• Aucun résultat trouvé

Stabilization of a rigid body moving in a compressible viscous fluid

N/A
N/A
Protected

Academic year: 2021

Partager "Stabilization of a rigid body moving in a compressible viscous fluid"

Copied!
28
0
0

Texte intégral

(1)

HAL Id: hal-02314870

https://hal.archives-ouvertes.fr/hal-02314870

Preprint submitted on 14 Oct 2019

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.

viscous fluid

Arnab Roy, Takéo Takahashi

To cite this version:

Arnab Roy, Takéo Takahashi. Stabilization of a rigid body moving in a compressible viscous fluid.

2019. �hal-02314870�

(2)

ARNAB ROY AND TAK ´ EO TAKAHASHI

Universit´ e de Lorraine, CNRS, Inria, IECL, F-54000 Nancy, France

Abstract.

We consider the stabilizability of a fluid-structure interaction system where the fluid is viscous and compressible and the structure is a rigid ball. The feedback control of the system acts on the ball and corresponds to a force that would be produced by a spring and a damper connecting the center of the ball to a fixed point

h1

. We prove the global-in-time existence of strong solutions for the corresponding system under a smallness condition on the initial velocities and on the distance between the initial position of the center of the ball and

h1

. Then, we show with our feedback law, that the fluid and the structure velocities go to 0 and that the center of the ball goes to

h1

as

t→ ∞.

Keywords. Fluid-structure interaction, compressible Navier-Stokes system, global solutions, stabili- tization.

AMS subject classifications. 35Q35, 35D30, 35D35, 35R37, 76N10, 93D15, 93D20.

Contents

1. Introduction and main result 1

Notation 5

2. Local in time existence of solutions 6

2.1. Lagrangian change of variables 6

2.2. Analysis of a linear problem 8

2.3. Estimates of the nonlinear terms 10

2.4. Proof of Theorem 2.1 14

3. Global in time existence of solutions 15

3.1. A priori estimates 15

3.2. Proof of Theorem 1.1 22

4. Proof of Theorem 1.2 24

References 27

1. Introduction and main result

Let Ω ⊂ R

3

be a bounded domain with C

4

boundary occupied by a fluid and a rigid body. We denote by B(t) ⊂ Ω, the domain of the rigid body and we assume it is an open ball of radius 1 and of center h(t), where t ∈ R

+

is the time variable. We suppose that the fluid domain F (t) = Ω \ B(t) is connected.

Date: October 13, 2019.

1

(3)

The fluid is modeled by the compressible Navier-Stokes system whereas the motion of the rigid body is governed by the balance equations for linear and angular momentum. We also assume the no-slip boundary conditions. The equations of motion of fluid-structure are:

∂ρ

∂t + div(ρu) = 0 t > 0, x ∈ F (t), (1.1) {continuity}

ρ ∂u

∂t + (u · ∇) u

− div σ(u, p) = 0 t > 0, x ∈ F (t), (1.2) {momentum}

m`

0

= − Z

∂B(t)

σ(u, p)N dΓ + w t > 0, (1.3) {linear:body}

J ω

0

= − Z

∂B(t)

(x − h(t)) × σ(u, p)N dΓ t > 0, (1.4) {angular:body}

h

0

= ` t > 0, (1.5)

u(t, x) = 0 t > 0, x ∈ ∂Ω, (1.6) {boundary f}

u(t, x) = `(t) + ω(t) × (x − h(t)) t > 0, x ∈ ∂B(t), (1.7) {boundary fs}

ρ(0, ·) = ρ

0

, u(0, ·) = u

0

in F(0), (1.8)

h(0) = h

0

, `(0) = `

0

, ω(0) = ω

0

. (1.9) {initial}

In the above equations, ρ = ρ(t, x) and u = u(t, x) represent respectively the density and the ve- locity of the fluid and the pressure of the fluid is denoted by p. We assume that the flow is in the barotropic regime and we focus on the isentropic case where the relation between p and ρ is given by the constitutive law:

p = aρ

γ

,

with a > 0 and the adiabatic constant γ >

32

. The Cauchy stress tensor is defined as:

σ(u, p) = 2µ D (u) + λ div u I

3

− p I

3

, where D(u) =

12

∇u + ∇u

>

denotes the symmetric part of the velocity gradient (∇u

>

is the transpose of the matrix ∇u) and λ, µ are the viscosity coefficients satisfying

µ > 0, λ + µ > 0.

Here ` and ω are the linear and angular velocities of the rigid body, N (t, x) is the unit normal to ∂B(t) at the point x ∈ ∂B(t), directed to the interior of the ball and m, J are the mass and the moment of inertia of the rigid ball respectively. The formulae for m and J are

m = 4

3 πρ

B

, J = 2m 5 I

3

, where ρ

B

> 0 is the constant density of the rigid ball.

Finally, w (in (1.3)) is our control that we take as a feedback control:

w(t) = k

p

(t)(h

1

− h(t)) − k

d

`(t), (1.10) {feedback}

where k

d

> 0 and k

p

(t) > 0 are well-chosen so that

t→∞

lim h(t) = h

1

,

(4)

whereas the velocities of the fluid and of the rigid ball go to 0:

t→∞

lim u(t) = 0, lim

t→∞

`(t) = 0, lim

t→∞

ω(t) = 0.

In literature, this type of control is known as Proportional-Derivative (PD) controller generated by a spring and a damper. The spring-damper is connected from the center of the ball to the fixed anchor point h

1

and it is attracting the ball towards the point h

1

.

In order to give the precise statement of stabilization (Theorem 1.2), we first need a global in time existence result for (1.1)–(1.10) with (1.10). Such a result in the case without control is given in [1]

by adapting a method introduced in [13].

Here we will prove again this existence result, with the same approach but with a special attention to the estimates on h(t) and with some modifications in the proof of [1] due to the feedback law (1.10).

In order to state our result we introduce ρ the mean-value of ρ

0

: ρ = 1

|F (0)|

Z

F(0)

ρ

0

(x) dx. (1.11) {mean value}

Note that, from equation (1.1) and Reynold’s Transport Theorem, we obtain Z

F(0)

ρ

0

(x) dx = Z

F(t)

ρ(t, x) dx.

For 0 6 T

1

< T

2

6 ∞, we introduce the following space:

S b

T1,T2

= n

(ρ, u, `, ω) | ρ ∈ L

2

(T

1

, T

2

; H

3

(F(t))) ∩ BC

0

([T

1

, T

2

]; H

3

(F (t))) ∩ H

1

(T

1

, T

2

; H

2

(F(t))

∩ BC

1

([T

1

, T

2

]; H

2

(F(t))) ∩ H

2

(T

1

, T

2

; L

2

(F(t))),

u ∈ L

2

(T

1

, T

2

; H

4

(F (t))) ∩ BC

0

([T

1

, T

2

]; H

3

(F(t))) ∩ H

1

(T

1

, T

2

; H

2

(F(t))

∩BC

1

([T

1

, T

2

]; H

1

(F(t))) ∩ H

2

(T

1

, T

2

; L

2

(F(t))),

` ∈ H

2

(T

1

, T

2

), ω ∈ H

2

(T

1

, T

2

) o

.

(1.12) {solution space}

Here BC

k

are the functions of class C

k

bounded with bounded derivatives. We set k(ρ, u, `, ω)k

SbT1,T2

= kρ − ρk

L(T1,T2;H3(F(t)))

+ kρ − ρk

H1(T1,T2;H2(F(t)))

+ kρ − ρk

W1,∞(T1,T2;H2(F(t)))

+kρ − ρk

H2(T1,T2;L2(F(t)))

+ kuk

L2(T1,T2;H4(F(t)))

+ kuk

L(T1,T2;H3(F(t)))

+ kuk

H1(T1,T2;H2(F(t)))

+kuk

W1,∞(T1,T2;H1(F(t)))

+ kuk

H2(T1,T2;L2(F(t)))

+ k`k

H2(T1,T2)

+ k`k

W1,∞(T1,T2)

+kωk

H2(T1,T2)

+ kωk

W1,∞(T1,T2)

,

(1.13) {notation:norm}

and for T > 0

k(ρ

0

, u

0

, `

0

, ω

0

)k

SbT ,T

= kρ

0

− ρk

H3(F(T))

+ ku

0

k

H3(F(T))

+ |`

0

| + |ω

0

|.

Since we are working with regular solutions of (1.1)–(1.10), we need to introduce the following com- patibility conditions at initial time:

u

0

(y) = `

0

+ ω

0

× (y − h

0

) for y ∈ ∂B(0), u

0

= 0 on ∂Ω, (1.14) {finalcompatibility-1}

− 1 ρ

0

div σ(u

0

, p

0

) = 0 on ∂Ω, (1.15) {finalcompatibility-2}

(5)

ω

0

× (ω

0

× (y − h

0

))

− 1

ρ

0

div σ(u

0

, p

0

)(y)

= 1 m

 Z

∂B(0)

σ(u

0

, p

0

)n dΓ − k

d

`

0

 +

 J

−1

Z

∂B(0)

(x − h

0

) × σ(u

0

, p

0

)n dΓ

x

 × (y − h

0

)

for y ∈ ∂B(0), (1.16) {finalcompatibility-3}

where

p

0

= aρ

γ0

. Finally, we introduce the following notation

0

:= {x ∈ Ω ; dist(x, ∂Ω) > 1} . Our hypotheses on k

p

and k

d

are the following ones:

k

p

∈ C

1

(R

+

, [0, 1]), k

p

(0) = 0, k

p

> 0 in (0, ∞), k

p

≡ 1 in [T

I

, ∞), 0 6 k

0p

< k

d

2T

I2

(1.17) {hypkp}

for some T

I

> 0.

global existence Theorem 1.1. Assume that Ω

0

is non empty and connected. Let h

1

∈ Ω

0

and ρ > 0. Assume w is given by the feedback law (1.10) with (k

p

, k

d

) satisfying (1.17). There exists δ > 0 such that for any

h

0

∈ Ω

0

, ρ

0

∈ H

3

(F (0)), ρ

0

> 0, u

0

∈ H

3

(F(0)), `

0

, ω

0

∈ R

3

, (1.18) {initial condition space:global}

satisfying the compatibility conditions (1.14)–(1.16) with k(ρ

0

, u

0

, `

0

, ω

0

)k

Sb0,0

+ |h

1

− h

0

| 6 δ, (1.19) {smallness}

the system (1.1)–(1.10) admits a unique strong solution (ρ, u, `, ω) ∈ S b

0,∞

, h ∈ L

(0, ∞). Moreover, there exist C, η > 0 such that

k(ρ, u, `, ω)k

Sb0,∞

+ k p

k

p

(h

1

− h)k

L(0,∞)

6 C

k(ρ

0

, u

0

, `

0

, ω

0

)k

Sb0,0

+ |h

1

− h

0

|

, (1.20) {final estimate}

dist(h(t), ∂Ω) > 1 + η (t > 0). (1.21) {1451}

We are now in a position to state our stabilization result.

asymptotic behavior Theorem 1.2. With the notations and assumptions of Theorem 1.1, the solution (ρ, u, h, `, ω) of (1.1)-(1.10) satisfies

t→∞

lim kρ(t, ·) − ρk

H2(F(t))

= 0, lim

t→∞

ku(t, ·)k

H2(F(t))

= 0, (1.22) {limit:fluid}

t→∞

lim h(t) = h

1

, lim

t→∞

`(t) = 0, lim

t→∞

ω(t) = 0. (1.23) {limit:solid}

During the last two decades, there has been a considerable interest in fluid-structure interaction problems involving moving interfaces. Broadly speaking, these types of models can be classified into two types: either the structure is moving inside the fluid or the structure is located at the boundary of the fluid domain. Since in this article we are interested in studying the motion of body inside the compressible fluid domain, below we mention related works from the literature concerning this case only.

The global-in-time existence (up to contact) of weak solutions for compressible viscous flow (for

γ > 2) in a bounded domain of R

3

interacting with a finite number of rigid bodies has been studied

by Desjardins and Esteban [6]. In [9], Feireisl established the global existence result (for γ > 3/2)

regardless of possible collisions of several rigid bodies or a contact of the rigid bodies with the exterior

(6)

boundary. Regarding strong solutions, the existence and uniqueness of global solutions for small initial data have been achieved in [1] in the Hilbert space framework by Boulakia and Guerrero as long as no collisions occur. Their work is based on a method proposed in [13] for a viscous compressible fluid (without structure). In a L

p

-L

q

setting, the authors in [12] proved the existence and uniqueness of local-in-time strong solutions for the system composed by rigid bodies immersed into a viscous compressible fluid and in [11], the authors establish the global in time existence up to contact.

Let us mention some works related to the large time behavior of fluid-structure interaction system.

In [17], the authors analyze the fluid-structure model in one space dimension where the fluid is governed by the viscous Burgers equation and the solid mass is moving by the difference of pressure at both sides of it. They obtain that the asymptotic profile of the fluid is a self-similar solution of the Burgers equation and the point mass enjoys the parabolic trajectory as t → ∞. An extension of this work in several space dimensions is obtained in [14] for the heat equations in interaction with a rigid body.

Their result is that as t → ∞, the fluid solution behaves as the fundamental solution of the heat equation and the ball goes to infinity in bidimensional case whereas the ball remains in a bounded domain in three dimension. Regarding the long-time behavior of a moving particle inside a Navier- Stokes fluid, the authors in [10] consider in particular the case of a ball falling over an horizontal plane and show that the velocity of the fluid goes to zero and the particle reaches the bottom of the container asymptotically in time. In [7], the authors analyze the case of a rigid disk immersed into a two-dimensional Navier-Stokes equations filling the exterior of the structure domain. They restrict to the case of a solid and a fluid with the same density and for the linear case.

Finally, let us mention two works using a control supported on the rigid body: [5] in the 1d case for a Burgers-particle system and [16] in the 3d case for a rigid ball moving into a viscous incompressible fluid. The main difference between this study and the two previous references come from the fact that in our case we need to deal with stronger solutions than in the incompressible case. In particular, to avoid compatibility conditions at t = 0 that involve the feedback control w, we take here k

p

depending on time with k

p

(0) = 0.

The plan of the paper is the following. In Section 2, we establish the local-in-time existence of solu- tions for the system (1.1)–(1.10). We then obtain a priori estimates in Section 3 to prove Theorem 1.1.

Finally Section 4 is devoted to the asymptotic analysis of the solutions in order to prove Theorem 1.2.

Notation. For any a ∈ R

3

, we set

B(a) = b {x ∈ R

3

| |x − a| < 1}, F b (a) = Ω \ B(a). b In particular,

B(t) = B(h(t)), b F (t) = F b (h(t)).

In this article, to shorten the notation, we write H

m

and L

2

instead of H

m

(F(0)) and L

2

(F (0)).

Assume X is Banach space. We need to consider a particular norm for H

m

(0, T ; X) if m ∈ N

and if T ∈ R

+

.

kfk

Hm

(0,T;X)

= kfk

Hm(0,T;X)

+ kf k

Wm−1,∞(0,T;X)

. (1.24) {HmX}

Using the Sobolev embedding, this norm is equivalent to the usual one, but the corresponding constants depend on T and that is the reason why we introduce such a notation.

Assume X

1

and X

2

are Banach spaces. We also introduce the following spaces H

m

(0, T ; X

1

, X

2

) = L

2

(0, T ; X

1

) ∩ H

m

(0, T ; X

2

) (m > 1).

In the case T ∈ R

+

, we also need to introduce the following norm for the above space:

kf k

H1

(0,T;H2,L2)

= kf k

L2(0,T;H2)

+ kf k

L(0,T;H1)

+ kf k

H1(0,T;L2)

, (1.25) {infinity-1}

(7)

kf k

H2

(0,T;H4,L2)

= kf k

L2(0,T;H4)

+ kf k

L(0,T;H3)

+ kf k

H1(0,T;H2)

+ kf k

W1,∞(0,T;H1)

+ kf k

H2(0,T;L2)

.

(1.26) {infinity-2}

Using interpolation results, we see again that the corresponding norm is equivalent to H

1

(0, T ; H

2

) but the corresponding constants depend on T .

2. Local in time existence of solutions sec:Local in time existence of solution

In order to prove Theorem 1.1, we first prove the existence and uniqueness of strong solutions of system (1.1)-(1.10) for small times. More precisely, we show in this section the following result:

local-in-time existence Theorem 2.1. Let h

1

∈ Ω

0

and ρ > 0. Assume w is given by the feedback law (1.10) with k

d

∈ R and k

p

∈ H

loc1

([0, ∞)). There exist δ

0

, C

, T

> 0 such that for any

h

0

∈ Ω

0

, ρ

0

∈ H

3

, u

0

∈ H

3

, `

0

, ω

0

∈ R

3

, (2.1) {initial cond req-1}

satisfying the compatibility conditions (1.14)–(1.16) with k(ρ

0

, u

0

, `

0

, ω

0

)k

Sb0,0

+ |h

1

− h

0

| 6 δ

0

, (2.2) {initial cond req-2}

the system (1.1)-(1.9) admits a unique strong solution (ρ, u, `, ω) ∈ S b

0,T

, h ∈ L

(0, T

) and k(ρ, u, `, ω)k

Sb0,T∗

+ kh

1

− hk

L(0,T)

6 C

k(ρ

0

, u

0

, `

0

, ω

0

)k

Sb0,0

+ |h

1

− h

0

|

. (2.3) {final local estimate}

Change of variables

2.1. Lagrangian change of variables. Firstly, we use a Lagrangian change of variables to rewrite the system (1.1)–(1.10) in a fixed spatial domain: let introduce the flow X(t, ·) : F(0) → F(t) defined by

∂X

∂t (t, y) = u(t, X(t, y)), X(0, y) = y.

Due to the boundary conditions, we have X(t, y) =

( h(t) + Q(t)(y − h

0

) if y ∈ ∂B(0),

y if y ∈ ∂Ω,

where Q(t) ∈ SO(3) is the rotation matrix associated to the angular velocity ω:

Q

0

= A (ω)Q, Q(0) = I

3

. For any ω ∈ R

3

, A (ω) is the skew-symmetric matrix:

A (ω) =

0 −ω

3

ω

2

ω

3

0 −ω

1

−ω

2

ω

1

0

 .

If u is regular enough, X is well-defined and X(t, ·) is a C

1

-diffeomorphism from F(0) onto F(t) for all t ∈ (0, T ). We denote by Y (t, ·) the inverse of X(t, ·) and we consider the following change of variables

u(t, y) = e Q(t)

>

u(t, X(t, y)), ρ(t, y) = e ρ(t, X(t, y)) − ρ, (2.4) {chng of var:fluid}

e h(t) = h(t) − h

1

, `(t) = e Q(t)

>

`(t), e ω(t) = Q(t)

>

ω(t). (2.5) {chng of var:body}

Note that now we have

X(t, y) = y +

t

Z

0

Q(s) e u(s, y)ds, ∀ y ∈ F (0). (2.6) {def:X}

(8)

Under the change of variables (2.4)-(2.5), the system (1.1)-(1.9) is transformed as follows:

∂ ρ e

∂t + ρ

0

div u e = F

1

( ρ, e u, e `, e ω, Q) e in (0, T ) × F (0), (2.7) {reform:fluiddensity}

∂ e u

∂t − µ ρ

0

∆ e u − λ + µ ρ

0

∇ (div u) = e F

2

( ρ, e u, e `, e ω, Q) e in (0, T ) × F (0), (2.8) {reform:fluidvelocity}

m ` e

0

= F

3

( ρ, e u, e e h, e `, ω, Q) e in (0, T ), (2.9) {reform:rigidlinear}

J ω e

0

= F

4

( ρ, e u, e `, e ω, Q) e in (0, T ), (2.10) {reform:rigidangular}

e h

0

= Qe `, Q

0

= Q A ( e ω) in (0, T ), (2.11) {reform:position}

u e = e ` + e ω × (y − h

0

) on (0, T ) × ∂B(0), (2.12) {reformboundary1:vel}

u e = 0 in (0, T ) × ∂Ω, (2.13) {reformboundary2:vel}

ρ(0, e ·) = ρ

0

(·) − ρ, e u(0, ·) = u

0

(·), in F(0), (2.14) {reforminitialcond:densityvel}

e h(0) = h

0

− h

1

, `(0) = e `

0

, ω(0) = e ω

0

, Q(0) = I

3

. (2.15) {reformbody:initial}

In the above equations, F

1

, F

2

, F

3

, F

4

are defined in the following way:

F

1

( ρ, e e u, `, e ω, Q) = e −( ρ e + ρ)∇ e u : h

((∇Y (X))Q)

>

− I

3

i − ( ρ e + ρ − ρ

0

) div u, e (2.16) {F1}

for i = 1, 2, 3:

(F

2

)

i

( ρ, e e u, `, e ω, Q) = e −( ω e × u) e

i

+ µ ρ e + ρ

X

p,l,m

2

e u

i

∂y

m

∂y

l

∂Y

m

∂x

p

(X) ∂Y

l

∂x

p

(X) − δ

mp

δ

lp

+ µ

ρ e + ρ X

p,l

∂ u e

i

∂y

l

2

Y

l

∂x

2p

(X) + µ∆ e u

i

ρ

0

− ( ρ e + ρ) ρ

0

( ρ e + ρ)

+ λ + µ ρ e + ρ

X

p,l

∂ u e

p

∂y

l

2

Y

l

∂x

p

∂x

i

(X) + λ + µ

ρ e + ρ X

p,l,m

2

e u

p

∂y

m

∂y

l

∂Y

m

∂x

p

(X) − δ

mp

∂Y

l

∂x

i

(X) + λ + µ ρ e + ρ

X

p,l

2

e u

p

∂y

p

∂y

l

∂Y

l

∂x

i

(X) − δ

li

+ (λ + µ)[∇(div u)] e

i

ρ

0

− ( ρ e + ρ) ρ

0

( ρ e + ρ)

+ aγ( ρ e + ρ)

γ−2

X

j,l

Q

ji

∂ ρ e

∂y

l

∂Y

l

∂x

j

(X), (2.17) {F2}

F

3

( ρ, e e u, e h, `, e ω, Q) = e −m( ω e × e `) − Z

∂B(0)

h µ

Q∇ u(∇Y e (X)) + (Q∇ e u(∇Y (X)))

>

+ λ (Q∇ u(∇Y e (X)) : I

3

) − a(ρ + ρ) e

γ

i

n dΓ − k

p

Q

>

e h − k

d

e `, (2.18) {F3}

F

4

( ρ, e e u, `, e ω, Q) = e − Z

∂B(0)

(y − h

0

) × h µ

Q∇ e u(∇Y (X)) + (Q∇ u(∇Y e (X)))

>

+ λ (Q∇ e u(∇Y (X)) : Id) − a(ρ + ρ) e

γ

i

n dΓ. (2.19) {F4}

Here n(y) = Q(t)

>

N (t, x) is the unit normal to ∂B(0) at the point y ∈ ∂B(0), directed to the interior

of the ball.

(9)

2.2. Analysis of a linear problem. In this section, we want to study the existence and regularity of the solution of the following linear system:

∂ ρ e

∂t + ρ

0

div u e = f

1

in (0, T ) × F (0), (2.20) {linearfluid:density}

∂ u e

∂t − µ ρ

0

∆ u e − λ + µ ρ

0

∇ (div e u) = f

2

in (0, T ) × F (0), (2.21) {linearfluid:vel}

me `

0

= f

3

in (0, T ), (2.22) {body:linearmom}

J ω e

0

= f

4

in (0, T ), (2.23) {body:angularmom}

u e = ` e + ω e × (y − h

0

) on (0, T ) × ∂B(0), (2.24) {boundary1:vel}

u e = 0 on (0, T ) × ∂Ω, (2.25) {boundary2:vel}

u(0, e ·) = u

0

(·) in F (0), (2.26) {initialcond:vel}

ρ(0, e ·) = ρ e

0

in F(0), (2.27) {initialcond:density}

e `(0) = `

0

, ω(0) = e ω

0

. (2.28) {linearbody:initial}

We introduce the following set for T > 0:

S

T

= n

( ρ, e e u, `, e ω) e | ρ e ∈ H

1

(0, T ; H

3

) ∩ C

1

([0, T ]; H

2

) ∩ H

2

(0, T ; L

2

), e u ∈ H

2

(0, T ; H

4

, L

2

), e ` ∈ H

2

(0, T ), ω e ∈ H

2

(0, T ), u e = 0 on ∂Ω, e u = ` e + ω e × (y − h

0

) on ∂B(0), ρ(0) = e ρ e

0

,

e u(0) = u

0

, `(0) = e `

0

, ω(0) = e ω

0

o

, (2.29) {13:40}

equipped with the norm k( ρ, e u, e `, e ω)k e

ST

:= k ρk e

H1

(0,T;H3)

+ k ρk e

W1,∞(0,T;H2)

+ k ρk e

H2(0,T;L2)

+ k uk e

H2

(0,T;H4,L2)

+ ke `k

H2

(0,T)

+ k e ωk

H2

(0,T)

. We recall that the norms k · k

H1

(0,T;H3)

, k · k

H2

(0,T)

are defined in (1.24) and k · k

H2

(0,T;H4,L2)

is defined in (1.26). The space S

T

is similar to S b

T1,T2

defined by (1.12) except that here F(t) is replaced by F(0) and we add the boundary and initial conditions.

Since ρ > 0, there exists δ

0

> 0 such that (2.2) implies ρ

0

> ρ

2 > 0.

In that case, the system (2.20)–(2.28) is well-posed:

linear system:existence Proposition 2.2. Let us assume ρ > 0, (2.2) with δ

0

as above and

( ρ e

0

, u

0

, `

0

, ω

0

) ∈ H

3

× H

3

× R

3

× R

3

, f

1

∈ L

2

(0, T ; H

3

) ∩ C([0, T ]; H

2

) ∩ H

1

(0, T ; L

2

), f

2

∈ H

1

(0, T ; H

2

, L

2

), f

3

∈ H

1

(0, T ), f

4

∈ H

1

(0, T )

with

u

0

= `

0

+ ω

0

× (y − h

0

) for y ∈ ∂B(0), u

0

= 0 on ∂Ω, (2.30) {comp cond1}

f

2

(0) + µ ρ

0

∆u

0

+ λ + µ ρ

0

∇ (div u

0

) = 0 on ∂Ω, (2.31) {comp cond2}

f

2

(0) + µ ρ

0

∆u

0

+ λ + µ ρ

0

∇ (div u

0

) = m

−1

f

3

(0) + J

−1

f

4

(0) × (y − h

0

) for y ∈ ∂B(0). (2.32) {comp cond3}

(10)

Then the system (2.20)–(2.28) admits a unique solution ( ρ, e e u, `, e ω) e ∈ S

T

. Moreover, there exists C

L

> 0 (nondecreasing with respect to T ) such that

k( ρ, e u, e `, e ω)k e

ST

6 C

L

kf

1

k

L2(0,T;H3)

+ kf

1

k

L(0,T;H2)

+ kf

1

k

H1(0,T;L2)

+ kf

2

k

H1

(0,T;L2,H2)

+ kf

3

k

H1

(0,T)

+ kf

4

k

H1

(0,T)

+ k ρ e

0

k

H3

+ ku

0

k

H3

+ |`

0

| + |ω

0

|

. (2.33) {est:linearsystemfull}

Proof. We solve (2.20)-(2.28) like a cascade system: first, (2.22)-(2.23) admits a unique solution (e `, ω) e with

ke `k

H2

(0,T)

+ k ωk e

H2

(0,T)

6 C kf

3

k

H1

(0,T)

+ kf

4

k

H1

(0,T)

+ |`

0

| + |ω

0

|

. (2.34) {est:body}

Next, we solve equation (2.21) with the boundary and initial conditions (2.24)-(2.26). First we consider a lifting operator R, such that for any a, b ∈ R

3

, R(a, b) ∈ C

(R

3

) satisfies

R(a, b) =

( a + b × (y − h

0

) on ∂B(0),

0 on ∂Ω.

Then e v = u e − R( e `, ω) satisfies e

 

 

∂ e v

∂t − µ ρ

0

∆ e v − λ + µ ρ

0

∇ (div e v) = F = f

2

+ µ ρ

0

∆R(e `, ω) + e λ + µ ρ

0

div R(e `, e ω)

− R(e `

0

, ω e

0

), e v = 0 on (0, T ) × ∂F(0),

e v(0, ·) = e v

0

= u

0

− R(`

0

, ω

0

) in F (0).

By using a standard Galerkin method (see [8, Chapter 7, Theorem 1, p.354]) and by using the regularity result of Lam´ e operator (see, for instance, [4, Theorem 6.3-6, p.296]), under the condition that ∂F(0) is of class C

4

, we can show the following result: if

F ∈ L

2

(0, T ; H

2

) ∩ H

1

(0, T ; L

2

), e v

0

∈ H

3

∩ H

01

, with the condition

F (0, ·) + µ

ρ

0

∆ e v

0

+ λ + µ

ρ

0

∇ (div e v

0

) = 0 on ∂F(0), (2.35) {13:38}

then there exists a unique solution e v ∈ H

2

(0, T ; H

4

, L

2

) with the estimate k vk e

H2

(0,T;H4,L2)

6 C

kF k

H1

(0,T;L2,H2)

+ k v(0)k e

H3

.

We note that condition (2.35) is equivalent to (2.31) and (2.32). We can use the relation u e = e v+R( e `, ω) e and the above estimate of e v to deduce the following estimate of e u:

k uk e

H2

(0,T;H4,L2)

6 C

kf

2

k

H1

(0,T;L2,H2)

+ kf

3

k

H1

(0,T)

+ kf

4

k

H1

(0,T)

+ ku

0

k

H3

+ |`

0

|+ |ω

0

|

. (2.36) {reg:u}

Now, with the help of equation (2.20) satisfied by ρ, we obtain e k ρk e

H1

(0,T;H3)

+ k ρk e

W1,∞(0,T;H2)

+ k ρk e

H2(0,T;L2)

6 C

kf

1

k

L2(0,T;H3)

+ kf

1

k

L(0,T;H2)

+ kf

1

k

H1(0,T;L2)

+ k uk e

L2(0,T;H4)

+ k e uk

L(0,T;H3)

+ k e uk

H1(0,T;H1)

+ k ρ e

0

k

H3

. (2.37) {est:fluiddensity}

Thus, we have proved the existence of solution in appropriate space for the system (2.20)-(2.28).

Thanks to (2.34), (2.36) and (2.37), we have also obtained our required estimate (2.33).

(11)

15:29

2.3. Estimates of the nonlinear terms. For T > 0 and R > 0, we define the following subset of S

T

:

S

T,R

= n

( ρ, e e u, `, e e ω) ∈ S

T

| k( ρ, e u, e `, e ω)k e

ST

6 R o

. (2.38) {ball for fixed point}

In what follows, R is fixed and the constants that appear can depend on R.

Assume ( ρ, e u, e `, e ω) e ∈ S

T ,R

. Then there exists a unique solution ( e h, Q) ∈ H

3

(0, T ) of the following equations

 

 

e h

0

= Q e ` in (0, T ), Q

0

= Q A ( ω) e in (0, T ), Q(0) = I

3

, e h(0) = h

0

− h

1

,

(2.39) {13:43}

and we can then define X by (2.6). From (2.38), there exists C = C(R) > 0 such that kQk

H3(0,T)

6 C, kQ − I

3

k

L(0,T)

6 CT,

ke hk

L(0,T)

6 |h

0

− h

1

| + CT

1/2

. (2.40) {est:h}

In particular, taking δ

0

small enough in (2.2), there exists T

1

= T

1

(R, δ

0

, dist(h

1

, ∂Ω)) > 0 and c

1

> 0 such that

dist( B( b e h(t) + h

1

), ∂Ω) > c

1

> 0 ∀t ∈ [0, T

1

]. (2.41) {12:07}

From now on, we assume T 6 T

1

and the constants may depend on T

1

. Combining (2.6) and (2.38), we also deduce

k∇X − I

3

k

L(0,T;H3)

6 CT

1/2

. (2.42) {inverse:X}

In particular, using the embedding H

3

(F(0)) , → W

1,∞

(F(0)) and (2.41), there exists T

2

6 T

1

such that X : F(0) → F(e b h(t) + h

1

) is invertible and its inverse is denoted by Y .

In the same spirit, using the initial condition on ρ e (see (2.29)), we have

k ρ e + ρ − ρ

0

k

L(0,T;H3)

6 T

1/2

R. (2.43) {15:06bis}

Using the embedding H

3

(F (0)) , → L

(F(0)) and (2.2) with δ

0

small enough, there exists T

3

6 T

2

such that

ρ

2 6 ρ e + ρ 6 3ρ

2 . (2.44) {15:06}

In particular, combining this with (2.38), for any α ∈ R, k( ρ e + ρ)

α

k

L(0,T;H3)

6 C,

Z

∂B(0)

( ρ e + ρ)

γ

ndΓ

H1(0,T)

6 CT

1/2

. (2.45) {rhoalpha}

From the above construction and assuming T 6 T

3

, we can define the terms F

1

, F

2

, F

3

, F

4

by (2.16)-(2.19). To estimate these terms, we first give some estimates of X and Y :

lem:estX Lemma 2.3. Assume ( ρ, e u, e e `, ω) e ∈ S

T,R

. There exists a positive constant C depending only on R, F(0) such that, for all 0 < T 6 T

3

,

k∇Y (X) − I

3

k

L(0,T;H3)

6 CT

1/2

, (2.46) {X-6}

2

Y

l

∂x

p

∂x

i

(X)

L(0,T;H2)

+

∂t (∇Y (X))

L(0,T;H2)

+

∂t

2

Y

l

∂x

p

∂x

i

(X)

L(0,T;H1)

6 C. (2.47) {X-3}

(12)

Proof. From (2.42) and the fact that L

(0, T ; H

3

) is an algebra, we deduce (2.46). This yields in particular that

k∇Y (X)k

L(0,T;H3)

6 C. (2.48) {18:04}

Writing

∂y

m

∂Y

l

∂x

i

(X)

= X

p

2

Y

l

∂x

p

∂x

i

(X) ∂X

p

∂y

m

and using (2.48), we deduce the estimate on

∂x2Yl

p∂xi

(X).

From the expression (2.6), we have

∂t

(∇X(t, ·)) = Q(t)∇ e u(t, ·), and using

∂t (∇Y (X)) = −∇Y (X) ∂

∂t (∇X)∇Y (X), we obtain the estimate of the second term in (2.47).

Finally, we write

∂y

m

∂t ∂Y

l

∂x

i

(X)

= X

p

∂t

2

Y

l

∂x

p

∂x

i

(X) ∂X

p

∂y

m

+ X

p

2

Y

l

∂x

p

∂x

i

(X) ∂

∂t ∂X

p

∂y

m

and from the previous estimate, we have

∂y

m

∂t ∂Y

l

∂x

i

(X)

L(0,T;H1)

+

X

p

2

Y

l

∂x

p

∂x

i

(X) ∂

∂t ∂X

p

∂y

m

L(0,T;H1)

6 C.

Thus, using (2.48), we deduce the estimate of the last term in (2.47).

Next we give some properties on F

1

, F

2

, F

3

, F

4

.

lip:F1 Proposition 2.4. There exist α > 0 and a positive constant C depending on R, k

p

, k

d

, ρ and the other physical parameters, and on F(0) such that, for all 0 < T 6 T

3

, for all

( ρ, e u, e e `, ω), e ( ρ e

1

, u e

1

, ` e

1

, ω e

1

), ( ρ e

2

, u e

2

, e `

2

, ω e

2

) ∈ S

T ,R

, kF

1

( ρ, e e u, `, e e ω, Q)k

L2(0,T;H3)∩L(0,T;H2)∩H1(0,T;L2)

6 CT

α

, kF

2

( ρ, e u, e e `, ω, Q)k e

H1

(0,T;L2,H2)

6 C

T

α

+ kω

0

× u

0

k

H1

+ kaγρ

γ−20

∇ρ

0

k

H1

, kF

3

( ρ, e e u, e h, `, e ω, Q)k e

H1

(0,T)

6 C (T

α

+ |ω

0

× `

0

| + |`

0

| + kρ

0

− ρk

H1

+ ku

0

k

H3

) , kF

4

( ρ, e e u, e `, e ω, Q)k

H1

(0,T)

6 CT

α

, and

kF

1

( ρ e

1

, u e

1

, e `

1

, ω e

1

, Q

1

) − F

1

( ρ e

2

, u e

2

, ` e

2

, ω e

2

, Q

2

)k

L2(0,T;H3)∩L(0,T;H2)∩H1(0,T;L2)

6 CT

α

k( ρ e

1

, u e

1

, ` e

1

, ω e

1

) − ( ρ e

2

, e u

2

, e `

2

, ω e

2

)k

ST

, kF

2

( ρ e

1

, u e

1

, e `

1

, ω e

1

, Q

1

) − F

2

( ρ e

2

, u e

2

, e `

2

, ω e

2

, Q

2

)k

H1

(0,T;L2,H2)

6 CT

α

k( ρ e

1

, e u

1

, ` e

1

, ω e

1

) − ( ρ e

2

, e u

2

, ` e

2

, ω e

2

)k

ST

, kF

3

( ρ e

1

, u e

1

, e h

1

, e `

1

, ω e

1

, Q

1

) −F

3

( ρ e

2

, e u

2

, e h

2

, ` e

2

, ω e

2

, Q

2

)k

H1

(0,T)

6 CT

α

k( ρ e

1

, e u

1

, e `

1

, ω e

1

)− ( ρ e

2

, e u

2

, ` e

2

, ω e

2

)k

ST

, kF

4

( ρ e

1

, u e

1

, ` e

1

, ω e

1

, Q

1

) − F

4

( ρ e

2

, e u

2

, ` e

2

, e ω

2

, Q

2

)k

H1

(0,T)

6 CT

α

k( ρ e

1

, u e

1

, e `

1

, ω e

1

) − ( ρ e

2

, u e

2

, ` e

2

, ω e

2

)k

ST

.

where Q, Q

1

, Q

2

, e h, e h

1

, e h

2

∈ H

3

(0, T ) are given by (2.39).

(13)

Proof. Using the definition (2.16) of F

1

, (2.43), (2.29), (2.38), (2.46) we have the following estimates kF

1

k

L2(0,T;H3)

6 Ck( ρ e + ρ)k

L(0,T;H3)

k∇ uk e

L2(0,T;H3)

k((∇Y )Q)

>

− I

3

k

L(0,T;H3)

+ Ck( ρ e + ρ − ρ

0

)k

L(0,T;H3)

k div e uk

L2(0,T;H3)

6 CT

α

,

∂F

1

∂t

L2(0,T;L2)

6 Ck( ρ e + ρ)k

L(0,T;H3)

(

∇ ∂ u e

∂t

L2(0,T;L2)

k((∇Y )Q)

>

− I

3

k

L(0,T;H3)

+k∇ uk e

L2(0,T;H3)

∂t ((∇Y (X))Q)

>

L(0,T;H2)

)

+ CT

1/2

∂ ρ e

∂t

L(0,T;H2)

k∇ uk e

L(0,T;H2)

((∇Y (X))Q)

>

L(0,T;H3)

+ Ck ρ e + ρ − ρ

0

k

L(0,T;H3)

div ∂ u e

∂t

L2(0,T;H1)

6 CT

α

, (2.49) {time derivative F1}

kF

1

k

L(0,T;H2)

6 Ck ρ e + ρk

L(0,T;H3)

k∇ uk e

L(0,T;H2)

k((∇Y )Q)

>

− I

3

k

L(0,T;H3)

+ Ck ρ e + ρ − ρ

0

k

L(0,T;H3)

k div uk e

L(0,T;H2)

6 CT

1/2

. Let us now estimate the L

2

(0, T ; H

2

) norm of F

2

. Here we only estimate some terms in (2.17), the other terms can be estimated similarly. Using (2.45), (2.29), (2.38), (2.46), (2.47),

1 ρ e + ρ

2

u e

i

∂y

m

∂y

l

∂Y

m

∂x

p

(X) ∂Y

l

∂x

p

(X) − δ

mp

δ

lp

L2(0,T;H2)

6 C

1 ρ e + ρ

L(0,T;H3)

2

u e

i

∂y

m

∂y

l

L2(0,T;H2)

∂Y

m

∂x

p

(X) ∂Y

l

∂x

p

(X) − δ

mp

δ

lp

L(0,T;H2)

6 CT

α

,

1 ρ e + ρ

∂ e u

i

∂y

l

2

Y

l

∂x

2p

(X)

L2(0,T;H2)

6 CT

1/2

1 ρ e + ρ

L(0,T;H3)

∂ e u

i

∂y

l

L(0,T;H2)

2

Y

l

∂x

2p

(X)

L(0,T;H2)

6 CT

α

,

( ρ e + ρ)

γ−2

∂ ρ e

∂y

l

∂Y

l

∂x

j

(X)

L2(0,T;H2)

6 CT

1/2

k( ρ e + ρ)

γ−2

k

L(0,T;H2)

∂ ρ e

∂y

l

L(0,T;H2)

∂Y

l

∂x

j

(X)

L(0,T;H2)

6 CT

α

.

For the estimate of the H

1

(0, T ; L

2

) norm of F

2

, we also only give the estimates the L

2

(0, T ; L

2

)

norm of some terms of the time derivative F

2

. Again, the other terms can be estimated similarly.

(14)

First, we write

∂t 1

ρ e + ρ

2

e u

i

∂y

m

∂y

l

∂Y

m

∂x

p

(X) ∂Y

l

∂x

p

(X) − δ

mp

δ

lp

= − 1

( ρ e + ρ)

2

∂ ρ e

∂t

2

u e

i

∂y

m

∂y

l

∂Y

m

∂x

p

(X) ∂Y

l

∂x

p

(X) − δ

mp

δ

lp

+ 1

ρ e + ρ

3

e u

i

∂t∂y

m

∂y

l

∂Y

m

∂x

p

(X) ∂Y

l

∂x

p

(X) − δ

mp

δ

lp

! +

1 ρ e + ρ

2

e u

i

∂y

m

∂y

l

∂t ∂Y

m

∂x

p

(X) ∂Y

l

∂x

p

(X)

,

∂t 1

ρ e + ρ

∂ u e

i

∂y

l

2

Y

l

∂x

2p

(X)

= − 1

( ρ e + ρ)

2

∂ ρ e

∂t

∂ u e

i

∂y

l

2

Y

l

∂x

2p

(X) + 1 ρ e + ρ

2

e u

i

∂t∂y

l

2

Y

l

∂x

2p

(X)

+ 1

ρ e + ρ

∂ u e

i

∂y

l

∂t ∂

2

Y

l

∂x

2p

(X)

. Using (2.45), (2.29), (2.38), (2.46), (2.47), we deduce that the above terms is estimated in L

2

(0, T ; L

2

) {use:1/f est}

{use:1/f est}

{use:1/f est}

{use:1/f est}

{use:1/f est}

{use:1/f est}

{use:1/f est}

{use:1/f est}

{use:1/f est}

{use:1/f est}

by CT

α

.

Finally, to obtain the L

(0, T ; H

1

) estimate of the term F

2

, we use the following inequality [15, Lemma 4.2]:

sup

t∈(0,T)

kF

2

(t)k

H1

6 C

kF

2

k

L2(0,T;H2)

+ kF

2

k

H1(0,T;L2)

+ kF

2

(0)k

H1

, and since

kF

2

(0)k

H1

6 kω

0

× u

0

k

H1

+ kaγρ

γ−20

∇ρ

0

k

H1

, we deduce the result for F

2

.

It remains to estimate F

3

and F

4

. We only consider F

3

, the analysis for F

4

is the same. From (2.18), we can see that the time derivative of F

3

involves the following terms (and similar ones)

( ω e × `) e

0

, (k

p

Q

>

e h)

0

, k

d

` e

0

, Z

∂B(0)

Q

0

∇ u∇Y e (X) + Q∇ ∂ u e

∂t ∇Y (X) + Q∇ u e ∂

∂t ∇Y (X)

! n dΓ

− aγ Z

∂B(0)

(ρ + ρ) e

γ−1

∂ ρ e

∂t n dΓ.

Almost all the terms can be estimated in a direct way in L

2

(0, T ) by using (2.40), (2.45), (2.29), (2.38), (2.46). We have nevertheless to take care of

Z

∂B(0)

Q∇ ∂ u e

∂t ∇Y (X)n dΓ.

For this term, we use standard interpolation result (see, for instance, [2, Lemma A.5]) to obtain

∇ ∂ e u

∂t

L8/3(0,T;H1/4)

6 C

∇ ∂ e u

∂t

1/4 L(0,T;L2)

∇ ∂ u e

∂t

3/4 L2(0,T;H1)

,

where C is independent of T . Using a trace result and (2.29), (2.38), we deduce an estimate of F

30

in L

2

(0, T ) of the form CT

α

. To end the estimate of F

3

, we use that

kF

3

k

L(0,T)

6 |F

3

(0)| + T

1/2

kF

3

k

H1(0,T)

.

(15)

We have the following estimate:

Z

∂B(0)

(ρ + ρ(0)) e

γ

n dΓ

=

Z

∂B(0)

ρ

γ0

n dΓ

=

Z

∂B(0)

γ0

− ρ

γ

)n dΓ

6 C Z

∂B(0)

0

− ρ| dΓ.

Thus,

|F

3

(0)| 6 C (|ω

0

× `

0

| + |`

0

| + kρ

0

− ρk

H1

+ ku

0

k

H3

) .

The estimates for the differences can be done in a similar way and we thus skip the corresponding

proof.

2.4. Proof of Theorem 2.1.

Proof. We are going to establish the local in time existence of (2.7)-(2.19). In order to do this we use a fixed-point argument.

Assume ρ > 0, δ

0

satisfying the smallness assumptions introduced in the above section and let us consider (ρ

0

, u

0

, h

0

, `

0

, ω

0

) satisfying (2.1), (2.2). Recall that from (2.44), we have ρ

2 6 ρ

0

6 3ρ 2 and thus, using Sobolev embeddings, there exists C

1

> 0 depending on ρ, δ

0

and the geometry such that

C

0

× u

0

k

H1

+ kaγρ

γ−20

∇ρ

0

k

H1

+ |ω

0

× `

0

| + |`

0

| + kρ

0

− ρk

H1

+ ku

0

k

H3

6 C

1

e δ

0

(2.50) {est:F20}

where C is the constant appearing in Proposition 2.4 and where we have set e δ

0

= kρ

0

− ρk

H3

+ ku

0

k

H3

+ |h

1

− h

0

| + |`

0

| + |ω

0

| 6 δ

0

. We now fix R > 0 as

R = 2C

L

C

1

e δ

0

, (2.51) {radius of ball}

where C

L

is the continuity constant in estimate (2.33). We take T 6 T

3

, where T

3

= T

3

(R) is the time obtained in the above section.

Let us define the following mapping

N : S

T ,R

→ S

T,R

(2.52) {10:12}

( ρ, e e u, `, e ω) e 7→ ( ρ, b u, b b `, ω). b (2.53) For ( ρ, e e u, `, e e ω) ∈ S

T ,R

, we define X by (2.6), e h and Q by (2.39) and F

1

, F

2

, F

3

, F

4

by (2.16)-(2.19).

Then ( ρ, b b u, b `, b ω) is the solution of

∂ ρ b

∂t + ρ

0

div u b = F

1

( ρ, e u, e `, e ω, Q) e in (0, T ) × F (0), (2.54) {fixedpoint:fluiddensity}

∂ b u

∂t − µ ρ

0

∆ b u − λ + µ ρ

0

∇ (div u) = b F

2

( ρ, e u, e `, e ω, Q) e in (0, T ) × F (0), (2.55) {fixedpoint:fluidvelocity}

mb `

0

= F

3

( ρ, e u, e e h, e `, ω, Q) e in (0, T ), (2.56) {fixedpoint:rigidlinear}

J ω b

0

= F

4

( ρ, e u, e `, e ω, Q) e in (0, T ), (2.57) {fixedpoint:rigidangular}

u b = ` b + ω b × (y − h

0

) on (0, T ) × ∂B(0), (2.58) {fixedpointboundary1:vel}

u b = 0 in (0, T ) × ∂Ω. (2.59) {fixedpointboundary2:vel}

ρ(0, b ·) = ρ

0

(·) − ρ, u(0, b ·) = u

0

(·) in F(0), (2.60) {fixedpointinitialcond:densityvel}

b `(0) = `

0

, b ω(0) = ω

0

. (2.61) {fixedpointbody:initial}

(16)

In order to show that N is well-defined, we apply Proposition 2.2 to the above system. First we note that (1.14)–(1.16) yield the compatibility conditions (2.30)–(2.32). More precisely, the first condition is exactly condition (1.14). Using the expression of F

2

in (2.17), we have

h

F

2

( ρ, e e u, `, e ω, Q) e i

(0, ·) = −ω

0

× u

0

+ 1 ρ

0

∇p

0

, where p

0

= aρ

γ0

. Thus, (1.15) yields the second condition.

On the other hand, using the expressions of F

3

and F

4

in (2.18) and (2.19), we have h

F

3

( ρ, e u, e `, e ω, Q) e i

(0, ·) = −m(ω

0

× `

0

) − Z

∂B(0)

σ(u

0

, p

0

)n dΓ − k

d

`

0

,

h

F

4

( ρ, e e u, `, e ω, Q) e i

(0, ·) = − Z

∂B(0)

(y − h

0

) × σ(u

0

, p

0

)n dΓ.

These expressions of F

3

(0, ·) and F

4

(0, ·) show that (1.16) gives the third condition (2.32). We thus deduce from Proposition 2.2 the existence and uniqueness of ( ρ, b u, b `, b ω) b ∈ S

T

. Combining (2.33), Proposition 2.4, (2.50) and (2.51), we obtain

k( ρ, b u, b `, b ω)k b

ST

6 R

2 + CT

α

. In particular, taking T small enough, we deduce that N is well defined.

Next we show that N is a contraction. Let ( ρ e

1

, e u

1

, ` e

1

, e ω

1

), ( ρ e

2

, e u

2

, ` e

2

, e ω

2

) ∈ S

T ,R

. For j = 1, 2, we set N ( ρ e

j

, u e

j

, ` e

j

, ω e

j

) := ( ρ b

j

, b u

j

, ` b

j

, b ω

j

). Using Proposition 2.2 and Proposition 2.4, we obtain

k( ρ b

1

, u b

1

, b `

1

, ω b

1

) − ( ρ b

2

, u b

2

, ` b

2

, ω b

2

)k

ST

6 CT

α

k( ρ e

1

, e u

1

, ` e

1

, ω e

1

) − ( ρ e

2

, u e

2

, ` e

2

, ω e

2

)k

ST

. Thus N is a contraction in S

T ,R

for T small enough.

Finally, using (2.51) and (2.39), we deduce k( ρ, e u, e `, e ω)k e

ST

+ kh

1

− hk

L(0,T)

6 C e δ

0

= C

0

− ρk

H3

+ ku

0

k

H3

+ |h

1

− h

0

| + |`

0

| + |ω

0

|

that yields (2.3).

3. Global in time existence of solutions sec:Global in time existence of solution

3.1. A priori estimates. We have already established a local-in-time existence result in Theorem 2.1.

In order to obtain the global in time existence of the solutions, we need an appropriate a priori estimates. We recall that k · k

Sb0,T

is introduced in (1.13). We also introduce the following notation to shorten the notation: for Z = L

p

or Z = W

k,p

, we set:

W

Tk,∞

(Z ) = W

k,∞

(0, T ; Z(F (t))), H

Tk

(Z) = H

k

(0, T ; Z(F (t))), for k = 1, 2, W

T0,∞

(Z) = L

T

(Z) = L

(0, T ; Z(F(t))), H

T0

(Z) = L

2T

(Z) = L

2

(0, T ; Z(F (t))).

The main tool to prove the global in time existence of the solutions is the following proposition:

aprori est Proposition 3.1. Let h

1

∈ Ω

0

and ρ > 0. Assume the feedback law (1.10) with (k

p

, k

d

) satisfying (1.17). There exist ε

0

, C

0

> 0 with ε

0

6 δ

0

such that if (ρ, u, h, `, ω) is a solution of system (1.1)–(1.10) with

k(ρ, u, `, ω)k

Sb0,T

6 ε

0

, (3.1) {smallness cond:apriori}

Références

Documents relatifs

Le second modèle proposé (chapitre 5 ) tire parti de l’analyse des forces des deux familles d’approches : d’une part notre modèle utilise la rapidité d’inférence des

Recently the case where Navier slip-with-friction boundary conditions are prescribed only on the body boundary ∂S(t) and no-slip conditions on ∂Ω has been studied: existence of

Si les soins palliatifs semblent désormais attractifs financièrement (notamment lorsqu’ils sont pratiqués par des équipes classiques, avec ou sans lits dédiés), rien n’a

We test our method by comparing the velocity field around a single squirmer and the pairwise interactions between two squirmers with exact solutions to the Stokes equations and

Section IV deals with the analysis of the performances obtained by means of this pulsed current source for the control of the UV emission; they confirm the very tight

ﻟا ﺮﯿﺛﺄﺗ ﺖﺤﺗ داﻮﻤﻟا هﺬھ ﻲﻓ ﺔﺟوﺰﻠﻟا ﺮﺴﻛ كﻮﻠﺳ ﻰﻠﻋ ﻞﻣاﻮﻌﻟا ﺾﻌﺑ ﺮﯿﺛﺄﺗ ﺢﯿﺿﻮﺘﻟ ﻦﻣ اﺪﯿﺟ ﺎﻔﺻو ﺐﻠﻄﺘﯾ ﺖﻠﻔﺳﻹا كﻮﻠﺳ ﻞﯿﻠﺤﺗ .ﺔﻔﻠﺘﺨﻤﻟا طﻮﻐﻀ ﺟإ حﺮﺘﻘﻧ ﺎﻨﻧإ ،ﺔﺳارﺪﻟا

Existence of contacts for the motion of a rigid body into a viscous incompressible fluid with the Tresca boundary conditions... Existence of contacts for the motion of a rigid body

The mathematical study of fluid-structure interaction systems has been the subject of an intensive re- search since around 2000. A large part of the articles devoted to this