• Aucun résultat trouvé

Analysis of a simplified model of rigid structure floating in a viscous fluid

N/A
N/A
Protected

Academic year: 2021

Partager "Analysis of a simplified model of rigid structure floating in a viscous fluid"

Copied!
37
0
0

Texte intégral

(1)

HAL Id: hal-01889892

https://hal.archives-ouvertes.fr/hal-01889892

Submitted on 8 Oct 2018

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.

Analysis of a simplified model of rigid structure floating in a viscous fluid

Debayan Maity, Jorge San Martin, Takéo Takahashi, Marius Tucsnak

To cite this version:

Debayan Maity, Jorge San Martin, Takéo Takahashi, Marius Tucsnak. Analysis of a simplified model

of rigid structure floating in a viscous fluid. Journal of Nonlinear Science, Springer Verlag, 2019, 29

(5), pp.1975-2020. �10.1007/s00332-019-09536-5�. �hal-01889892�

(2)

Analysis of a simplified model of rigid structure floating in a viscous fluid.

Debayan Maity

†1

, Jorge San Mart´ın

2

, Tak´ eo Takahashi

∗3

, and Marius Tucsnak

†1

1

Institut de Math´ ematiques, Universit´ e de Bordeaux, Bordeaux INP, CNRS, F-33400 Talence, France

2

DIM, Universidad de Chile , Blanco Encalada, Chile

3

Universit´ e de Lorraine, CNRS, Inria, IECL, F-54000 Nancy, France debayan.maity@u-bordeaux.fr, jorge@dim.u-chile.cl, takeo.takahashi@inria.fr,

marius.tucsnak@u-bordeaux.fr

October 8, 2018

Abstract

We study the interaction of surface water waves with a floating solid constraint to move only in the vertical direction. The first novelty we bring in is that we propose a new model for this interaction, taking into consideration the viscosity of the fluid. This is done supposing that the flow obeys a shallow water regime (modeled by the viscous Saint-Venant equations in one space dimension) and using a Hamiltonian formalism. Another contribution of this work is establishing the well-posedness of the obtained PDEs/ODEs system in function spaces similar to the standard ones for strong solutions of viscous shallow water equations. Our well- posedness results are local in time for every initial data and global in time if the initial data are close (in appropriate norms) to an equilibrium state. Moreover, we show that the linearization of our system around an equilibrium state can be described, at least for some initial data, by an integro-fractional differential equation related to the classical Cummins equation and which reduces to the Cummins equation when the viscosity vanishes and the fluid is supposed to fill the whole space. Finally, we describe some numerical tests, performed on the original nonlinear system, which illustrate the return to equilibrium and the influence of the viscosity coefficient.

Key words. Viscous shallow water equations, floating structure, fluid-structure interaction, strong solutions, return to equilibrium.

AMS subject classifications. 35Q35, 74F10.

Contents

1 Introduction 2

2 Modelling 3

3 Energy estimate and main result 8

Tak´eo Takahashi acknowledges the support of the Agence Nationale de la Recherche, project IFSMACS, ANR- 15-CE40-0010.

Debayan Maity and Marius Tucsnak acknowledge the support of the Agence Nationale de la Recherche - Deutsche Forschungsgemeinschaft (ANR - DFG), project INFIDHEM, ID ANR-16-CE92-0028.

(3)

4 Equations (3.1)–(3.12) as a parabolic system onE 10

5 Local in time existence and uniqueness 12

6 Global in time existence and uniqueness 16

6.1 Linearization around a stationary state. . . 16 6.2 Study of the Linear Problem . . . 19 6.3 Fixed point . . . 25

7 Return to the Equilibrium 27

8 Numerical Simulations 32

1 Introduction

The motion of rigid bodies in a fluid is a widely studied subject in both engineering and mathe- matical literature. This is due to important applications (naval and aerospace engineering, biology, medicine,. . .) and to the mathematical challenges, namely the existence of free boundaries, raised by the corresponding mathematical models. The study of the case in which the rigid bodies are completely immersed in the fluid goes back to Euler, Kelvin and Kirchhoff. Due to the important number of works (namely in the last two decades) devoted to this subject, the corresponding math- ematical questions may be considered by now well understood, for various types of fluids (ideal, viscous incompressible, compressible,. . .). We refer, for instance, to [4] and references therein for a concise description of recent progress in this field. The case when the solids are floating, thus only partially immersed and interacting with the water waves, has been much less studied in the literature, essentially in the case of an ideal fluid. We refer to the classical work of John [11, 12]

or to the monograph Falnes [7] for the linearized theory and to the review paper of Lannes [14] for the description of recent progress in the field. As far as we know, the case of a viscous fluid has not been tackled, at least from a mathematical view point. The aim of this work is to partially fill this gap by considering a PDE system modeling the coupled motion of a free surface viscous fluid and of a solid, constrained to move in the vertical direction only, which is floating on this free surface. The main contributions of this work are:

• Proposing, via a Hamiltonian formalism, a new mathematical model which takes into account the viscosity of the fluid, for the coupled of a free boundary shallow fluid and of a rigid body floating on the surface.

• Proving the existence and uniqueness of strong solutions (locally in time or globally in time for small data).

• Proposing a reduced model and performing numerical simulations in the linear case.

The content of the subsequent sections is outlined below. In Section2 we derive the governing equations using a Hamiltonian formalism. In Section 3 we show that our model is energetically consistent and we state our main existence and uniqueness results. In Section4 we write the gov- erning equation in an equivalent form obtained by “eliminating” the unknown functions involving the floating solid. Section 5 is devoted to the proof of the local in time existence and uniqueness result, whereas in Section6 we prove global in time existence and uniqueness of solutions for ini- tial data which are close to an equilibrium state. Finally, in Section 7 we analyze the return to equilibriumproblem and in Section8, we present some numerical simulations for this problem.

(4)

h(t, x)

H (t) M

0 a b `

Figure 1: Configuration

2 Modelling

In this section, we derive, using a Hamiltonian formalism, the simplified model which is analyzed in the remaining part of this work. The fluid is described using the viscous Saint Venant equations, whereas the solid obeys Newton’s second law. To couple the two models we use a Hamiltonian formalism, passing by the following steps:

• Introduce the total energy of the fluid-floating object system within the Saint-Venant ap- proximation and define conservation of mass as a constraint, following Petit and Rouchon [17].

• Use of a Hamiltonian formalism to derive the governing equations, by combining the ap- proach introduced in [17] for the inviscid case with the methodology used in Gay-Balmaz and Yoshimura [8] in the case the Navier-Stokes-Fourier system, which allows us to include the viscous effects in the system.

To describe our model we need some notation, which is introduced below. Denote respectively byρ(t, x),v(t, x) andh(t, x) the density, the velocity, and the height of the free surface of the fluid (see Figure 1). These quantities depend on the time t>0 and on the positionx∈[0, `].We also denote by H(t) the height of the cylinder. The mass M of the rigid body is a positive constant.

We also assume that the cylinder moves only vertically and we set

I:= [a, b] := the projection of the cylinder on the flat bottom, E:= (0, `)\ I, (2.1) withI ⊂(0, `).

The functionH, which depends only on time is supposed to satisfy 0< H(t)<min(h(t, a), h(t, b+)) (t>0),

so that the cylinder is immersed into the fluid and does not touch the bottom.

We are now on a position to derive the governing equations. The first one is the equation of mass conservation:

∂h

∂t +∂(hv)

∂x = 0 (t >0, x∈[0, `]). (2.2)

Forx∈ I the height of fluid is given by the position of the cylinder, i.e.:

h(t, x) =H(t) (t >0, x∈ I). (2.3)

Equations (2.2) and (2.3) appear below as constraints in our Hamiltonian formalism.

(5)

In order to derive the other equations, we introduce the kinetic energy Kf and the potential energyUf of the fluid:

Kf = 1 2 Z

(0,`)

ρhv2dx,

Uf = Z

(0,`)

h1

2ρgh2+e(s(t, x))i dx,

where g is the gravity acceleration, and e(s) represents the internal energy density which is a function of the density of entropys. We assume that the fluid is homogeneous and incompressible, so that that its densityρis a constant.

We also introduce the kinetic energyKf and the potential energyUf of the solid:

Ks= 1

2MH˙2 Us=M gH.

We obtain below the governing equations by writing the stationarity for the total action of the system under the constraints (2.2) and (2.3). To this aim, it seems more convenient, in particular to treat the continuity restriction (2.2), to introduce the variableqdefined by

q(t, x) =h(t, x)v(t, x) (t >0, x∈[0, `]). (2.4) The total action of the system is given by

A(h, v, H) = (Kf+Ks)−(Uf+Us)

= Z T

0

nZ ` 0

h1

2ρhv2−1

2ρgh2−e(s(t, x))i dx+1

2MH˙2−M gHo dt.

In order to include the restrictions (2.2)–(2.3), we introduce two Lagrange multipliers:

λ1(t, x) (t>0, x∈[0, `]), λ2(t, x) (t>0, x∈ I).

With the above notation, the governing equations are obtained as stationarity conditions for the Lagrangian defined by

L(h, q, H, λ1, λ2) = Z T

0

(Z ` 0

h1 2ρq2

h −1

2ρgh2+ρλ1

∂h

∂t +∂q

∂x

−e(s(t, x))i dx

+ Z

I

2

H−hi dx+1

2MH˙2−M gH )

dt. (2.5) To achieve this aim, we consider virtual displacements of the trajectory, given by the indepen- dent variations δh(x, t), δH(t) andδq. Since q is a velocity type quantity, whereas hand H are displacements, we introduce the functionϕand its variationδϕdefined by

∂ϕ

∂t =q and ∂(δϕ)

∂t =δq. (2.6)

The governing equations are obtained by imposing that δL = 0 for any virtual displacement δh(t, x),δH(t) andδϕ(t, x) such that

δh(0, x) =δh(T, x) =δH(0) =δH(T) =δϕ(0, x) =δϕ(T, x) = 0, (2.7)

δϕ(t,0) =δϕ(t, `) = 0. (2.8)

(6)

Moreover, to take in consideration friction forces inside the fluid, we follow the approach in [8] by assuming that the entropy trajectorysand its variationδssatisfy the variational constraint

∂e

∂sδs=µ h

∂q

∂x

∂(δϕ)

∂x (2.9)

where the constantµ >0 is the viscosity of the fluid, together with the phenomenological constraint

∂e

∂ss˙=µ h

∂q

∂x 2

.

Note that the above expression of this friction term leads to an nonstandard form of the viscous Saint-Venant equations, see Remark2.1below.

Using (2.5) we get

δL= Z T

0

(Z `

0

h ρq

h(δq)−1 2ρq2

h2(δh)−ρgh(δh) +ρλ1∂(δh)

∂t +∂(δq)

∂x

i−∂e

∂s(δs) dx

+ Z

I

h λ2

δH −δhi

dx+MH˙(δH)˙ −M g(δH) )

dt.

The above formula and (2.6) yield:

δL= Z T

0

(Z `

0

h ρq

h

∂(δϕ)

∂t −1 2ρq2

h2(δh)−ρgh(δh) +ρλ1∂(δh)

∂t + ∂

∂x

∂(δϕ)

∂t

i−∂e

∂s(δs) dx

+ Z

I

h λ2

(δH)−(δh)i

dx+MH(δ˙ H˙)−M g(δH) )

dt. (2.10) Considering (2.7) and integrating by parts in time the formula (2.10), we find

δA= Z T

0

nZ ` 0

h−ρ∂

∂t q

h

δϕ−ρ∂λ1

∂t

∂xδϕ−∂e

∂sδsi dx +

Z ` 0

h−1 2ρq2

h2 −ρgh−ρ∂λ1

∂t δhi

dx− Z

I

2δhi dx +Z

I

λ2dx−MH¨ −M g δHo

dt. (2.11) Using the model (2.9) in (2.11) we get

δA= Z T

0

(Z `

0

h−ρ∂

∂t q

h

(δϕ)−ρ∂λ1

∂t

∂x(δϕ)−µ1 h

∂q

∂x

∂(δϕ)

∂x i

dx +

Z ` 0

h−1 2ρq2

h2 −ρgh−ρ∂λ1

∂t

(δh)i dx−

Z b a

h

λ2(δh)i dx +Z b

a

λ2dx−MH¨ −M g (δH)

)

dt. (2.12)

Now we integrate by parts in space but we need to take care about the fact that some quantities can present discontinuities at the interfaces solid-liquid. We thus introduce the notation

[f]a:=f(a+)−f(a), [f]b:=f(b+)−f(b) (2.13)

(7)

and we use the following formula for the integration by parts with discontinuities:

Z ` 0

uv0 dx=− Z `

0

u0v dx−[uv]a−[uv]b. (2.14) We thus obtain

δA= Z T

0

nZ ` 0

h−ρ∂

∂t q

h

+ρ ∂

∂x ∂λ1

∂t

+ ∂

∂x µ

h

∂q

∂x i

(δϕ) dx +h

ρ∂λ1

∂t +µ h

∂q

∂x i

a

(δϕ)(t, a) +h ρ∂λ1

∂t +µ h

∂q

∂x i

b

(δϕ)(t, b) +

Z ` 0

h−1 2ρq2

h2 −ρgh−ρ∂λ1

∂t

(δh)i dx−

Z b a

h

λ2(δh)i dx +Z b

a

λ2dx−MH¨ −M g (δH)o

dt. (2.15) SinceδA= 0 for any virtual displacement, we get the following system of equations

∂t q

h − ∂

∂x

∂λ1

∂t = µ ρ

∂x 1

h

∂q

∂x

(t >0, x∈[0, `]), (2.16) h

ρ∂λ1

∂t +µ h

∂q

∂x i

a= 0 (t >0), (2.17)

h ρ∂λ1

∂t +µ h

∂q

∂x i

b= 0 (t >0), (2.18)

−∂λ1

∂t = 1 2

q2

h2 +gh+λ2

ρ (t >0, x∈ I), (2.19)

−∂λ1

∂t = 1 2

q2

h2 +gh (t >0, x∈ E) (2.20)

MH¨ = Z b

a

λ2dx−M g (t >0). (2.21)

We remark that equation (2.21) gives an interpretation to the Lagrange multiplierλ2as the pressure force exerted by the fluid on the solid. For this reason, we write in what follows

p(t, x) :=λ2(t, x).

Combining equations (2.16)–(2.20), we can eliminate the Lagrange multiplierλ1and we obtain:

∂h

∂t +∂q

∂x = 0 (t >0, x∈ I ∪ E), (2.22)

∂t q

H + ∂

∂x 1

2 q2

H2 +gH+p ρ

= µ ρH

2q

∂x2 (t >0, x∈ I), (2.23)

∂t q

h

+ ∂

∂x 1

2 q2 h2 +gh

= µ

ρ

∂x 1

h

∂q

∂x

(t >0, x∈ E), (2.24) h1

2 q2

H2 +gH+p ρ

− µ Hρ

∂q

∂x i

(a+) = h1 2

q2 h2 +gh

− µ hρ

∂q

∂x i

(a) (t >0), (2.25) h1

2 q2

H2+gH+p ρ

− µ Hρ

∂q

∂x i

(b) = h1 2

q2 h2 +gh

− µ hρ

∂q

∂x i

(b+) (t >0), (2.26) MH(t)¨ = −M g+

Z b a

p(t, x) dx (t >0), (2.27) where we have also rewritten (2.2) and used (2.3) in the solid part. In the solid part we can simplify once again this system, using the fact that ash(t, x) =H(t) does not depend onx, then equation (2.22) implies thatq is a first degree polynomial.

(8)

Then our model for the rigid rectangle solid, constrained to move in the vertical direction, which is floating at the surface of a viscous fluid is

∂h

∂t +∂q

∂x = 0 (t >0, x∈ I ∪ E), (2.28)

∂t q

H + ∂

∂x 1

2 q2

H2 +gH+p ρ

= 0 (t >0, x∈ I), (2.29)

∂t q

h

+ ∂

∂x 1

2 q2 h2 +gh

= µ

ρ

∂x 1

h

∂q

∂x

(t >0, x∈ E), (2.30) h1

2 q2

H2 +gH+p ρ

− µ Hρ

∂q

∂x i

(t, a+) = h1 2

q2 h2 +gh

− µ hρ

∂q

∂x i

(t, a) (t >0), (2.31) h1

2 q2

H2+gH+p ρ

− µ Hρ

∂q

∂x i

(t, b) = h1 2

q2 h2 +gh

− µ hρ

∂q

∂x i

(t, b+) (t >0), (2.32) MH(t)¨ = −M g+

Z b a

p(t, x) dx (t >0).(2.33) Remark 2.1. As mentioned above, our modeling approach yields a nonstandard viscous term in the Saint-Venant system, more precisely in the right hand side of (2.31). Indeed, different forms of this term are generally used in most of the literature on viscous Saint-Venant equations. We mention in this direction Kloeden [13], Sundbye [19], Gerbeau and Perthame [9] or the review paper Bresch [5], which consider, once written in dimension one and with our notation, the term

µ

∂x h∂x qh

in the right hand side of (2.31). However, other viscous terms have been considered in the literature, see, for instance, [5], Bernardi and Pironneau [3], Orenga [16] or Rodr´ıguez and Taboada-V´azquez [18]. Comparing the viscosity term we introduce with the more commonly one used used, for instance, in [9], shows that their difference is given by

µ ρ

∂x 1

h

∂q

∂x

− µ hρ

∂x

h ∂

∂x q

h

= µq ρh2

2h

∂x2 − µq ρh3

∂h

∂x 2

.

Passing to dimensionless variables and denoting byεthe size of the solution, the term in the right hand side of the above formula is of higher order inεthan the other terms in the momentum equa- tion. Therefore, within the global precision of the shallow water approximation, the two viscosity terms are equivalent. Our choice of the viscous term in the right hand side of (2.31) seems an appropriate one in the case of viscous shallow water model of a fluid interacting with a floating solid for at least two reasons:

• It is energetically consistent, see Proposition 3.1 below.

• In our case, the smoothing effect coming from the viscous term applies to the flux qinstead of the velocity field v as in [13] and [19] (note that in our case, unlike in the case of non viscous Saint Venant equations, qand v have different regularity properties). An advantage of our approach is that, since we use the variables handq we do not need any information on the sign of the velocity v at the solid-fluid interface.

(9)

3 Energy estimate and main result

By extending the pressure by 0 in the fluid part, by taking ρ= 1 and by adding boundary and initial conditions, we can rewrite system (2.28)–(2.33) as follows:

∂h

∂t +∂q

∂x = 0 (t >0, x∈ I ∪ E), (3.1)

∂t q

h

+ ∂

∂x q2

2h2 +gh+p

−µ ∂

∂x 1

h

∂q

∂x

= 0 (t >0, x∈ I ∪ E), (3.2) q(t,0) =q(t, `) = 0 (t >0), (3.3) p(t, x) = 0 (t>0, x∈ E), (3.4) h(t, x) =H(t) (t>0, x∈ I), (3.5) [q(t,·)]a = [q(t,·)]b= 0 (t >0), (3.6)

p(t,·) + q2(t, a)

2h2(t,·)+gh(t,·)−µ h

∂q

∂x(t,·)

a

= 0 (t>0), (3.7)

p(t,·) + q2(t, b)

2h2(t,·)+gh(t,·)−µ h

∂q

∂x(t,·)

b

= 0 (t>0), (3.8)

MH¨(t) =−M g+ Z b

a

p(t, x) dx (t >0), (3.9) h(0, x) =h0(x) (x∈ E), (3.10) q(0, x) =q0(x) (x∈[0, `]), (3.11)

H(0) =H0. (3.12)

Note that from (3.1) and (3.5) it follows that

∂q

∂x(t, x) =−H˙(t) (t>0, x∈ I). (3.13) In particular, the initial velocity of the solid is given by

H˙(0) =−∂q0

∂x.

Let us note that the spatial derivatives in the above equations and in the whole paper correspond to the spatial derivatives of the restrictions for each subdomain. In particular, we have the classical formula for any function f, smooth onE ∪ I:

Z ` 0

df

dx dx=f(`)−f(0)−[f]b−[f]a. (3.14) We first show that the system (3.1)–(3.12) satisfies an energy estimate:

Proposition 3.1. Let us write E(t) = 1

2 Z `

0

q2(t, x)

h(t, x) +gh2(t, x)

dx+1

2MH˙2(t) +M gH(t) (t>0) (3.15) and let us consider a smooth solution(h, q, H, p)of system(3.1)–(3.12). Then we have the following relation:

E(t) =˙ −µ Z `

0

1 h(t, x)

∂q

∂x(t, x) 2

dx (t>0). (3.16)

(10)

Proof. From (3.15), it follows that E(t) =˙

Z ` 0

1 2

q h

2∂h

∂t +q∂

∂t q

h

+gh∂h

∂t

dx+MH˙(t) ¨H(t) +M gH(t).˙ By combining the above formula with (3.1), (3.2) and (3.4) it follows that

E(t) =˙ − Z `

0

q2 2h2 +gh

∂q

∂x +q ∂

∂x q2

2h2 +gh

−qµ ∂

∂x 1

h

∂q

∂x

dx

− Z b

a

q(t, x)∂p

∂x(t, x) dx+MH˙(t) ¨H(t) +M gH(t)˙ (t>0). (3.17) Since

q2 2h2 +gh

∂q

∂x+q ∂

∂x q2

2h2 +gh

= ∂

∂x

q q2

2h2+gh

, we can combine (3.3), (3.6) and (3.17) to obtain

E(t) =˙ q(t, a)

q2(t, a)

2h2(t,·)+gh(t,·)

a

+q(t, b)

q2(t, b)

2h2(t,·)+gh(t,·)

b

+ Z `

0

qµ ∂

∂x 1

h

∂q

∂x

dx− Z b

a

q(t, x)∂p

∂x(t, x) dx+MH(t) ¨˙ H(t) +M gH˙(t) (t>0).

Integrating by parts in the two integrals in the above formula and using (3.7) and (3.8), together with (3.6), it follows that

E(t) =˙ −µ Z `

0

1 h

∂q

∂x 2

dx+ Z b

a

p(t, x)∂q

∂x(t, x) dx+MH(t) ¨˙ H(t) +M gH˙(t) (t>0).

Finally, inserting (3.13) and (3.9) in the last formula we obtain the desired energy estimate (3.16).

One can check that the stationary solutions of (3.1)–(3.9) can be parametrized by H > 0 by setting

h:=H+ M

b−a, p:= M g

b−a (3.18)

and in that case

hS(x) =

(h x∈ E

H x∈ I , qS(x) = 0 and pS(x) =

(0 x∈ E

p x∈ I . (3.19)

In the statements below, we also need the following notation: we writePk(I),k>0 the set of polynomial functions of degree6k. Let us state our main results:

Theorem 3.2. Let us assume that H0, h0, q0>

∈R×H1(E)×H1(0, `) with

H0>0, h0>0 in E, H0<min(h0(a), h0(b+)) (3.20)

q0(0) =q0(`) = 0. (3.21)

Let K >0be such that

|H0|+kh0kH1(E)+kq0kH1(0,`)6K, 1

K 6h0(x)6K forx∈ E. (3.22)

(11)

Then, there exists T >0, depending only onK such that the system (3.1)–(3.12) admits a unique strong solution

H∈H2(0, T), h∈H1(0, T;H1(E))∩C1([0, T];L2(E)), (3.23)

q∈C0([0, T];H1(0, `)) (3.24)

q|E ∈H1(0, T;L2(E))∩C0([0, T];H1(E))∩L2(0, T;H2(E)), (3.25)

q|I∈H1(0, T;P1(I)), (3.26)

p|I ∈L2(0, T;P2(I)). (3.27)

Moreover, there exists a constantKT >0 such that 1

KT 6h0(t, x), H(t)6KT, for all t∈(0, T), x∈ E, H(t)<min(h(t, a), h(t, b+)) for all t∈(0, T).

Theorem 3.3. For anyH >0, the system (3.1)–(3.12)is locally well-posed around the stationary state defined by (3.18)-(3.19). More precisely, there existsη0>0such that, for allη∈(0, η0)there exists two constants ε >0andC >0,such that, for any

H0, h0, q0>

∈R×H1(E)×H1(0, `) with

Z

E

h0(x) dx+H0(b−a) =M|E|

|I|+H`, q0(0) =q0(`) = 0, (3.28)

|H0−H|+kh0−hkH1(E)+kq0kH1(0,`)6ε, (3.29) there exists a unique solution of (3.1)–(3.12) with

H ∈H+H2(0,∞), h∈h+H1(0,∞;H1(E))∩Cb1([0,∞);L2(E)), (3.30)

q∈Cb([0,∞);H1(0, `)) (3.31)

q|E ∈H1(0,∞;L2(E))∩Cb([0,∞);H1(E))∩L2(0,∞;H2(E)), (3.32)

q|I∈H1(0,∞;P1(I)) (3.33)

p|I ∈ M g

|I| +L2(0,∞;P2(I)) (3.34)

Z

E

h(t, x) dx+H(t)(b−a) =M|E|

|I|+H` (t>0), (3.35) h(t, x)>h

2, H(t)>H

2 , for allt∈(0,∞), x∈ E, (3.36) H(t)<min(h(t, a), h(t, b+)) for all t∈(0,∞). (3.37) satisfying the estimate

keη(·) H−H

kH2(0,∞)+keη(·) h−h

kH1(0,∞;H1(E))+keη(·)qkL(0,∞;H1(0,`))

+keη(·)q|EkH1(0,∞;L2(E))∩L2(0,∞;H2(E))6Cε.

4 Equations (3.1)–(3.12) as a parabolic system on E

In this section we eliminate pressure from (3.1)–(3.12) and obtain an equivalent system involving the restrictions ofhandq toE, together with the traces of qat x=a, b andH. More precisely, let us set

qa(t) :=q(t, a) =q(t, a+), qb(t) :=q(t, b) =q(t, b+). (4.1)

(12)

From (3.13), we deduce

H˙ =−qb−qa

b−a inI, (4.2)

and

q(t, x) =qa(t)x−b a−b

+qb(t)x−a b−a

, ∂q

∂x(t, x) =qb(t)−qa(t)

b−a (t>0, x∈ I). (4.3) Combining (3.2), (3.5) and (3.13), we get

∂p

∂x =−H ∂

∂t q

H2

=−q˙

H + 2qH˙

H2 (t>0, x∈ I). (4.4)

We recall that iff ∈ P2([a, b]), then we have the following standard formulas Z b

a

f(x)dx=f(a)(b−a) +f0(a)(b−a)2

3 +f0(b)(b−a)2

6 , (4.5)

Z b a

f(x)dx=f(b)(b−a)−f0(a)(b−a)2

6 −f0(b)(b−a)2

3 . (4.6)

In particular, since pis a second degree polynomial with respect tox, we deduce Z b

a

p(t, x)dx

=p(t, a)(b−a) + −q˙a(t)

H(t)+ 2qa(t) ˙H(t) H(t)2

!(b−a)2

3 + −q˙b(t)

H(t)+ 2qb(t) ˙H(t) H(t)2

!(b−a)2

6 (4.7)

and Z b

a

p(t, x)dx

=p(t, b)(b−a)− −q˙a(t)

H(t)+ 2qa(t) ˙H(t) H(t)2

!(b−a)2

6 − −q˙b(t)

H(t)+ 2qb(t) ˙H(t) H2

!(b−a)2

3 . (4.8) Combining (4.7), (4.8), (4.2) and (3.9) yields

h

M +(b−a)3 3H

i

˙ qa−h

M −(b−a)3 6H

i

˙

qb=−M g(b−a) +p(·, a)(b−a)2 +

"

2qa2−qaqb−qb2

#(b−a)2

3H2 (t >0), (4.9) and

−h

M −(b−a)3 6H

i

˙ qa+h

M+(b−a)3 3H

i

˙

qb=M g(b−a)−p(·, b)(b−a)2

"

2q2b−qaqb−q2a

#(b−a)2

3H2 (t >0). (4.10) Inverting the above linear system, we get

a

˙ qb

=S(H)

−M g(b−a) +p(·, a)(b−a)2+h

2q2a−qaqb−q2bi(b−a)2

3H2

M g(b−a)−p(·, b)(b−a)2−h

2qb2−qaqb−qa2i(b−a)2

3H2

, (4.11)

(13)

where for allH >0,S(H) is the following symmetric positive matrix:

S(H) := H

(b−a)3

M +(b−a)12H3

"

M +(b−a)3H3 M −(b−a)6H3 M −(b−a)6H3 M +(b−a)3H3

#

. (4.12)

Considering equations (3.7)–(3.8) together with (4.3) and with (4.11), we deduce, we obtain:

a

˙ qb

=R

H, q(·, a), ∂

∂xq(·, a), h(·, a), q(·, b), ∂

∂xq(·, b+), h(·, b+)

, (4.13)

where R

H, q(·, a), ∂

∂xq(·, a), h(·, a), q(·, b), ∂

∂xq(·, b+), h(·, b+)

=S(H)

−M g(b−a) +p(·, a)(b−a)2+h

2qa2−qaqb−qb2i(b−a)2

3H2

M g(b−a)−p(·, b)(b−a)2−h

2qb2−qaqb−q2ai(b−a)2

3H2

, (4.14) andp(·, a) andp(·, b) are given by:

p(t, a) = q2(t, a)

2h2(t, a)+gh(t, a)−µ h

∂q

∂x(t, a)−q2(t, a)

2H2(t)−gH(t) +µqb(t)−qa(t)

H(b−a) (4.15) p(t, b) = q2(t, b)

2h2(t, b+)+gh(t, b+)−µ h

∂q

∂x(t, b+)−q2(t, b)

2H2(t)−gH(t) +µqb(t)−qa(t)

H(b−a) . (4.16) Finally, the system (3.1)–(3.12) writes in the equivalent form

H˙ =−qb−qa

b−a (t>0), (4.17)

∂h

∂t(t, x) + ∂q

∂x(t, x) = 0 (t >0, x∈ E), (4.18)

∂q

∂t + ∂

∂x q2

h +gh2 2

=h∂

∂x µ

h

∂q

∂x

(t >0, x∈ E), (4.19)

q(t,0) =q(t, `) = 0 (t >0), (4.20)

q(t, a) =qa(t), q(t, b) =qb(t) (t >0), (4.21) q˙a

˙ qb

=R

H, q(·, a), ∂

∂xq(·, a), h(·, a), q(·, b), ∂

∂xq(·, b+), h(·, b+)

, (4.22)

withR defined by (4.14), (4.15) and (4.16).

5 Local in time existence and uniqueness

In this section we prove Theorem3.2. In view of the reduction performed in Section 4, our main result in Theorem3.2can be restated as:

(14)

Theorem 5.1. Assume that

H0, h0, q0, qa,0, qb,0>

∈R×H1(E)×H1(E)×R2, with

H0>0, h0>0 in E, H0<min(h0(a), h0(b+)), (5.1) q0(0) =q0(`) = 0, q0(a) =qa,0, q0(b) =qb,0. (5.2) Let K >0be such that

|H0|+kh0kH1(E)+kq0kH1(E)+|q0,a|+|q0,b|6K, 1

K 6h0(x)6K forx∈ E. (5.3) Then, there existsT >0,depending only onK, such that the system (4.17)–(4.22)admits a unique strong solution

H ∈H2(0, T), h∈H1(0, T;H1(E))∩C1([0, T);L2(E)), (5.4) q∈H1(0, T;L2(E))∩C([0, T);H1(E))∩L2(0, T;H2(E)), (5.5) qa∈H1(0, T), qb∈H1(0, T) (5.6) Moreover, there exists a constantKT >0 such that

1

KT 6h(t, x), H(t)6KT (t∈(0, T), x∈ E), H(t)<min(h(t, a), h(t, b+)) (t∈(0, T)).

The proof of the above theorem relies on classical estimates on linear parabolic problems, com- bined with the use of the Banach fixed point theorem. The strategy we adopt here is based on the fact that the system (4.17)–(4.22) can be rewritten as

H˙ =−qb−qa

b−a (t >0), (5.7)

∂h

∂t + ∂q

∂x = 0 (t >0, x∈ E), (5.8)

∂q

∂t −µ∂2q

∂x2 =F1(h, q) (t >0, x∈ E), (5.9)

q(t,0) =q(t, `) = 0, (t >0), (5.10)

q(t, a) =qa(t), q(t, b) =qb(t) (t >0), (5.11) q˙a

˙ qb

=F2(H, q, qa, qb) (t >0), (5.12)

h(0, x) =h0(x), q(0, x) =q0(x) (x∈ E), (5.13) H(0) =H0, qa(0) =qa,0, qb(0) =qb,0, (5.14) where

F1(h, q) =−2q h

∂q

∂x+ q2 h2

∂h

∂x+gh∂h

∂x −µ h

∂h

∂x

∂q

∂x, (5.15)

and

F2(H, q, qa, qb) =R

H, q(·, a), ∂

∂xq(·, a), h(·, a), q(·, b), ∂

∂xq(·, b+), h(·, b+)

, (5.16) with R defined by (4.14), (4.15) and (4.16). Next, by replacingF1 and F2 by given source term f1 andf2we obtain the following linear system :

(15)

H˙ =−qb−qa

b−a (t >0), (5.17)

∂h

∂t + ∂q

∂x = 0 (t >0, x∈ E), (5.18)

∂q

∂t −µ∂2q

∂x2 =f1 (t >0, x∈ E), (5.19)

q(t,0) =q(t, `) = 0, (t >0), (5.20)

q(t, a) =qa(t), q(t, b) =qb(t) (t >0), (5.21) q˙a

˙ qb

=f2 (t >0), (5.22)

h(0, x) =h0(x), q(0, x) =q0(x) (x∈ E), (5.23) H(0) =H0, qa(0) =qa,0, qb(0) =qb,0. (5.24) We have the following regularity result for the system (5.17)-(5.24)

Theorem 5.2. Let us assume that

H0, h0, q0, qa,0, qb,0>

∈R×H1(E)×H1(E)×R2and satisfies the compatibility condition (5.2). Then, for all0< T 61, f1∈L2(0, T;L2(E))and f2∈L2(0, T;R2) the system (5.17)-(5.24)admits a unique solution

H ∈H2(0, T), h∈H1(0, T;H1(E))∩C1([0, T);L2(E)), q∈H1(0, T;L2(E))∩C([0, T);H1(E))∩L2(0, T;H2(E)),

qa∈H1(0, T), qb∈H1(0, T).

Moreover, there exists a constantC independent of T such that

kHkH2(0,T)+khkH1(0,T;H1(E))+khkL(0,T;H1(E))+kqkL2(0,T;H2(E))+kqkH1(0,T;L2(E))

+kqkL(0,T;H1(E))+k(qa, qb)kH1(0,T;R2)6C

|H0|+keh0kH1(E)+kq0kH1(E)

+|qa,0|+|qb,0|+

(f1, f2)

L2(0,T;L2(E)×

R2)

, (5.25) for all0< T 61.

Proof. Let us remark that the linear system (5.17)-(5.24) can be solved in “cascades”: Eq. (5.22) can be solved independently and admits a unique solution (qa, qb)∈H1(0, T;R2). Next we solve Eq. (5.17) and we obtainH ∈H2(0, T). Using the standard regularity results for parabolic equa- tions with non-homogeneous boundary conditions, we also have that

q∈L2(0, T;H2(E))∩H1(0, T;L2(E))∩L(0, T;H1(E)).

Finally, using the regularity ofq,we obtain the desired regularity ofhfrom Eq. (5.18). To obtain estimates with continuity constant independent of time T we can proceed as in [10, Theorem 5.3].

Now we estimate the nonlinear terms defined in (5.15)-(5.16).

Proposition 5.3. Let us assume that

H0, h0, q0, qa,0, qb,0>

∈R×H1(E)×H1(E)×R2 such that (5.1)-(5.3)holds. There existTe∈(0,1),δ >0 andC=C(M,Te)>0such that for T ∈(0,Te]and for(f1, f2)∈L2(0, T;L2(E)×R2)satisfying

(f1, f2)

L2(0,T;L2(E)×R2)61, the solution(H, h, q, qa, qb)of (5.17)-(5.24)verifies

F1

L2(0,T;L2(E))+ F2

L2(0,T;

R2)6CTδ.

(16)

Proof. We chooseT <e 1 andT ∈(0,Te].The constant appearing in this proof depends only onK and independent ofT.First of all, from (5.25) we first obtain

kHkH2(0,T)+khkH1(0,T;H1(E))+khkL(0,T;H1(E))+kqkL2(0,T;H2(E))+kqkH1(0,T;L2(E))

+kqkL(0,T;H1(E))+k(qa, qb)kH1(0,T;R2)6C. (5.26) Applying H¨older’s inequality together with the above estimate we deduce that

kH−H0kL(0,T)6C

T , kh−h0kL(0,T;H1(E)) 6C

√ T . In particular, there existsTesuch that for allT ∈(0,Te] the following holds:

H(t)> H0

2 , h(t, x)> 1

2K, t∈(0, T), x∈ E, (5.27)

khkL((0,T)×E)+kHkL(0,T)6C, (5.28)

khkL2(0,T;H1(E))+ kHkL2(0,T)6C√

T , (5.29)

1 h

L(0,T;H1(E)) +

1 h

L((0,T)×E) +

1 H

L(0,T)6C. (5.30)

In a similar manner, we also obtain

kqakL(0,T)+kqbkL(0,T)6C, kqakL2(0,T)+kqbkL2(0,T)6C√

T . (5.31)

On the other hand, using an interpolation inequality one has following estimate (see, for instance estimate, (6.13) of [10])

kqkL2(0,T;H1+s(E)) 6CT(1−s)/4, s∈(0,1). (5.32) Using the above estimate, (5.26) and the Sobolev embedding we get

kqkL((0,T)×E)6C,

∂q

∂x

L2(0,T;L(E))

6CT(1−s)/4, s∈(1/2,1) (5.33) We are now in a position to estimate the non-linear terms defined in (5.15)-(5.16).

Estimate of F1:

kF1kL2(0,T;L2(E)) 6CTδ, for some δ >0. (5.34)

•Estimate of the first term ofF1: Using (5.26), (5.27) and (5.33), we have

−2q h

∂q

∂x

L2(0,T;L2(E))

6CkqkL2(0,T;H1(E))6C√

TkqkL(0,T;H1(E))6C√ T .

•Estimate of the second and third term ofF1: Using (5.26), (5.27) and (5.33) we have

q2 h2

∂h

∂x +gh∂h

∂x

L2(0,T;L2

(E))

6CkhkL2(0,T;H1(E))6C√ T .

•Estimate of the last term ofF1: Using (5.26), (5.27) and (5.33) we obtain

µ h

∂h

∂x

∂q

∂x,

L2(0,T;L2

(E))

6C

∂h

∂x

L(0,T;L2

(E))

∂q

∂x

L2(0,T;L

(E))

6CT(1−s)/4, for somes∈(1/2,1).Thus we have proved (5.34).

Estimate of F2:

kF2kL2(0,T;R2)6CTδ, for someδ >0. (5.35)

(17)

•Estimate ofS(H) defined in (4.12) : Using (5.27) - (5.29), we have the following estimates kS(H)kL(0,T;R4)6C, kS(H)kL2(0,T;R4)6C√

T . (5.36)

•Estimatep(t, a) andp(t, b) defined in (4.15) - (4.16): We claim that

kp(·, a)kL2(0,T)+kp(·, b)kL2(0,T)6CTδ, for some δ >0. (5.37) We provide only the estimate of third term of p(t, a). Estimates of the other terms are similar.

Using (5.30) and (5.32) we have

µ h

∂q

∂x(·, a)

L2(0,T)6C 1 h

∂q

∂x

L2(0,T;Hs(0,a))6C 1 h

L(0,T;H1(0,a))

∂q

∂x

L2(0,T;Hs(0,a))6CT(1−s)/4, for somes∈(1/2,1).Thus we have (5.37). Finally, combining (5.36), (5.37) and (5.31), we obtain (5.35).

We are now in a position to prove one of Theorem5.1.

Proof of Theorem 5.1. LetTebe the constant in Proposition 5.3. For T ∈(0,Te], we consider the ball

BT :=

(f1, f2)∈L2(0, T;L2(E)×R2) ; (f1, f2)

L2(0,T;L2

(E)×R2)61

(5.38) and the map

Ξ : (f1, f2)∈BT 7→(F1(h, q),F2(H, q, qa, qb)), (5.39) where (H, h, q, qa, qb) is the solution to the system (5.17)-(5.24) and whereF1andF2are given by (5.15)-(5.16). By Proposition 5.3, we have that Ξ(BT) ⊂BT for T small enough. With similar calculation as Proposition5.3, we can also show that for smallT, Ξ|BT is a strict contraction. This completes the proof of Theorem5.1.

6 Global in time existence and uniqueness

In this section, we linearize the system (4.17)–(4.22) around a stationary state and we study the corresponding linear system. Note that, in order to prove global existence for small data of our original problem (4.17)–(4.22) we need this linearization to be exponentially stable, so that we cannot use the linearized problem (5.17)-(5.24), introduced for the local existence result.

6.1 Linearization around a stationary state

We consider the stationary state given by (3.19) and we define

He :=H−H, eh:=h−hS =

(h−h inE He inI . From (4.15)–(4.16) and (3.18), we deduce

p(t, a) = q2(t, a) 2

1

(eh+h)2(t, a)− 1 (He+H)2(t)

!

+g(eh(t, a)−He(t)) +g M b−a

−µ h

∂q

∂x(t, a) +µqb(t)−qa(t) H(b−a) +µ

h eh (eh+h)

∂q

∂x

!

(t, a)−µqb(t)−qa(t) b−a

He

(He+H)H (6.1)

(18)

p(t, b) =q2(t, b) 2

1

(eh+h)2(t, b+)− 1 (He +H)2(t)

!

+g(eh(t, b+)−He(t)) +g M b−a

−µ h

∂q

∂x(t, b+) +µqb(t)−qa(t) H(b−a) +µ

h eh (eh+h)

∂q

∂x

!

(t, b+)−µqb(t)−qa(t) b−a

He

(He+H)H. (6.2) Thus

−M g(b−a) +p(·, a)(b−a)2+h

2q2a−qaqb−q2bi(b−a)2 3H2

= (b−a)2 (

g(eh(·, a)−He)−µ h

∂q

∂x(·, a) +µ(qb−qa) H(b−a) +q2(·, a)

2

1

(eh+h)2(·, a)− 1 (He+H)2

!

+µ h

eh (eh+h)

∂q

∂x

!

(t, a)−µqb−qa

b−a He

(He+H)H +2qa2−qaqb−qb2 3(He +H)2

) (6.3) and

M g(b−a)−p(·, b)(b−a)2−h

2qa2−qaqb−qb2i(b−a)2 3H2

= (b−a)2 (

−g(eh(·, b+)−H) +e µ h

∂q

∂x(·, a)−µ(qb−qa) H(b−a)

−q2(·, b) 2

1

(eh+h)2(·, b+)− 1 (He +H)2

!

+µ h

eh (eh+h)

∂q

∂x

!

(t, b+) +µqb−qa

b−a He

(He+H)H −2qb2−qaqb−qa2 3(He +H)2

) (6.4)

Then (H,e eh, q, qa, qb) satisfies

˙

He =−qb−qa

b−a (t>0), (6.5)

∂eh

∂t(t, x) +∂q

∂x(t, x) = 0 (t >0, x∈ E), (6.6)

∂q

∂t +gh∂eh

∂x −µ∂2q

∂x2 =F1(eh, q) (t >0, x∈ E), (6.7)

q(t,0) =q(t, `) = 0 (t >0), (6.8)

q(t, a) =qa(t), q(t, b) =qb(t) (t >0), (6.9)

a

˙ qb

= (b−a)2S(H)

g(eh(·, a)−H)e −µ h

∂q

∂x(·, a) + µ H

q(·, b)−q(·, a) b−a

−g(eh(·, b+)−H) +e µ h

∂q

∂x(·, b+)− µ H

q(·, b)−q(·, a) b−a

+F2(H,e eh, q, qb, qa) (6.10) where

F1(eh, q) :=− ∂

∂x q2

eh+h+geh2 2

!

− µ

(h+eh)2

∂eh

∂x

∂q

∂x, (6.11)

(19)

F2(H,e eh, q, qb, qa) := (b−a)2

S(He +H)−S(H)

g(eh(·, a)−H)e −µ h

∂q

∂x(·, a) + µ H

qb−qa

b−a

−g(eh(·, b+)−H) +e µ h

∂q

∂x(·, a)− µ H

qb−qa

b−a

+ (b−a)2S(He+H)

"

F3(H,e eh, q, qb, qa) F4(H,e eh, q, qb, qa)

#

, (6.12)

F3(H,e eh, q, qb, qa) :=qa2 2

1

(eh+h)2(·, a)− 1 (He+H)2

! +µ

h eh (eh+h)

∂q

∂x

! (·, a)

−µqb−qa

b−a

He

(He +H)H +2q2a−qaqb−q2b

3(He+H)2 (6.13)

F4(H,e eh, q, qb, qa) :=−q2b 2

1

(eh+h)2(·, b+)− 1 (He+H)2

!

−µ h

eh (eh+h)

∂q

∂x

! (·, b+)

+µqb−qa

b−a

He

(He +H)H −2q2b−qaqb−q2a

3(He+H)2 (6.14) Using (3.10), (3.11) and (3.12), we add the following initial conditions to system (6.5)–(6.14):

eh(0, x) =eh0(x) (x∈ E), (6.15)

q(0, x) =q0(x) (x∈ E), (6.16)

He(0) =He0, qa(0) =qa,0 qb(0) =qb,0, (6.17) where

He0=H0−H, eh0:=h0−h, qa,0:=q0(a), qb,0:=q0(b).

In order to study the above equations, we consider the following linear system

˙

He =−qb−qa

b−a (t>0), (6.18)

∂eh

∂t(t, x) + ∂q

∂x(t, x) = 0 (t >0, x∈ E), (6.19)

∂q

∂t +gh∂eh

∂x−µ ∂

∂x ∂q

∂x

=f1 (t >0, x∈ E), (6.20)

q(t,0) =q(t, `) = 0 (t >0), (6.21) q(t, a) =qa(t), q(t, b) =qb(t) (t >0), (6.22)

a

˙ qb

=S0(H)

 µ H

q(·, b)−q(·, a) b−a +g

eh(·, a)−He

−µ h

∂q

∂x(·, a)

−µ H

q(·, b)−q(·, a) b−a −g

eh(·, b+)−He +µ

h

∂q

∂x(·, b+)

+f2 (6.23) where

S0(H) = H

(b−a)

M+(b−a)12H3

"

M+(b−a)3H3 M−(b−a)6H3 M−(b−a)6H3 M+(b−a)3H3

#

, (6.24)

and wheref1andf2are now given.

(20)

6.2 Study of the Linear Problem

In this section we use the methodology developed in [15] in order to prove that the linear problem introduced above can be described by an analytic C0 semigroup in an appropriate Hilbert space and then we show that this semigroup is exponentially stable. To this end we begin by defining the following spaces:

X =n

[H, h, q]>∈C×H1(E)×L2(E) | Z

E

h0(x) dx+H0(b−a) = 0}, U =C2, (6.25)

Z=n

[H, h, q]>∈C×H1(E)×H2(E) | Z

E

h0(x) dx+H0(b−a) = 0, q0(0) =q0(`) = 0o . (6.26) We also introduce the following operators:

L:D(L)→X, L

 H

h q

=

q(b)−q(a)b−a

dqdx

−ghdhdxdxd2q2

, D(L) =Z, (6.27)

G:D(G)→ U, G

 H

h q

= q(a)

q(b)

, D(G) =Z, (6.28) and

A:=L|Ker(G). (6.29)

The above definition yields

D(A) =C×H1(E)×(H2∩H01)(E) and A

 H

h q

=

0

dqdx

−ghdhdxdxd2q2

. (6.30)

Lemma 6.1. The operator A:D(A)→X generates an analytic semigroup.

Proof. It clearly suffices to prove that the operatorAedefined by D(A) =e H1(E)×(H2∩H01)(E),

Ae h

q

=

"

dqdx

−ghdhdxdxd2q2

#

h q

∈ D(A)e

,

generates an analytic semigroup in Xe =H1(E)×L2(E). To this aim, noticing that the operator defined by

P h

q

=

"

0

−ghdhdx

#

is in L(X), it suffices to prove that the operatore A1 defined by D(A1) =D(A),e A1

h q

=

"

dxdq µdxd2q2

#

h q

∈ D(A)e

,

Références

Documents relatifs

The paper deals with projection estimators of the density of the stationary solution X to a differential equation driven by the fractional Brownian motion under a

Note nally that for the nonlinear Fokker{Planck equation (also called Landau equation) (Cf. [Lif, Pi], [De, Vi]), no rigorous deriva- tion from an N-particle system exists, even in

[28] ——— , “Computation of the difference-differential Galois group and differential relations among solutions for a second-order linear difference equation”, Commun.. Singer

We deal with the case where the fluid equations are the non stationary Stokes system and using the enclosure method, we can recover the convex hull of the obstacle and the distance

We show hereafter that the dynamic programming approach, already implemented by Fabbri and Gozzi (2008) in the delayed case, provides a much finer characterization of the

Tucsnak, Existence of solutions for the equations modelling the motion of a rigid body in a viscous fluid, Comm.. Coron, On the controllability of 2-D incompressible perfect

The literature reveals four major uses of the BM concept: to map specific business sectors or industries and create taxonomies; to describe the

Our result can be contrasted with the result from [13] (see also [14] for a gentle exposition) regarding the controllability of the fluid velocity alone in the case of