• Aucun résultat trouvé

Linearization of a coupled system of nonlinear elasticity and fluid

N/A
N/A
Protected

Academic year: 2021

Partager "Linearization of a coupled system of nonlinear elasticity and fluid"

Copied!
19
0
0

Texte intégral

(1)

HAL Id: inria-00442954

https://hal.inria.fr/inria-00442954

Submitted on 24 Dec 2009

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.

Linearization of a coupled system of nonlinear elasticity

and fluid

Lorena Bociu, Jean-Paul Zolesio

To cite this version:

Lorena Bociu, Jean-Paul Zolesio. Linearization of a coupled system of nonlinear elasticity and fluid. [Research Report] Inria. 2009. �inria-00442954�

(2)

NONLINEAR ELASTICITY AND FLUID

LORENA BOCIU AND JEAN-PAUL ZOL´ESIO

Abstract. We model the coupled system formed by an incompressible, irrotational fluid and a nonlinear elastic body. We work with large dis-placement, small deformation elasticity (or St Venant elasticity), which makes the problem very interesting from the physical point of view. The elastic body is three-dimensional Ω ∈ R3, and thus it can not be reduced

to its boundary Γ (like in the case of a membrane or a shell). In this paper, we study the static problem, which contrary to common belief, it is more subtle than the dynamical one (since in real life, evolution is more plausible than equilibrium).

1. Introduction

1.1. The model and the problem. We consider a model of fluid-structure interaction on a bounded domain D ∈ R3. We assume that D is comprised

of two open domains D = Ω ∪ ΩC, and has smooth boundary ∂D = Γ′

∪ Γin∪

Γout. The elastic body occupies domain Ω with sufficiently smooth boundary

Γ ∪ Γ′

, and is described by a nonlinear elastic equation in terms of the displacement u. The fluid occupies domain ΩC with boundary Γ ∪ Γin∪ Γout,

and is described by a Navier-Stokes equation in terms of the velocity of the fluid v and the pressure p. The interaction takes place on the common boundary Γ and is realized via suitable transmission boundary conditions. We assume that there is a flux ~f coming into D through Γin, that will

determine the velocity of the fluid v (see Figure 1).

One specific example of the above mentioned model is a 3D tube with elastic walls through which a fluid is flowing. From the physical point of view, this is a very important model with a lot of applications in mathematical biology, more precisely, the study of arterial diseases (the tube represents the artery, the elastic body is the wall of the artery and the fluid is the blood).

Date: December 24, 2009.

2010 Mathematics Subject Classification. Primary: 35A01, Secondary:

Key words and phrases. nonlinear elasticity, Navier-Stokes, linearization, coupled sys-tem.

The research of Lorena Bociu was supported by the National Science Foundation under International Research Fellowship OISE-0802187 .

(3)

Γ′ Γ′ Γin Γout Γ Γ Ω Ω ΩC ~ f →

We begin the description of our model by introducing the notation and the basic assumptions on the two equations present in the system.

• Nonlinear, 3D elasticity: We work with large displacement, small deformation elasticity (or St Venant elasticity [11]), which makes the problem difficult from the mathematical point of view, and very interesting from the physical point of view. The elastic body is three-dimensional Ω ∈ R3, and thus it can not be reduced to its boundary Γ (like in the case of a membrane or a shell).

At rest, the elastic body occupies a reference configuration O ∈ R3, where O is a bounded, open, connected set in R3 with sufficiently smooth

bound-ary S ∪ Γ′

. When subjected to applied forces, the elastic body occupies a deformed configuration Ω = ϕ(O), with smooth boundary Γ ∪ Γ′

(where Γ′

is fixed). The deformation of the reference configurations is given by the map ϕ : O → R3, that is smooth enough, injective (except possibly on the boundary of the set O), and orientation-preserving (i.e. det∇ϕ(x) > 0, for all x ∈ O).

Together with the deformation ϕ, we introduce the displacement u : O → R3, defined as usual as ϕ = I + u, where I denotes identity map I : O → R3. It is well known that a body occupying a deformed configuration Ω, and subjected to zero applied body forces in its interior Ω and to applied surface forces on the boundary Γ, is in static equilibrium if the fundamental stress principle of Euler and Cauchy is satisfied:

(

−divT = 0 in Ω

T nϕ = gϕ on Γ (1.1)

where ˆg represents the density of the applied surface force, nϕ is the unit outer normal vector along Γ, and the tensor T is the Cauchy stress tensor. The above equilibrium equations over Ω are equivalent to the equilibrium

(4)

equations over the reference configuration O: (

−divP = 0 in O

Pn = g on S (1.2)

where n denotes the unit outer normal vector along S, gda = ˆgdaϕ, and P : O → M3 is the Piola transform of the Cauchy stress tensor field, defined

by

P(x) = T (xϕ)Cof∇T (x) = det(∇ϕ(x))T (xϕ)(∇ϕ)−∗ (1.3) From the constitutive equations, we have that P(x) = ∇ϕ(x)Σ(u(x)), where Σ defines the second Piola-Kirchhoff stress tensor. In terms of the displacement u, Σ is given by

Σ(σ(u)) = λ(trσ(u))I + 2µσ(u) (1.4)

where λ and µ are the Lame constants of the material, and the Green-St Venant strain tensor σ(u) is given by

σ(u) = 1 2(Du

+ Du + Du∗Du) (1.5)

Therefore equations (1.2) can be rewritten as (

−div[(I + ∇u)Σ(σ(u))] = 0 in O

(I + ∇u)Σ(σ(u))n = g on S (1.6)

The advantage of the equilibrium equations over the reference configura-tion (1.2) or (1.6) over (1.1) is the fact that they are written in terms of the Lagrange variable x that is attached to the reference configuration, instead of the Euler variable xϕ= ϕ(x), which is precisely one of the unknowns.

Nevertheless, we want to stress the fact that equations (1.1) play a critical role when dealing with elastic body - fluid systems, where the coupling is taking place on the boundary interface between the two media. This interface is precisely the boundary Γ of the deformed configuration of the elastic body Ω and thus the coupling requires the continuity of the velocities and the normal stress tensors across Γ. Therefore, we need a relationship between the Cauchy stress tensor T and the strain tensor σ(u), that will provide us with the correct matching of the two dynamics on the common interface.

Recalling the relations between P, T , and Σ(u) we obtain that

T = 1

det(∇ϕ)∇ϕ · Σ(σ(u)) · (∇ϕ)

∗

◦ ϕ−1 (1.7)

• Potential fluid: We assume that the fluid present in the system is incompressible and irrotational. This means that the fluid is a potential fluid v= ∇φ, where v is the velocity of the fluid and φ (the velocity potential of the fluid) satisfies the Laplace equation ∆φ = 0 (due to the incompressibility condition which translates into ∇ · u = 0).

(5)

Now if p represents the pressure of the fluid and η the viscosity, then the flow is described by the following Navier-Stokes equation

v· ∇v − η∆v + ∇p = ρ~g (1.8)

where ρ is the density, and ~g is the gravitational acceleration. Due to the fluid being irrotational (vorticity curl v = 0), the convective acceleration reduces to v · ∇v = ∇kvk

2

2 

and thus (1.8) becomes

∇(1 2k∇φk

2+ p − ρgz) = 0

This provides us with the formula for the pressure p of the fluid:

p= c +1 2k∇φk

2− ρgz (1.9)

1.2. Fluid-structure interaction: the mathematical model. Now we couple the two dynamics described above and we require continuity of both the velocities and the normal stress tensors across the common boundary Γ. Recall that D = Ω ∪ ΩC, with boundary ∂D = Γ′ ∪ Γin∪ Γout, the elastic

body occupies domain Ω, and the fluid occupies ΩC (see Figure 1). We obtain the following PDE model in variables (φ, u):

               ∆φ = 0 ,ΩC −div(T ) = 0 ,Ω ∇φ · n = 0 ,Γ T · n = (c +12k∇φk2− ρgz)n ,Γ u= 0 ,Γ′ (1.10)

where n is the unit outer normal vector along Γ, and f is the flux coming in D through Γin. Recall that T is the Cauchy stress tensor, given by

T = 1

det(∇ϕ)∇ϕ · Σ(σ(u)) · (∇ϕ)

∗

◦ ϕ−1

Thus the two equations on Γ present in (1.10) are equivalent to ( ∇φ · n = 0 , on Γ 1 det(∇ϕ)∇ϕ · Σ(σ(u)) · (∇ϕ) ∗ · n ◦ ϕ = [(c + 12k∇φk2− ρgz)n] ◦ ϕ , on Γ (1.11)

1.2.1. Boundary conditions on Γin. Now we describe the flow (which so far

we called f ) coming into the domain through Γin. Let c(x) be a given,

smooth function defined on Γin such that

(

c(x) = 0 on ∂Γin, ∂

∂nφ= c(x) on Γin

(6)

Then it follows that 0 = Z Ωc div(∇φ)dx = Z Γin ∂φ ∂nin dΓin+ Z Γout ∂φ ∂nout dΓout (1.13)

At this point, we choose α ∈ R verifying      α= ∂n∂φ out on Γout, R Γoutα dΓout= − R Γinc(x) dΓin (1.14)

2. Description of the Static Problem

2.1. Parameter of Perturbation s. As previously mentioned, in this pa-per we are interested in the static problem associated with system (1.10). In a subsequent paper, we will consider the dynamical case. Contrary to common belief, the steady problem is more subtle than the dynamical one, since in real life, evolution is more plausible than equilibrium.

We assume that the flux entering the domain D is dependent on a param-eter of perturbation s, i.e. c(x) is a given, smooth function defined on Γin

such that ( c(x) = 0 on ∂Γin, ∂ ∂nφs = (1 + s)c(x) on Γin (2.1)

Then it follows that

0 = Z Ωc s div(∇φs)dx = Z Γin ∂φs ∂nin dΓin+ Z Γout ∂φs ∂nout dΓout (2.2)

For any s ≥ 0, we choose αs∈ R verifying

     αs = ∂n∂φouts on Γout, R Γoutα dΓout= −(1 + s) R Γinc(x) dΓin, for all s ≥ 0 (2.3)

If the elastic body occupies a reference configuration O ∈ R3 with smooth

boundary S ∪ Γ′

, then, when subjected to applied forces, it occupies a de-formed configuration Ωs= ϕs(O), with smooth boundary Γs∪ Γ′ (where Γ′

is fixed). The deformation map in this case is dependant on the parameter s: ϕs: O → R3, but nevertheless is smooth enough, injective, and

orientation-preserving. The displacement us: O → R3 becomes us= ϕs− I, where I is

the identity map I : O → R3.

Similarly, for the fluid present in the system, its velocity and pressure are now functions of s: vs= ∇φs, and ps= cs+12k∇φsk2− ρgz.

(7)

Therefore we are interested in the following PDE model:                    ∆φs = 0 ,ΩCs −div(Ts) = 0 ,Ωs ∇φs· ns= 0 ,Γs Ts· ns= (cs+ 21k∇φsk2− ρgz)n ,Γs u= 0 ,Γ′ R Γoutα dΓout= −(1 + s) R Γinc(x) dΓin, for all s ≥ 0 (2.4)

where ns is the unit outer normal vector along Γs, Ts is the Cauchy stress

tensor (associated to s), given by

T = 1

det(∇ϕs)

∇ϕs· Σ(σ(us)) · (∇ϕs)∗



◦ ϕ−s1 (2.5)

and the boundary conditions are equivalent to ( ∇φs· ns= 0 , on Γs 1 det(∇ϕs)∇ϕs· Σ(σ(us)) · (∇ϕs) ∗ · ns◦ ϕs= [(cs+12k∇φsk2− ρgz)ns] ◦ ϕs , on Γs (2.6) Our goal is to study, for any given value of s, the equilibrium problem associated with the coupled system (2.4).

2.2. The moving boundary Γs. Recall that the deformation map ϕ maps

the reference boundary S to Γ. Similarly, the deformation ϕs associated

with ˆf = f + s maps S to Γs.

At this point it is convenient to introduce the map Ts: Γ → Γsthat builds

the moving boundary Γs:

Ts= ϕs◦ ϕ−1 (2.7)

and the speed V (s, ·) associated with the flow mapping Ts:

V(s, ·) =∂ ∂s

Ts◦ Ts−1 = ∂ ∂s

ϕs◦ ϕ−s1 (2.8)

This means that Ts(V ) : X → x(s), where x(s) satisfies the following

differential equation ( ∂ ∂sx= V (s, x(s)) x(0) = X (2.9) which is equivalent to x(s) = X + Z s 0 V(t, x(t))dt.

2.2.1. Transport of scalar operators. Let D be a bounded domain in RN and T be a one-to-one transformation from D onto D and from ¯D onto

¯

D. Let S = T−1 be the inverse mapping. As T ◦ S = I, we obtain that

DT◦ S.DS = I, and therefore

DS= (DT )−1◦ S, and

(8)

For any φ ∈ H1(D) we have (∇φ) ◦ T = (DT )−∗.∇(φ ◦ T ) We also have: d dsTs= V (s) ◦ Ts⇒ d dsTs s=0= V (0). d dsDTs(X) = DV (s, Ts(X))DTs(X), DT0(X) = I ⇒ d dsDTs s=0= DV (0) and d ds(DTs) −1 s=0= −DV (0). d

dsdetDTs(X) = trDV (s, TS(X))detDTs(X) = divV (s, Ts(X))detDTs(X),

⇒ d dsdet(DTS) s=0= divV (0).

Now let ~Ebe a C1 vector field defined over D. Then we have the following property:

Property 2.1.

(divE) ◦ T = det(DT )−1 div( det(DT ) (DT )−1.(E ◦ T ) ) Proof. Let φ ∈ C∞

comp(D). Using the change of variable y = T (x) (or x =

S(y)), we obtain: Z D (divE) ◦ T (x) φ(x) dx = Z D

divE(y) φ ◦ S(y) det(DS)(y) dy

= − Z

D

< E(y), ∇( φ ◦ S(y) det(DS)(y) ) > dy

= − Z

D

< E(y), ∇( φ ◦ S(y) ) det(DS)(y) ) + φ ◦ S(y) ∇( det(DS)(y) ) > dy

[Using the identity ∇(φ ◦ S) = (DS)∗.(∇φ) ◦ S, we obtain:]

= − Z

D

< E(y), (DS)∗.(∇φ)◦S det(DS)(y) ) + φ◦S(y) ∇( det(DS)(y) ) > dy

[Transposing of the matrix DS∗

, we can rewrite as follows:]

= − Z

D

{ < det(DS)(y) ) DS.E(y), (∇φ)◦S > + < ∇( det(DS)(y) ) E(y), φ◦S(y) > }dy

[Performing the change of variable y = T (x), we obtain:]

= − Z

D

{< det(DT )det(DS)◦T DS◦T.E◦T, ∇φ > + < det(DT )(∇det(DS))◦T E◦T, φ >}dx

[As (DS)◦T = (DT )−1 ⇒ det(DS)◦T = det( (DS)◦T ) = det( (DT )−1) =

( detDT )−1

(9)

= − Z

D

{< (DT )−1.E◦ T, ∇φ > + < det(DT )(∇det(DS)) ◦ T E ◦ T, φ >}dx

[Using the fact that (∇det(DS))◦T = (DT )−∗.∇(det(DS)◦T ) = (DT )−∗.∇( 1 detDT ) = − 1 (detDT )2 (DT ) −∗ .∇(detDT ), we obtain:] = − Z D

{< (DT )−1.E◦T, ∇φ > − < det(DT )−1 (DT )−∗.∇(detDT ).E◦T, φ >}dx

= Z D {div((DT )−1.E◦ T ) + det(DT )−1 < (DT )−∗ .∇(detDT ), E ◦ T >}φdx = Z D

{div((DT )−1.E◦ T ) + det(DT )−1 < ∇(detDT ), (DT )−1.E◦ T >}φdx

= Z

D

det(DT )−1{det(DT ) div((DT )−1.E◦T ) + < ∇(detDT ), (DT )−1.E◦T >}φdx

[For any scalar function a and any vector function ~A we have div(a ~A) = a div ~A+ < ∇a, ~A >R3. Therefore we have that det(DT ) div((DT )−1.E ◦

T) + < ∇(detDT ), (DT )−1.E◦ T > = div(det(DT ) (DT )−1.E◦ T ), which gives us the desired conclusion:]

= Z

D

det(DT )−1 div(det(DT ) (DT )−1.E◦ T ) φdx

 Similarly, we can prove the following proposition:

Property 2.2. For any φ∈ H1(D), we have the following identity:

∆φ ◦ T = det(DT )−1div(det(DT )(DT )−1(DT )−∗∇(φ ◦ T )) (2.10)

2.2.2. Transport of Vector Operators. Let T be a N × N matrix function defined on D . We consider the vector Divergence operator Div ~T being defined as the vector whose ith component is the (scalar) divergence of the vector composed of the ith line of the matrix T :

(Div ~T )i = div(Ti,.) = Σj=1,...,N

∂ ∂xj

Ti,j

From the previous section we obtain that

( (Div ~T ) ◦ T )i= (Div ~T )i◦ T = det(DT )−1 div( det(DT ) (DT )−1.(Ti,.) ◦ T )

It turns out that (Ti,.) ◦ T is the ith column vector of the matrix T∗◦ T so

that (DT )−1

.(Ti,.) ◦ T is the ith column of the matrix (DT )−1.T∗◦ T , and

thus the ithline of the matrix (DT )−1.T ◦T . Therefore we have the following

(10)

Property 2.3.

(Div ~T ) ◦ T = det(DT )−1 Div( det(DT ) (DT )−1.(T ◦ T ) )

2.2.3. Boundary change of variable. Let Γ = ∂Ω be a C1 manifold: there

exists a covering Γ ⊂ ∪i=1,...,MOi, open subsets and charts ci: Oi → B where

B is the unit ball of RN such that ci(Γ ∩ Oi) ⊂ B0= {x = (x′,0) ∈ B} and

ci(Ω ∩ Oi) ⊂ B + 0 = {x = (x′, z) ∈ B s.t.z > 0 }. Let ri ∈ C∞(D) ith

compact support in Oi such that 0 ≤ ri ≤ 1, Σri = 1 in a neibourhood of

the boundary Γ. For any f ∈ L1(Γ) we have

Z Γ f dΓ = Σ Z Γ∩Oi rif dΓ = Σ Z B0 rioc−1f oc−1||cof (D(c−i 1)).n0|| dx′

Where for any square matrix A, the cofactor’s matrix is

cof(A) = detA A−∗, cof(A−1) = 1 detA A ∗ We get D(c−i 1) = (Dci)−1◦ c−i 1 then cof(D(c−i 1)) = cof ((Dci)−1) ◦ c−i 1 = ( 1 detDci (Dci)∗ ) ◦ c−i 1

It can be easily verified that if T is a smooth enough transformation we have, with Σ = T (Γ), Z T (Γ) f dΣ = Z Γ f◦ T ω dΓ

Where, n being the unitary normal field on Γ,

ω = ||cof (DT ).n|| = |det(DT )| ||(DT )−∗.n|| Also, we have the following lemma ([15]):

Lemma 2.1. If the mapping s→ Ts(V ) is in C1([0, τ ]; Ck(D, Rn)), then

s→ ns◦ Ts=

DTs−∗n kDTs−∗nk

is in C1([0, τ ]; Ck(Γ))

where n and ns are the outward normal fields respectively to Ω and Ωs, on

Γ and Γs. Moreover, its derivative is given by:

d

ds(ns◦ Ts) =< DV · ns, ns>◦Tsns◦ Ts− DV

(11)

2.2.4. Shape derivatives. Assume that the transformation under considera-tion Ts(V ) is the flow mapping of a Lipschitz-continuous vector field V (s, x).

Then we get

∀x ∈ Γ, ω(s, x) = det(DTs(V )) ||(DTs(V ))−∗.n||

and

∀x ∈ Γ, ∂

∂sω(s, x)|s=0= H(x) < V (0, x), n(x) >

Where H is the mean curvature of Γ, H = T race(D2bΩ)|Γ = (∆bΩ)|Γ and

v=< V (0, x), n(x) > is the so-called normal speed of the moving boundary Γs.

3. Material Derivatives

3.1. Existence results for the derivative. Recall that O is the reference domain whose boundary is S and let Ocbe its complementary. The mapping I+ us is invertible from O onto Ωs as soon as det(I + Dus) > 0 over O. We

transport the harmonic problem whose solution is φs∈ H1(Ωs). Let

φs:= φs◦ (I + us) ∈ H1( ˆΩ)

From the Transport Lemma we know that

div( A(s).∇φs) = 0 in Oc (3.1)

where

A(s) = det(I + Dus) (I + Dus)−1.(I + Dus)−∗

Moreover, we have the following boundary condition

<n, A(s).∇φˆ s>= 0 on S (3.2) Concerning the elastic boundary condition on Γs, we have:

Ts.ns=

1

2|∇Γsφs|

2+ f on Γ

s (3.3)

where the forcing term f , may be due to the gravity acceleration and can take the form f (x) = ρgx3 in R3. The change of variable leads to

Ts◦(I+us).ns◦(I+us) =

1 2det(I + Dus)

< A(s).∇φs,∇φs >+ f ◦(I+us) on Γs

(3.4) The stress tensor is the matrix

Ts◦ (I + us) = ( 1 det(I + Dus) (I + Dus).Σ(σ(us)).(I + D∗us) ) (3.5) Where Σ(σ(us)) = Cλ,µ..(Dus+ D∗us+ 1 2Dus.D ∗ us) (3.6) ns◦ (I + us) = (I + Dus)−∗.∇(bΩs◦ (I + us) )

Equation (3.1), (3.2), (3.3), (3.4), and (3.5) above form a system that we can rewrite as

(12)

where the mapping

F : [0, s1[× (H1( ˆΩ, RN) × H1( ˆΩ) → H−1( ˆΩ, RN) × H−1( ˆΩ)

verifies the Implicit Function theorem assumptions so that the derivative u′ := ∂s∂us|s=0exists in H1( ˆΩ, RN) and also ˙φ:= ∂s∂φs|s=0exists in H1( ˆΩ, R).

4. Weak Formulation of the Problem

Now we write the variational form associated with system (2.4). ∀Ψ ∈ H1(D), ∀R ∈ H1(D, RN), we have Z Ωs T r(Ts.DR)dx + Z Ωc s <∇φs,∇Ψ > dx = Z Γs {1 2|∇φs| 2+ ρg x 3} < ns, R > dΓs

Our goal is to compute the s derivatives (at s = 0). Let T′

= [∂s∂ Ts]s=0,

v=< V (0), n > on Γ, and n′

= dsd(∇bΩs)s=0. Recall that ns= ∇bΩs is the

unit normal vector field on Γsout going to the (elastic) set Ωs, and H = ∆bΩ

is the mean curvature of Γ. We obtain: Z Ω T r(T′.DR)dx + Z Γ T r(T.DR) v dΓ + Z Ωc <∇φ′,∇Ψ > dx + Z Γ <∇φ, ∇Ψ > v dΓ = Z Γ { < ∇φ′,∇φ > < n, R > dΓ + Z Γ {1 2|∇φ| 2+ ρg x 3} < n′, R > dΓ + Z Γ ∂ ∂n{ ( 1 2|∇φ| 2+ ρg x 3) < ∇bΩ, R >} v dΓ + Z Γ H (1 2|∇φ| 2+ ρg x 3) < ∇bΩ, R > v dΓ

We recall the by part integration formula: Z Ω T r(T′.DR)dx = − Z Ω < ~Div(T′), R > dx + Z Γ <T′.n, R > dΓ

Then, taking Ψ (repectively R) with compact support in Ωc (respectively

in Ω) we get the following equations:

−∆φ′= 0 in Ωc − ~Div(T′) = 0 in Ω But ∂ ∂n{ ( 1 2|∇φ| 2+ ρg x 3) < ∇bΩ, R >} =< n, ∇{ (1 2|∇φ| 2+ ρg x 3) < ∇bΩ, R >} >

(13)

= (< n, D2φ.∇φ > + ρg n3) < n, R > + ( 1 2|∇φ| 2+ρg x 3) < n , D2bΩ.R+D∗R.n > Notice that < n, D2bΩ.R >=< D2bΩ.n, R >= 0, as D2bΩ.n= 0

Therefore, concerning the boundary conditions, and taking Ψ ∈ H1(D) and R = 0, we obtain:

∂ ∂nφ

= divΓ( v ∇Γφ)

In addition, taking Ψ = 0 and R ∈ H1(D, RN) with DR.n = 0 on Γ, we have:

< n , D∗R.n >=< DR.n, n >= 0. Hence we have the following identity:

∂ ∂n{ ( 1 2|∇φ| 2+ ρg x 3) < ∇bΩ, R >} = (< n, D2φ.∇φ > + ρg n3) < n, R >

Finally, we obtain the following variational problem at the boundary: ∀R ∈ H1(Γ, RN) Z Γ <T′.n, R > dΓ + Z Γ T r(T.DR) v dΓ = Z Γ <∇φ′,∇φ > < n, R > dΓ + Z Γ {1 2|∇φ| 2+ ρg x 3} < n′, R > dΓ + Z Γ (< n, D2φ.∇φ > + ρg n3) < n, R > v dΓ + Z Γ H (1 2|∇φ| 2+ ρg x 3) < ∇bΩ, R > v dΓ

We have to perform a tangential by part integration in the term Z Γ T r(T.DR) v dΓ = Z Γ Ti,j ∂ ∂xjRiv dΓ = Z Γ <Ti,.,∇Ri> v dΓ

But as for all i we have ∂n∂ Ri = 0,

∇Ri= ∇ΓRi + ∂ ∂nRi ~n= ∇ΓRi, and Z Γ T r(T.DR) v dΓ = Z Γ <Ti,.,∇ΓRi> v dΓ = − Z Γ div(v Ti,.) Ri dΓ = − Z Γ < ~Div(v T ), R > dΓ,

(14)

∀R ∈ H1(Γ, RN) Z Γ <T′.n, R > dΓ = + Z Γ < ~Div(v T ), R > dΓ + Z Γ <∇φ′,∇φ > < n, R > dΓ + Z Γ {1 2|∇φ| 2+ ρg x 3} < n′, R > dΓ + Z Γ (< n, D2φ.∇φ > + ρg n3) < n, R > v dΓ + Z Γ H (1 2|∇φ| 2+ ρg x 3) < ∇bΩ, R > v dΓ Therefore, we obtain T′ .n = ~Div(v T ) + < ∇φ′ ,∇φ > ~n + {1 2|∇φ| 2+ ρg x 3} ~n′ + (< n, D2φ.∇φ > + ρg n3) v ~n + H ( 1 2|∇φ| 2+ ρg x 3) v ~n

4.1. Calculus of the tangent vector n′. From [?], [?], [?], we know that in some neighborhood U of Σ = ∪0<s<s1{s} × ∂Ωs the oriented distance

function solves the convection equation ∂

∂sbΩs+ ∇bΩs.V(s)opΓs = 0

where pΓs = Id − bΩs∇bΩs is the projection mapping onto Γs.

Then we get n′ := ∂ ∂s(∇bΩs)s=0= ∇( ∂ ∂sbΩs)s=0= ( ∇(−∇bΩs.V(s)opΓs) )s=0 = −( D2bs.V(s)opΓs )s=0− ( D ∗ (V (s)opΓs).∇bΩs )s=0 s= 0, x ∈ Γ, n′(x) = − D2b(x).V (0, x) − DΓ∗V(0, x).n(x) = −∇Γv(x)

where again v(x) =< V (0, x), n(x) > on Γ is the normal speed of the bound-ary.

As we have

~

Div(T ) = 0 in Ω,

assuming the boundary Γ smooth we get that this equation holds true up to the boundary, so that the term ~Div(vT ) simplifies and we get

~

Div(vT ) = v ~Div(T ) + T .∇v. Thus we have the following new expression:

T′.n = T .∇v + < ∇Γφ′,∇Γφ > ~n− { 1 2|∇Γφ| 2+ ρg x 3} ∇Γv + (< n, D2φ.∇Γφ > + 1 2H |∇Γφ| 2+ (1 + H )ρg n 3) v ~n (4.1)

(15)

Regarding the term ∇v on Γ, we have:

∇v = ∇(< V (0), ∇bΩ >) = D2bΩ.V(0) + D∗V(0).∇bΩ

Therefore it follows that

<∇v, n >=< DV (0).n, n > and ∇v = ∇Γv+ < DV (0).n, n > ~n Now using T .n = p~n = (1 2|∇Γφ| 2+ ρg x

3)~n, equation (4.1) simplifies to:

T′.n = [ T − {1 2|∇Γφ| 2+ ρg x 3} ].∇Γv + < ∇Γφ′,∇Γφ > ~n + (< n, D2φ.∇Γφ > + 1 2H |∇Γφ| 2+ (1 + H )ρg n 3) v ~n + < DV (0).n, n > (1 2|∇Γφ| 2+ ρg x 3) ~n (4.2)

4.2. The stress tensor Ts◦Ts. Recall that Ts = ϕs◦ϕ−1, where ϕs= I +us,

and that V (s) = ∂s∂ Ts◦ T−1 s = ∂s∂ϕs◦ ϕ −1 s . Then we get V(s) = ∂ ∂suso(I + us) −1. If we let u′ := ( ∂ ∂sus )s=0 then we obtain V(0) = u′◦ (I + u)−1 On the boundary Γ we have

v=< u′◦ (I + u)−1, n > so that

∇Γv= ∇Γ(< (u′◦(I+u)−1), n >) = DΓ∗(u ′

◦(I+u)−1).n + D2b.(u′◦(I+u)−1)Γ

Remark 4.1. We design by K ∈ L(H1/2(Γ), H−1/2(Γ)) the self adjoint

Dirichlet to Neuman operator at the boundary associated with harmoniques functions in Ω. Then we get

φ′|Γ = K.(divΓ(v∇Γφ))

The stress tensor is the matrix TsoTs= (

1 det(I + Dus)

(I + Dus).Σ(σ(us)).(I + D∗us) )o(I + u)−1 (4.3)

where Σ(σ(us)) = Cλ,µ..(Dus+ D∗us+ 1 2Dus.D ∗ us) (4.4)

In (4.4) we assumed the four entries elasticity tensor to be governed by the Lam´ecoefficients.

(16)

Taking derivative w.r.t s in (4.3), we obtain:

( [ ∂

∂sTsoTs ]s=0 )o(I + u) = −

div(u′

)

det(I + Du) (I + Du).Cλ,µ..(σ(u)).(I + D

u)

+( 1

det(I + Du)D(u

).Cλ,µ..(σ(u)).(I + D∗u) )

+( 1

det(I + Du)(I + Du).Cλ,µ..(σ

).(I + D∗u) )

+( 1

det(I + Du)(I + Du).Cλ,µ..(σ(u)).D

∗ (u′) ) where σ′ = D(u′) + D∗(u′) +1 2( D(u ′ ).D∗u+ Du.D∗(u′) ) Now we let ˙ T = [ ∂ ∂sTsoTs ]s=0 = T′ + DT .(u′◦ (I + u)−1)

where D is a “new coming“: it is a three entries tensor, representing the gradient of the matrix T . Its contraction with the vector (u′

◦ (I + u)−1

) gives the matrix DT .(u′

◦ (I + u)−1). Then we have

T′ = { − div(u

)

det(I + Du) (I + Du).Cλ,µ..(σ(u)).(I + D

u) }o(I + u)−1

+{ ( 1

det(I + Du)D(u

).Cλ,µ..(σ(u)).(I + D∗u) ) }o(I + u)−1

+{ ( 1

det(I + Du)(I + Du).Cλ,µ..(σ

).(I + D∗u) ) }o(I + u)−1

+{ ( 1

det(I + Du)(I + Du).Cλ,µ..(σ(u)).D

(u′) ) }o(I + u)−1

− DT .(u′◦ (I + u)−1)

5. Linearization around “rest”

The most “popular” framework (among mathematicians) consists in con-sidering u = 0. With this assumption, we have the following simplification:

T′ = Cλ,µ..(σ′) − DT .(u′◦ (I + u)−1)

Since u = 0, we have that T = 0 and thus DT = 0. Moreover, σ′ = D(u′◦ (I + u)−1) + D∗(u′◦ (I + u)−1) Therefore T′ = Cλ,µ..(D(u′◦ (I + u)−1) + D∗(u′◦ (I + u)−1)) T′ = λ det(D(u′ ◦ (I + u)−1)) I + µ (D(u′ ◦ (I + u)−1) + D∗ (u′ ◦ (I + u)−1))

(17)

So now we are in the classical linear elasticity framework: U := (u′

◦ (I + u)−1

) is the linearized displacement, while ¯T = T′

= λ det(DU ) I + µ (DU + D∗U

) = is the associated stress function.

5.1. The linearization of the coupled fluid-structure problem around stress less steady structure. We assume small variations around the rest state u = 0, T = 0 for the elastic body occupying the volume Ω in D ⊂ R3.

The fluid speed v is steady, but not zero. We assume this flow to be irrota-tional so that v derives from an harmonic potential in Ωc = D \ ¯Ω, that is vs= ∇φs. From (4.2), with the following notation

Φ= φ′,

we obtain the following system for the fluid-structure coupling (with ¯T = λ det(DU ) I + µ (DU + D∗U )): ∆Φ = 0 in Ωc, ~ Div( ¯T ) = 0 in Ω, ∂ ∂nΦ= − + divΓ( U.n ∇Γφ) on Γ, ¯ T .n = − {1 2|∇Γφ| 2+ ρg x 3} .∇Γ(U.n) + < ∇ΓΦ,∇Γφ > ~n + (< n, D2φ.∇Γφ > + 1 2H |∇Γφ| 2+ (1 + H )ρg n 3) U.n ~n + < DU.n, n > (1 2|∇Γφ| 2+ ρg x 3) ~n (5.1)

Note that the coupling on the boundary interface Γ takes into account the mean curvature of the boundary H. This is an important point.

References

[1] R.A. Adams, “Sobolev Spaces”, Academic Press, New York-London, 1975.

[2] G. Avalos and R. Triggiani, Semigroup Well-posedness in the Energy Space of a Parabolic-Hyperbolic Coupled Stokes-Lame PDE System of Fluid-Structure Interac-tion, Discrete and Continuous Dynamical Series S, Volume 2, Number 3, September 2009.

[3] V. Barbu, Z. Grujic, I. Lasiecka, and A. Tuffaha, Smoothness of Weak Solutions to a Nonlinear Fluid-structure Interaction Model Indiana University Mathematics Journal, Vol. 57, No. 3, 2008.

[4] L. Bociu, Local and global wellposedness of weak solutions for the wave equation with nonlinear boundary and interior sources of supercritical exponents and damping, , Nonlin. Analysis TMA. (2008). [doi:10.1016/j.na.2008.11.062]

[5] L. Bociu and I. Lasiecka, Blow-up of weak solutions for the semilinear wave equations with nonlinear boundary and interior sources and damping Applicationes Mathemat-icae, 35 (2008), 281–304.

[6] L. Bociu and I. Lasiecka, Uniqueness of Weak Solutions for the Semilinear Wave Equations with Supercritical Boundary/Interior Sources and Damping, Discrete Con-tin. Dyn. Syst., 22 (2008), 835–860.

[7] L. Bociu and I. Lasiecka, Hadamard wellposedness for nonlinear wave equations with supercritical sources and damping, Submitted to JDE, 2009.

(18)

[8] L. Bociu, M. Rammaha and D. Toundykov, On a wave equation with supercriti-cal interior and boundary sources and damping terms, Submitted to Mathematische Nachrichten, 2009.

[9] S. Boisg´erault and J.-P. Zol´esio. Shape derivative of sharp functionals governed by Navier-Stokes flow. In Partial differential equations (Praha, 1998), volume 406 of Chapman & Hall/CRC Res. Notes Math., pages 49–63. Chapman & Hall/CRC, Boca Raton, FL, 2000.

[10] S. Boisg´erault and J.P. Zol´esio. Boundary variations in the Navier-Stokes equations and Lagrangian functionals. In Shape optimization and optimal design (Cambridge, 1999), volume 216 of Lecture Notes in Pure and Appl. Math., pages 7–26. Dekker, New York, 2001.

[11] P.G. Ciarlet, “Mathematical Elasticity Volume I: Three-dimensional Elasticity, Noth Holland, Amsterdam-New York-Oxford-Tokyo, 1988.

[12] D. Coutand and S. Shkoller, Motion of an Elastic Solid inside and Incompressible Viscous Fluid, Arch. Rational Mech. Anal. 176 (2005), p. 25-102.

[13] G. Da Prato and J.-P. Zol´esio. Existence and optimal control for wave equation in moving domain. In Stabilization of flexible structures (Montpellier, 1989), volume 147 of Lecture Notes in Control and Inform. Sci., pages 167–190. Springer, Berlin, 1990. [14] Giuseppe Da Prato and Jean-Paul Zol´esio. Boundary control for inverse free

bound-ary problems. In Boundbound-ary control and boundbound-ary variation (Sophia-Antipolis, 1990), volume 178 of Lecture Notes in Control and Inform. Sci., pages 163–173. Springer, Berlin, 1992.

[15] M.C. Delfour and J.P. Zol´esio, “Shapes and Geometries: Analysis, Differential Cal-culus and Optimization, Advances in Design and Control, SIAM 2001.

[16] Michel C. Delfour and Jean-Paul Zol´esio. Shape analysis via oriented distance func-tions. J. Funct. Anal., 123(1):129–201, 1994.

[17] M. C. Delfour and J.-P. Zol´esio. Dynamical free boundary problem for an incom-pressible potential fluid flow in a time-varying domain. J. Inverse Ill-Posed Probl., 12(1):1–25, 2004.

[18] Michel C. Delfour and Jean-Paul Zol´esio. Boundary evolution. Honoring Da Prato, 4(1):29–52, 2006.

[19] F. R. Desaint and Jean-Paul Zol´esio. Manifold derivative in the Laplace-Beltrami equation. J. Funct. Anal., 151(1):234–269, 1997.

[20] Fabrice Desaint and Jean-Paul Zol´esio. Shape boundary derivative for an elastic mem-brane. In Advances in mathematical sciences: CRM’s 25 years (Montreal, PQ, 1994), volume 11 of CRM Proc. Lecture Notes, pages 481–491. Amer. Math. Soc., Providence, RI, 1997.

[21] F. R. Desaint and J. P. Zol´esio. Shape derivative for the Laplace-Beltrami equation. In Partial differential equation methods in control and shape analysis (Pisa), volume 188 of Lecture Notes in Pure and Appl. Math., pages 111–132. Dekker, New York, 1997.

[22] Fabrice Desaint and Jean-Paul Zol´esio. Sensitivity analysis of all eigenvalues of a symmetrical elliptic operator. In Boundary control and variation (Sophia Antipolis, 1992), volume 163 of Lecture Notes in Pure and Appl. Math., pages 141–160. Dekker, New York, 1994.

[23] C. Grandmont, Existence et unicite de solutions d’un probleme de couplage fluide-structure bidimensionnel stationnaire, C.R.Acad. Sci. Paris, t. 326, Serie I (1998), p. 651-656.

[24] C. Grandmont, Existence for a Three-Dimensional Steady State Fluid-Structure In-teraction Problem, J. math. fluid. mech. 4 (2002), p. 76-94.

[25] J. Soko lowski and J.-P. Zol´esio. Shape design sensitivity analysis of plates and plane elastic solids under unilateral constraints. J. Optim. Theory Appl., 54(2):361–382, 1987.

(19)

[26] J. Soko lowski and J.-P. Zol´esio. Shape sensitivity analysis of contact problem with prescribed friction. Nonlinear Anal., 12(12):1399–1411, 1988.

[27] Jan Soko lowski and Jean-Paul Zol´esio. Shape sensitivity analysis of unilateral prob-lems. SIAM J. Math. Anal., 18(5):1416–1437, 1987.

[28] M. Souli and J.-P. Zol´esio. Shape derivative of discretized problems. Comput. Methods Appl. Mech. Engrg., 108(3-4):187–199, 1993.

[29] M. Souli and J. P. Zol´esio. Finite element method for free surface flow problems. Comput. Methods Appl. Mech. Engrg., 129(1-2):43–51, 1996.

[30] M. Souli, J. P. Zol´esio, and A. Ouahsine. Shape optimization for non-smooth geometry in two dimensions. Comput. Methods Appl. Mech. Engrg., 188(1-3):109–119, 2000. [31] Jean-Paul Zol´esio. Weak shape formulation of free boundary problems. Ann. Scuola

Norm. Sup. Pisa Cl. Sci. (4), 21(1):11–44, 1994. E-mail address: lbociu2@math.unl.edu

CNRS-INLN and INRIA, Sophia Antipolis, France, 06560 E-mail address: Jean-Paul.Zolesio@sophia.inria.fr

Références

Documents relatifs

Chapter 7: This chapter presents the design of a backstepping control feedback in order to exponentially sta- bilize the same wave equation considered in Part I, but assuming that

We give the material for deducing from the Lipschitz regularity result for viscosity solutions and a right hand side continuous, that the same holds true for weak solutions and f ∈ L

In summation, for the studied interconnected nonlinear systems, an interconnected observer design method is proposed by combining both actuator and process subsystem state

Konotop, Nonlinear modes in a generalized PT -symmetric discrete nonlinear Schr¨ odinger equations, J.. For each line the first plot stands for the initial condition, the second

The authors thank E. - Free vibrations and dynamic buckling of the extensible beam, J. 2014 Nonlinear vibrations of beams and rectangular plates, Z. A.) .2014 Nonlinear

By opposition to the power-like case [5], where the initial trace was constructed by duality arguments based upon H¨older inequality and delicate choice of test func- tions, our

1. Fibich [7] noted that in the nonlinear optics context, the origin of the nonlinear damping is multiphoton absorption. Damped Nonlinear Schr¨ odinger Equation, Blow-up,

Fine topology and fine trace on the boundary associated with a class of quasilinear differential equations, Comm... F., Spatial branching processes, random snakes and