+1,s1∩B˙αs+2,s)2

×

Cb(R+,B˙2,1s1)∩L1(R+,B˙2,1s+1)d

endowed with the norm k(q, c, u)kFαs =k(q, c, u)kFαs() where for allt we denote (recall that ν0= min(µ, ν))

k(q, c, u)kFαs(t)

def= kukLet B˙s−12,1 +kqkLet B˙s−12,1 +νkqkLet B˙2,1s +kckLet B˙s−12,1 +νkckLet B˙2,1s

0kukLe1tB˙2,1s+1+νkqkLe1tB˙αs+1,s−12kqkLe1tB˙αs+2,s+νkckLe1tB˙αs+1,s−12kckLe1tB˙αs+2,s (1.2) Theorem 2 Let α > 0 and assume min(µ,2µ+λ) > 0. There exist two positive con- stants ηOP and C only depending ond,µ,λ,κ, and P(ρ) such that for allη ≤ηOP, if ρ0−ρ∈B˙

d 21 2,1 ∩B˙

d

2,12 ,u0∈B˙

d 21

2,1 ,c0 is defined by −∆c02c02ρ0 and kρ0−ρk˙

B

d2−1 2,1 B˙

d2 2,1

+ku0k˙

B

d2−1 2,1

≤η

then system(N SOPα)has a unique global solution (ρα, cα, uα)with(ρα−ρ, cα−ρ, uα)∈ F

d

α2 such that:

k(ρα−ρ, cα−ρ, uα)k

F

d2 α

≤C0 def= C(kρ0−ρk˙

B

d2−1 2,1 B˙

d2 2,1

+ku0k˙

B

d2−1 2,1

).

Moreover we have the global in time results:





kcα−ραke

L(R+,B˙

d 2−1

2,1 )+νkcα−ραke

L(R+,B˙

d 2,12 )α−→

→∞0, νkcα−ραk

L1(R+,B˙

d2−1

2,1 )2kcα−ραk

L1(R+,B˙

d2

2,1) ≤C0α2.

The following result deals with the convergence in α: when the initial data are small enough (so that we have global solutions for (N SK) and (N SOPα)) the solution of (N SOPα) goes to the solution of (N SK) whenα goes to infinity.

Theorem 3 With the same assumptions as before, there exists 0< η0 ≤min(ηK, ηOP) such that for allη ≤η0, if

0−ρk˙

B

d2−1 2,1 B˙

d2 2,1

+ku0k˙

B

d2−1 2,1

≤η,

(6)

then systems(N SK)and(N SOPα)both have global solutions andk(ρα−ρ, cα−ρ, uα− u)k

F

d2 α

goes to zero as α goes to infinity. Moreover, with the same notations as before, there exists a constant C =C(η, κ, ρ, P(1))>0 such that for all h∈]0,1[ (if d= 2) or h∈]0,1](if d≥3)

k(ρα−ρ, cα−ρ, uα−u)k

F

d2−h α

≤Cαh,

Remark 1 We can assume that ε= 1 without loss of generality. If not we just have to replace κ byε2 andα by α/ε.

Remark 2 As the order parameter cα goes toρα, we formally get that when α goes to infinity, the capillary term goes to κρ∇∆ρ.

1.3 outline of the paper

The article is structured the following way: section 2 is devoted to the proof of theorem 2. We first introduce an interaction potential φα that allows us to rewrite the system into a non-local shape. As we want precise estimates we follow the methods from [9]:

we first obtain estimates on the linearized system and then on the advected linear sys- tem thanks to a Lagrangian change of variable. The rest of the proof is classical, we define approximated solutions thanks to the Friedrichs’ scheme and obtain existence and uniqueness like in [7]. In section 3 we prove theorem 3 and in the appendix, we first re- call basic properties of Besov spaces, then we provide estimates for the flow of a smooth vectorfield. The last part of the appendix is devoted to Bessel functions that are needed for the expression of our new interaction potential.

2 Proof of theorem 2

2.1 Interaction potential

As announced in the introduction, we first rewrite the system in a non-local shape. Let us focus on the last equation, we can write that (for more clarity we drop the subscripts with α):

−ε2∆(ρ−c) +α2(ρ−c) =−ε2∆ρ which leads to:

α2(ρ−c) =−α2(−∆ +α2

ε2)1∆ρ=−(−∆ α2 + 1

ε2)1∆ρ So that in Fourier variable:

α2\(ρ−c)(ξ) = |ξ|2

|ξ|2 α2 +ε12

b

ρ(ξ) =ε2·α2

ε2(1− 1

ε2

α2|ξ|2+ 1)ρ(ξ).b

Then up to choose κ=ε2 and replace αby α/ε, from now on we assume thatε= 1 and then if we introduce:

D[ρ] =α2(c−ρ),

(7)

we have

D[ρ](ξ) =d −|ξ|2

|ξ|2

α2 + 1ρ(ξ) =b α2( 1

|ξ|2

α2 + 1−1)ρ(ξ).b (2.3) As a consequence, when α is large, α2(ρ−c) formally goes to ∆ρ as for the non-local capillary term from [9], and the object of this article is to prove that the solutions of this system will go to the solutions of the local Korteweg model. Let us now define the interaction potential φα by:

φcα(ξ) = 1

|ξ|2

α2 + 1. (2.4)

We have R

Rφα(x)dx= 1 and D[ρ] =α2α∗ρ−ρ). If we put φ=φ1 then φ(ξ) =b 1

|ξ|2+ 1, φcα=φ(b·/α), andφαdφ(α·). (2.5) In some cases we have explicit expressions for this inverse Fourier transform: for all x, φ(x) =Ce−|x| when d = 1, φ(x) = Ce−|x||x| when d = 3 (we refer to [33]). In the other cases the expression of φ involves Bessel functions. Let us begin by recalling that the fourier transform of a radial function is also radial, more precisely (see for example [33]

page 213) there exists a constant Cd such that if f(x) = f0(|x|) for all x ∈Rd, then its Fourier transform satisfies for all ξ ∈Rd,fb(ξ) =F0(|ξ|) where for all ρ >0

F0(ρ) = Cd ρd21

Z

0

Jd

21(ρr)f0(r)rd2dr,

whereJν denotes the general Bessel function of real indexν. This formulation is related to the Hankel transform, we refer to the appendix for more details and properties on Bessel functions. Coming back to our problem, we then obtain that for all x∈Rd,

φ(x) = Cd

|x|d21 Z

0

Jd

21(r|x|) rd2

1 +r2dr. (2.6)

And thanks to the Hankel-Nicholson integrals (we refer for example to [27] page 330 or [36] page 434), under the following assumptions:

a >0, Re(z) >0, −1< Re(ν)<2Re(µ) +3 2, we have the identity:

Z

0

tν+1Jν(at)

(t2+z2)µ+1dt= aµzνµ

2µΓ(µ+ 1)Kνµ(az),

whereKν denotes the modified Bessel function of the second kind and indexν(also called Hankel, Schl¨afti or Weber function). This allows us to finally write that for all x ∈Rd provided thatd∈ {1,2,3,4} (from the previous conditions withν = d2−1,µ= 0, z= 1, a=|x|),

φ(x) = Cd

|x|d21Kd

21(|x|). (2.7)

(8)

Remark 3 Another way to understand the limitation on the dimension consists in ob- serving in the integral (2.6), that if we roughly approximate the Bessel function by cos(r)r1/2 at infinity, then the integrated function has the following asymptotic expan- sion at infinity: cos(r)rd/25/2 (|x|= 1 for more simplicity).

In fact (2.7) is also valid for dimensions d≥5 (Like the Fourier transform, the Hankel transform can be generalized for tempered distributions). Let us compute the Fourier transform: for all ξ∈Rd

Z

Rd

eix·ξ Kd

21(|x|)

|x|d21 dx= Z

0

rd2Kd

21(r) Z

Sd−1

eirω·ξ

dr, and thanks to the radial symmetry:

Z

Sd−1

eirω·ξdω= Z

Sd−1

eir|ξ|ω·e1dω.

Performing a d-dimensional spherical change of variable we obtain that (with λ=r|ξ|):

Z

Sd−1

eiλω·e1dω= Z π

0

Z

0

...

Z

0

ecosθ1sind2θ1sind3θ2...sinθd21...dθd1

=Cd Z π

0

ecosθ1sind2θ11. (2.8) Using the following integral representation of function Iν (for Re(ν) > −1/2 see the appendix for modified Bessel functions Iν and Kν)

Iν(z) = zν 2νπ12Γ(ν+12)

Z π

0

ezcostsind2tdt.

Then thanks to the following identity (here a=i|ξ|and b= 1):

Z

zIν(az)Kν(bz)dz= z

a2−b2 (aIν+1(az)Kν(bz) +bIν(az)Kν+1(bz)), we obtain that:

Z

Rd

eix·ξ Kd

21(|x|)

|x|d21 dx= 2d21Γ(d2) 1 +|ξ|2 .

so that in (2.7), Cd= (2d21Γ(d2))1, Considering the asymptotics of functionKd

21,φis continuous on Rs− {0}. Near 0, and for d≥3, we have φ(x)∼Cd|x|2d so that|x|φ(x) is aL1 function on Rd.

2.2 Reformulation of the system

We are now able to write the system into a non-local form:

(N SOP(α)

tρα+ div (ραuα) = 0,

tαuα) + div (ραu⊗uα)− Auα+∇(P(ρα)) =ρακα2∇(φα∗ρα−ρα),

(9)

with

φαdφ(α·) with φ(x) = Cd

|x|d21Kd

21(|x|).

Remark 4 From the previous computations, we immediately get that cα−ρα = (−∆ +αId)1∆ραα∗ρα−ρα

that iscαα∗ρα. This is why we cannot choose any initial data for the order parameter and take c0α∗ρ0.

As we consider initial data close to an equilibrium state (ρ,0) we begin with the classical change of function ρ = ρ(1 +q). For simplicity we take ρ = 1. The previous system becomes (also denoted by (N SOPα)):

(N SOPα)

(∂tqα+uα.∇qα+ (1 +qα)divuα= 0,

tuα+uα.∇uα− Auα+P(1).∇qα−κα2∇(φα∗qα−qα) =K(qα).∇qα−I(qα)Auα, where K and I are the real-valued functions defined onRgiven by:

K(q) =

P(1)−P(1 +q) 1 +q

and I(q) = q q+ 1. The functional spaces we will really use are the following:

Definition 2 The spaceEαs is the set of functions(q, u) in

Cb(R+,B˙2,1s1∩B˙2,1s )∩L1(R+,B˙s+1,sα 1∩B˙αs+2,s)

×

Cb(R+,B˙2,1s1)∩L1(R+,B˙2,1s+1)d

endowed with the normk(q, u)kEαs =k(q, u)kEαs() where for all twe denote (recall that ν0 = min(µ, ν))

k(q, u)kEαs(t)

def= kukLet B˙s−12,1 +kqkLet B˙2,1s−1 +νkqkLet B˙2,1s

0kukLe1tB˙2,1s+1+νkqkLe1tB˙αs+1,s−12kqkLe1tB˙αs+2,s (2.9) Remark 5 Due to obvious simplifications we slightly changed the notations for Eαs and B˙αs+2,s: with the notations from [9] these spaces would have been respectively denoted by E1/αs and B˙1/αs+2,s.

We will now follow the tracks of [7] and [9] to prove the results. Classically in the study in critical spaces of compressible Navier-Stokes-type systems (see [12, 6, 18]), the proofs of theorems 2 and 3 (see [7] section 2) rely on key a priori estimates on the following advected linear system (α >0 is fixed and for more simplicity we write (q, u) instead of (qα, uα)):

(LOPα)

(∂tq+v.∇q+ divu=F,

tu+v.∇u− Au+p∇q−κα2∇(φα∗q−q) =G.

(10)

With

Au=µ∆u+ (λ+µ)∇divu.

Although the potential function is different from the gaussian from [7], we can easily adapt the energy methods and results from this paper. Here we will directly focus on more refined estimates as in [9] and use them in the proof of the last theorem: we can prove that the estimates are similar up to slight changes in the constants:

Theorem 4 Letα >0,−d2+ 1< s < d2+ 1,I = [0, T[or[0,+∞[andv∈L1(I,B˙

d 2+1 2,1 )∩ L2(I,B˙

d

2,12 ). Assume that(q, u)is a solution of System(LOPα) defined onI. There exists α0 >0, a constant C >0 depending on d, ssuch that if α ≥α0, for all t∈I (denoting ν =µ+ 2λ andν0 = min(ν, µ)),

kukLet B˙s−12,1 +kqkLet B˙2,1s−1+νkqkLet B˙2,1s0kukLe1tB˙2,1s+1+νkqkLe1tB˙s+1,s−1α2kqkLe1tB˙s+2,sα

≤Cp,ν2

e

Cp,ν2Cvisc Z t

0

(k∇v(τ)k˙

B

d2 2,1

+kv(τ)k2

B˙

d 2,12

)dτ

×

ku0kB˙2,1s−1 +kq0kB˙s−12,1 +νkq0kB˙s2,1+kFkLe1tB˙2,1s−1+νkFkLe1tB˙s2,1 +kGkLe1tB˙2,1s−1

. (2.10)

where 





Cp,ν2 =Cmax(√p, 1

√p) max(4κ ν2,(ν2

4κ)2), Cvisc = 1 +|λ+µ|+µ+ν

ν0 + max(1, 1

ν3).

Remark 6 The coefficient Cvisc satisfies:

Cvisc= (1+2ν

µ + max(1,ν13) Ifλ+µ >0,

1+2µ

ν + max(1,ν13) Ifλ+µ≤0 and when both viscosities are small, we simply haveCvisc ≤max(1,ν13

0).

2.3 Linear estimates

As in [6] and [9] the first step to prove theorem 4 is to obtain estimates for the following linearized system:

(OPα)

(∂tq+ divu=F,

tu− Au+p∇q−κα2∇(φα∗q−q) =G.

With

Au=µ∆u+ (λ+µ)∇divu.

In this article, we will use the following frequency-localized estimate:

(11)

Proposition 1 Letα >0,s∈R,I = [0, T[or [0,+∞[. Assume that(q, u)is a solution of System (Oε) defined on I. There exists α0 >0, a constant C >0depending on d,s, c0 and C0 such that if α ≥α0, for all t∈ I (we recall that ν0 and C

p,ν2 are defined in the previous theorem), and for all j∈Z,

k∆˙jukLt L2022jk∆˙jukL1tL2 + (1 +ν2j)

k∆˙jqkLt L2+νmin(α2,22j)k∆˙jqkL1tL2

≤C

p,ν2

(1 +ν2j)k∆˙jq0kL2+k∆˙ju0kL2 + (1 +ν2j)k∆˙jFkL1tL2 +k∆˙jGkL1tL2

(2.11)

Remark 7 Let us precise that these linear estimates are valid for all dimension, the limitation d≤4 only appears in the advected case.

2.3.1 Eigenvalues and eigenvectors

In this article, since the methods are very close to [7] and [9] we will only point out what is different and refer to these articles for details. As in [12] or [6] we first introduce the Helmholtz decomposition of u. Defining the pseudo-differential operator Λ by Λf = F1(|.|fb(.)), we set: (

v= Λ1divu,

w= Λ1curlu (2.12)

thenu=−Λ1∇v+ Λ1divw and the system turns into:

(Lα)





tq+ Λv=F,

tv−ν∆v−pΛq+κα2Λ(φα∗q−q) = Λ1divG,

tw−µ∆w= Λ1curlG.

The last equation is a decoupled heat equation, easily estimated in Besov spaces (see [2] chapter 2). Moreover, as the external forces appear through homogeneous pseudo- differential operators of degree zero, we can compute the estimates in the caseF =G= 0 and deduce the general case from the Duhamel formula. So we can focus on the first two lines and compute the eigenvalues and eigenvectors of the matrix associated to the Fourier transform of the system:

t qb

b v

=A(ξ) qb

b v

with A(ξ) :=

 0 −|ξ|

|ξ|(p+κ|ξ||ξ2|2

α2 +1) −ν|ξ|2

.

The discriminant of the characteristic polynomial of A(ξ) is:

∆(ξ) =|ξ|2 ν2|ξ|2−4(p+κ |ξ|2

|ξ|2 α2 + 1)

! , and thanks to the variations of function

fα :x7→ν2x−4(p+κ x

x

α2 + 1) =ν2x−4(p+κα2) + 4κα2

x α2 + 1,

(12)

we obtain the existence of a unique threshold xα>0 such that

∆(ξ)

(<0 if |ξ|2 < xα,

>0 if |ξ|2 > xα.

We emphasize that this function has the same variations as in the case of [9]: when

ν2

4K ≥ 1, fα is an increasing function on R+, and when 4Kν2 < 1, fα is decreasing in [0, α2(2νK −1)] and then increasing.

Proposition 2 Under the same assumptions, we have:

xα α

→∞















 4p

ν2−4κ if ν2 >1, α

rp

κ if ν2 = 1, (4κ

ν2 −1)α2 if ν2 <1.

Proof: It is simpler than in [9] because here we have explicit expressions for the threshold: if we put A=ν2−κ−4p/α2, then

xα= α2

2 −A+ r

16pν2 α2 +A

! .

Next introducing the following function, we obtain the expressions of bq and bv exactly as in [9]:

gα(x) = fα(x)

ν2x = 1− 4

ν2x(p+κ x

x

α2 + 1) (2.13)

-For the low frequencies (∆<0), when|ξ|<√xα, we have:







 b

q(ξ) = 12

(1 +S(ξ)i )e+ + (1− S(ξ)i )e b

q0(ξ)−ieν|+ξ|S(ξ)evb0(ξ), b

v(ξ) =i

p+κ|ξ||ξ2|2 α2+1

e+e

ν|ξ|S(ξ) qb0(ξ) + 12

(1−S(ξ)i )e++ (1 + S(ξ)i )e b v0(ξ), with:

S(ξ) =p

−gα|ξ|2) = s 4

ν2|ξ|2(p+κ |ξ|2

|ξ|2

α2 + 1)−1 (2.14) and

λ±=−ν|ξ|2

2 (1±iS(ξ)).

-For the high frequencies (∆>0), when |ξ|>√xα, we have:







 b

q(ξ) = 12

(1− R(ξ)1 )e++ (1 + R(ξ)1 )e b

q0(ξ) +eν|+ξ|R(ξ)evb0(ξ), b

v(ξ) =−

p+κ|ξ||ξ2|2 α2+1

e+e

ν|ξ|R(ξ) qb0(ξ) +12

(1 +R(ξ)1 )e+ + (1− R(ξ)1 )e b v0(ξ),

(13)

with:

R(ξ) =p

gα(|ξ|2) = s

1− 4

ν2|ξ|2(p+κ |ξ|2

|ξ|2

α2 + 1) (2.15)

and

λ± =−ν|ξ|2

2 (1±R(ξ)).

Remark 8 As in [9] it is crucial for the time integration to observe that p+κ |ξ|2

|ξ|2

α2 + 1 = ν2|ξ|2

4 (1−R(ξ))(1 +R(ξ)).

2.3.2 Thresholds

As in [9] we can find another threshold frequency yα > xα of size α2 (in each case for

ν2

) that will enable us to push the parabolic regularization until frequencies of size α.

In the present paper we will have explicit expressions. Let us first remark that rewriting functiongαinto the following form immediately implies that this is an increasing function from [0,∞[ to ]− ∞,1[:

gα(x) = 1−4p ν2

1 x −α2

M 1

x+α2 with M def= ν2

4κ (2.16)

If β ∈ [0,1[ we easily compute that there is a unique positive solution of the equation gα(x) =β given by:

xα,β= 1

2 −A+

s 16p

(1−β)ν2α2+A2

!

where A= 1

2(M− 1

1−β)− 1 1−β

4p ν2. so that we immediately have the following asymptotics whenα is large:

xα,β

















1

M(1−β) −1

α2 if M < 11β, 2

ν r p

1−βα if M = 11β, 4p

ν2

M

(1−β)M−1 if M > 11β.

Remark 9 The previous proposition is obviously a particular case of this result.

We are now able to define the second threshold yα:

gα(yα) =



 1

2 if M ≤1, 1− 1

2M ≥ 1

2 if M ≥1,

where M = ν2

4κ. (2.17)

Using the previous result for β = 12 or 1− 2M1 according to the case forM, we obtain

(14)

Proposition 3 With the same notations we have that:

• IfM ≥1,yα α

→∞α2 and for allα,

α2 ≤yα≤2α2.

• IfM ≤1,yα α

→∞(M2 −1)α2 and for allα, α2 ≤( 2

M −1)α2 ≤yα≤( 2 M −1

2)α2. 2.3.3 Pointwise estimates

Now that we have defined the frequency thresholds xα and yα we have the following estimates. Up to the values of m and the second exponential from the density in the second case, they are the same as in [9] to where we refer for details or proofs:

Proposition 4 Under the previous notations, there exists a constant C, such that for all j ∈ Z and all ξ ∈ 2jC where C is the annulus {ξ ∈ Rd, c0 = 34 ≤ |ξ| ≤ C0 = 83}, we have the following estimates (we denote by fj = ˙∆jf and we refer to the appendix for details on the Littlewood-Paley theory):

• If|ξ|<√xα:





(1 +ν2j)|qbj(ξ)| ≤Ceνtc

2022j 4

(1 +ν2j)|dq0,j(ξ)|+ (1 + 1p)|dv0,j(ξ)| ,

|vbj(ξ)| ≤Ce

νtc2 022j

4 (1 +ν2j)(1 +√p)(1 + ν2)|dq0,j(ξ)|+|dv0,j(ξ)| .

• If√xα<|ξ|<√yα:





(1 +ν2j)|qbj(ξ)| ≤ 1Cmeνtc

2022j

4 (1m)

(1 +ν2j)|dq0,j(ξ)|+ (1 +1p)|dv0,j(ξ)| ,

|vbj(ξ)| ≤ 1Cmeνtc

2022j

4 (1m) ν2j|dq0,j(ξ)|+|dv0,j(ξ)| , wherem=p

gα(yα) = 1

2 if M = ν2 ≤1,m=q

1− 2M1 ifM ≥1.

• If|ξ|>√yα>√xα:









(1 +ν2j)|qbj(ξ)| ≤C

eνt|ξ|

2

2 +eκα2t (1 +ν2j)|dq0,j(ξ)|+ (1 + 1p)|dv0,j(ξ)| ,

|vbj(ξ)| ≤C

eνtc

2022j

4 + 1−p

gε(c2022j) e

νtc2 022j

2 1

gε(C2022j)

ν2j|dq0,j(ξ)|+|dv0,j(ξ)| .

(15)

2.3.4 Time estimates

As in [9], due to the choice c0 = 3/4 and C0 = 8/3 (see the appendix), we can observe that there exist at most two indices j

α =jα−1 or j

α =jα such that √yα ∈2j[c0, C0] forj ∈ {j

α, jα}.

We refer to [9] for the proof of the following proposition that implies Proposition 1.

Proposition 5 Under the same assumptions as in Proposition 1, there exists a constant C such that for allj∈Z (denotingM = ν2):

• For allj ≤jα,

kvjkLt L2 +ν22jkvjkL1tL2+ (1 +ν2j)

kqjkLt L2 +ν22jkqjkL1tL2

≤ Cmax( 1

M, M2)

(1 +ν2j)(1 +√p)kq0,jkL2 + (1 + 1

√p)kv0,jkL2

,

(2.18)

• For allj > jα,

kvjkLt L2 +ν22jkvjkL1tL2+ (1 +ν2j)

kqjkLt L2 + ν

ε2kqjkL1tL2

≤ Cmax(1, M)

(1 +ν2j)kq0,jkL2 + (1 + 1

√p)kv0,jkL2

.

(2.19)

2.4 Advected linear estimates

The difficulties and methods exposed here are the same as in [9], so we will roughly explain them and focus on what is new.

In order to prove Theorem 4, a natural idea is to use Proposition 1 and put the advection terms as external forces. Unfortunately, there are some obstacles: the main problem is that inv·∇q, the term ˙Tv∇qcan be estimated in ˙B2,1s1but not in ˙B2,1s because it is not enough regular in high frequencies.

A direct use of the linear estimates will be useful only for the low frequencies (j≤0), and in the high frequency regime (j > 0), we will perform a Langrangian change of variable (as in [22], [23], [15], [6], [9]) in order to get rid of v· ∇q. We then aim to use on the new system our linear estimates but we have to be careful with the external force terms introduced by the change of variable. Most of the work in [9] was to provide estimates on the commutator of the non-local operator from the capillarity term and the Lagrangian change of variable.

For all j∈Z and t∈I, we introduce:

Uj(t) =k∆˙jukLt L2022jk∆˙jukL1tL2+(1+ν2j)

k∆˙jqkLt L2 +νmin(1

ε2,22j)k∆˙jqkL1tL2

(2.20)

Références

Documents relatifs

The solution u(t) of the integral equation (III) constructed in Theorem 1 or 2 may be regarded as a solution of the operator differential equation (II) in a

First introduced by Faddeev and Kashaev [7, 9], the quantum dilogarithm G b (x) and its variants S b (x) and g b (x) play a crucial role in the study of positive representations

10, the resulting hydrodynamics (later on referred to as hydrodynamic model for self-alignment with precession) consists of a conservation law for the density and a

(9) Next, we will set up an aggregation-diffusion splitting scheme with the linear transport approximation as in [6] and provide the error estimate of this method.... As a future

Another result concerning the local exact controllability of the Navier-Stokes system with Navier slip boundary conditions was studied in [12].. Let us define the concepts

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

Finally, in our previous work [3] we proved the existence and uniqueness of strong solutions in the case of an undamped beam equation but for small initial deformations.. The outline

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