<ph f="cmex"> </ph><ph f="cmbx">On the interaction between quasilinear elastodynamics and the Navier-Stokes equations</ph>
E-mail address : coutand@math.ucdavis.edu E-mail address : shkoller@math.ucdavis.edu
Daniel Coutand and Steve Shkoller Department of Mathematics University of California at Davis Davis, CA 95616

1 Introduction

We establish the existence and uniqueness in Sobolev spaces of strong solutions to the unsteady fluid-structure interaction problem consisting of a nonlinear large-displacement elastic solid coupled to a viscous incompressible Newtonian fluid. The fluid motion is governed by the incompressible Navier-Stokes equations, while the solid, which can be either compressible or incompressible, is modeled by the celebrated St. Venant-Kirchhoff constitutive law (although our method can be applied to more general quasilinear hyperelastic models).
The first fluid-solid interaction problems solved were for the case of a rigid body inside of a viscous flow in a bounded domain (see [7, [12, [4, [13), and the case of a rigid body inside of a viscous flow in an infinite domain ([22, [20, [15). Later, the elastic body was modeled with the restriction of either a finite number of modes ([8) or with a hyperviscous type law for the solid ([2, [10), essentially by the same type of Eulerian global variational methods developed in [7. For the steady-state problem, which is elliptic in both phases, [11solved the case of solid modeled as a St. Venant-Kirchhoff material. In [18, an Eulerian approach was used for the case in which the solid is a visco-hyperelastic material, which is a regularization of a hyperbolic model of solid deformation.
With the exception of our recent well-posedness result for the case of a linear elastic solid in [5, there are no known existence results for fluid-structure interaction when the solid is modeled by a standard second-order hyperbolic equation. This may be attributed to the difficulties associated with coupling a parabolic PDE for the fluid with a hyperbolic PDE for the solid through continuity of the velocity and traction vectors across the moving material interface. As we explained in [5, an iteration scheme between fluid and solid phases fails to converge due to a regularity loss induced by the hyperbolic phase (this divergent behavior has been computationally noted as well in [14), and so we developed a method comprised of the following new ideas: first, a functional framework which scales in a hyperbolic fashion for both the fluid and solid phases. This scaling leads to additional compatibility conditions in the fluid phase (when compared to the use of the classical parabolic framework), and is absolutely crucial for obtaining consistent energy estimates.
Second, we developed a regularity theory founded upon central trace estimates for the velocity vector restricted to the interface, rather than traditional interior regularity arguments which do not work for our problem. Third, we were forced to bypass the use of the frozen (or constant) coefficient basic linear problem, which requires estimates on one more time derivative of the pressure function than the initial data allows, and created a new method wherein the solution was found as a limit of a sequence of penalized problems set in the Lagrangian framework.
The penalization scheme approximates the divergence-free constraint, whereas the Lagrangian framework alleviates the difficulties associated with the lack of a priori estimates in the solid phase for the frozen coefficient problem; this method indeed differs significantly from the classical methods used in fluid-fluid interface problems (see for instance [21, [1).
The fundamental difficulty in extending our result to the case of nonlinear elasticity is the absence of any method of analysis for quasilinear elastodynamics which is compatible with the general scheme of [5, involving a global Lagrangian variational formulation and the use of difference quotients to track the regularity of interface data. We remind the reader that unlike the analysis of elastostatic motion, direct inverse function theorem arguments cannot be applied directly to the case of quasilinear elastodynamics due to the fact that the perturbation term arising from the nonlinear operator is not an element of the appropriate function space for optimal regularity. Alternatively, one might attempt a fixed-point approach, wherein a portion of the nonlinear elasticity operator is viewed as a forcing function coming from a given velocity v   , and then try to solve a linear problem for an unknown w   . The difficulty in this approach stems from the fact that one has to find exact time derivatives of elastic energies for the forcing term associated to the elasticity operator, which is complicated by the inner-product of a term involving 0 t v   and a term involving w   . This difficulty is overcome in [6, by a clever and essential use of the Dirichlet boundary condition in order to reformulate the problem in a non-standard way. As it turns out, the variety of known methods that have been used in the well-studied area of quasilinear elasticity, such as those in [6and [17for the Dirichlet boundary condition, or in [19and [9for the Neumann boundary condition, require a priori knowledge of the boundary data regularity, and are hence instrinsically incompatible with fluid-structure interaction analysis (in fact, the methods devised for Dirichlet conditions do not work for Neumann conditions and vice versa). Indeed, of these various methodologies, only [6and [17use a variational approach, the others employing either semi-group techniques as in the early work of [16in the full space, or the technical paradifferential calculus in [19for the two dimensional Neumann case.
In this paper, we develop a new method for quasilinear elastodynamics, variational in nature, that is compatible with our method in [5; we proceed in two steps.
First, we add a specific artificial viscosity to the solid phase which regularizes the system, thus converting our hyperbolic PDE into a parabolic one, transforming the fluid-structure interaction into a fluid-fluid interface-type problem for which existence and uniqueness of solutions is already known on a time interval that a priori shrinks to zero as the artificial viscosity κ   tends to zero. Second, and this is where the primary difficulty rests, we prove that our specific choice of parabolic smoothing renders the time interval (on which a unique solution exists) independent of κ   ; furthermore, our a priori estimates allow us to construct a solution by weak convergence in strong norms. We note that the use of higher-order operators in the artificial viscosity term, while providing the necessary a priori control of the regularity of the moving interface, would not yield κ   -independent estimates which are essential here. Also, as our parabolic reguralization method is not specialized to any particular boundary condition, it thus provides a unified approach to the classical problem of quasilinear elastodynamics when the solid is not coupled to a fluid.
We now proceed to the formulation of our problem. Let Ω R 3   denote an open, bounded, connected and smooth domain with smooth boundary Ω   which represents the fluid container in which both the solid and fluid move. Let Ω s ( t ) ¯ Ω   denote the closure of an open and bounded subset representing the solid body at each instant of time t [ 0 , T ]   with Ω f ( t ) : = Ω / Ω s ( t ) ¯   denoting the fluid domain at each t [ 0 , T ]   . Note that in our analysis Ω s ( t )   is not necessarily connected, which allows us to handle the case of several elastic bodies moving in the fluid.
Remark 1. If a function u   is defined on all of Ω   , we will denote u f = u 1 Ω ¯ 0 f   and u s = u 1 Ω ¯ 0 s   . This allows us to indicate from which phase the traces on Γ ( 0 ) : = Ω f ( 0 ) ¯ Ω s ( 0 ) ¯   of various discontinuous terms arise, and also to specify functions that are associated with the fluid and solid phases.
For each t ( 0 , T ]   , we wish to find the location of these domains inside Ω   , the divergence-free velocity field u f ( t , )   of the fluid, the fluid pressure function p ( t , )   on Ω f ( t )   , the fluid Lagrangian volume-preserving configuration η f ( t , ) : Ω f ( 0 ) = Ω 0 f Ω f ( t )   , and the elastic Lagrangian configuration field η s ( t , ) : Ω s ( 0 ) = Ω 0 s Ω s ( t )   such that Ω = η s ( t , Ω 0 s ¯ ) η f ( t , Ω 0 f )   , where η t f ( t , x ) = u f ( t , η f ( t , x ) )   and u f   solves the Navier-Stokes equations in Ω f ( t )   :
u t f + ( u f ) u f = div T f + f f ,
div u f = 0 ,
with
T f = ν Def u f p I , (1.1)
and η s   solves the elasticity equations on Ω s ( 0 )   η ¨ s = div T s + f s ,   with T s = λ 2 Tr ( ( η s ) T η s I ) I + μ ( ( η s ) T η s I ) ,   and where the equations are coupled together by the continuity of the normal component of stress along the material interface Γ ( t ) : = Ω s ¯ ( t ) Ω f ¯ ( t )   expressed in the Lagrangian representation on Γ 0 : = Γ ( 0 )   as T s N = [ T f η f ] [ ( η f ) 1 N ] ,   and the continuity of particle displacement fields along Γ 0   η f = η s ,   together with the initial conditions u ( 0 , x ) = u 0 ( x )   , and η ( 0 , x ) = x   and the Dirichlet (no-slip) condition on the boundary Ω   of the container u f = 0   , where ν > 0   is the kinematic viscosity of the fluid, λ > 0   and μ > 0   denote the Lamé constants of the elastic material, N   is the outward unit normal to Γ 0   and Def u   is twice the rate of deformation tensor of u   , given in coordinates by u i , j + u j , i   . All Latin indices run through 1 , 2 , 3   , the Einstein summation convention is employed, and indices after commas denote partial derivatives.
We now briefly outline the proof. As the solid and fluid phases are naturally expressed in the Lagrangian and Eulerian framework, respectively, we begin by transforming the fluid phase into Lagrangian coordinates, leading us to the system of equations ( 3.2 ) and, as in [5, we work in an hyperbolic framework in order to accomodate the dual nature of the problem (parabolic in the fluid and hyperbolic in the solid).
In order to solve ( 3.2 ), in Section  7 , we first add a particular form of artificial viscosity to the quasilinear hyperbolic equation in the solid, transforming the hyperbolic phase into a parabolic one; specifically, we add the term κ L ( η t )   , where L   denotes the linearized (about the identity) elasticity operator and η t   is the material velocity.
We hence obtain an interface problem that is parabolic in nature in both phases, and can be thought as a fluid-fluid parabolic interface problem for which well-posedness is classical, save for the fact that both phases are required to scale in an hyperbolic fashion. The time interval of existence [ 0 , T κ ]   for this parabolic system a priori shrinks to zero as κ 0   .
In Section  8 , we establish κ   -independent estimates on the solutions v κ   to the regularized parabolic problem on the time interval [ 0 , T κ ]   by identifying exact time derivatives of elastic energies, and establish regularity of the interface. A direct fixed-point approach for ( 3.2 ) does not appear to yield these exact time derivatives for the elastic energy, whereas the the regularized problem ( 7.1 ) does indeed lead to them. An essential key for obtaining estimates independent of κ   inside the solid is Lemma  6.1 . Whereas the trace estimates could be carried with other choices of artificial viscosity, we absolutely need the special choice made in our analysis in order to recover the regularity inside the solid independently of κ   . In particular, a different choice of a regularizing operator either of the same order such as η t   or of higher order such as L 2 η   or L 2 η t   would not provide κ   -independent estimates.
In Section  9 , we then explain how our estimates allow the construction of solutions v κ   on a time interval independent of κ   , still with energy estimates independent of κ   . The existence of a solution to ( 3.2 ) then follows by weak convergence as κ 0   .
Uniqueness is established in Section  11 in the same functional framework used for existence.
As our method seemingly requires more regularity on the initial data in the solid than it should, due to the artificial viscosity in the compatibility conditions, we explain in Section  12 how this extra regularity can be removed, thus leading to the result with optimal regularity.
Section  13 is dedicated to the case where the incompressibility constraint is added to the solid. The additional difficulty with respect to the compressible case comes from the fact that we control the velocity uniformly in κ   in function spaces which possess less regularity than in the fluid, whereas the pressure is controlled uniformly in the same regularity spaces in both phases. Also, we cannot use Lemma  6.1 in the most optimal form for the regularity of the pressure in the solid phase.

2 Notational simplification

Although a fluid with a Neumann (free-slip) boundary condition indeed obeys the constitutive law ( 1.1 ), we will replace for notational convenience ( 1.1 ) with
T f = ν u f p I ; (2.1)
this amounts to replacing the energy Ω 0 f Def u f : Def v   by Ω 0 f u f : v   , which is not a problem mathematically due to the well-known Korn inequality. Henceforth, we shall take ( 2.1 ) as the fluid constitutive law.

3 Lagrangian formulation of the problem

In regards to the forcing functions, we shall use the convention of denoting both the fluid forcing f f   and the solid forcing f s   by the same letter f   . Since f f   has to be defined in Ω   (because of the composition with η   ), and f s   must be defined in Ω 0 s   , we will assume that the forcing f   is defined over the entire domain Ω   .
Let
a ( x ) = [ Cof η f ( x ) ] T , (3.1)
where ( η f ( x ) ) j i = ( η f ) i / x j ( x )   denotes the matrix of partial derivatives of η f   .
Clearly, the matrix a   depends on η   and we shall sometimes use the notation a j i ( η )   to denote the formula ( 3.1 ).
Let v = u η   denote the Lagrangian or material velocity field, q = p η   is the Lagrangian pressure function (in the fluid), and F = f f η f   is the fluid forcing function in the material frame. Then, as long as no collisions occur between the solids (if there are initially more than one) or between a solid and Ω   , the problem can be reformulated as
η t = v in ( 0 , T ) × Ω , (3.2a)
v t i ν ( a l j a l k v i , k ) , j + ( a i k q ) , k = F i in ( 0 , T ) × Ω 0 f , (3.2b)
a i k v i , k = 0 in ( 0 , T ) × Ω 0 f , (3.2c)
v t c m j k l [ ( η , m η , j δ m j ) η , k ] , l = f in ( 0 , T ) × Ω 0 s , (3.2d)
c m j k l ( η , m η , j δ m j ) η i , k N l = ν v i , k a l k a l j N j q a i j N j on ( 0 , T ) × Γ 0 , (3.2e)
v ( t , ) H 0 1 ( Ω ; R 3 ) a.e. in ( 0 , T ) , (3.2f )
v = u 0 on Ω × { t = 0 } , (3.2g)
η = Id on Ω × { t = 0 } , (3.2h)
where N   denotes the outward-pointing unit normal to Γ 0   (pointing into the solid phase), and c i j k l = λ δ i j δ k l + μ ( δ i k δ j l + δ i l δ j k ) .   Throughout the paper, all Greek indices run through 1 , 2   and all Latin indices run through 1 , 2 , 3   . Note that the continuity of the velocity along the interface is satisfied in the sense of traces on Γ 0   by condition ( 3.2f ), whereas the continuity of the normal stress along the interface is represented by ( 3.2e ).

4 Notation and conventions

We begin by specifying our notation for certain vector and matrix operations.
  • We write the Euclidean inner-product between two vectors x   and y   as x y   , so that x y = x i y i   .
  • The transpose of a matrix A   will be denoted by A T   , i.e., ( A T ) j i = A i j   .
  • We write the product of a matrix A   and a vector b   as A b   , i.e, ( A b ) i = A j i b j   .
  • The product of two matrices A   and S   will be denoted by A S   , i.e., ( A S ) j i = A k i S j k   .
For T > 0   and k N   , we set
V f k ( T ) = { w L 2 ( 0 , T ; H k ( Ω 0 f ; R 3 ) ) | t n w L 2 ( 0 , T ; H k n ( Ω 0 f ; R 3 ) ) , n = 1 , . . . , k } ,
with V s k ( T )   defined with Ω s 0   replacing Ω f 0   .
In order to specify the initial data for the weak formulation, we introduce the space L d i v , f 2 = { ψ L 2 ( Ω ; R 3 ) | div ψ = 0 in Ω 0 f , ψ N = 0 on Ω } ,   which is endowed with the L 2 ( Ω ; R 3 )   scalar product.
The space of velocities, X T   , where the solution to ( 3.2 ) lives, is defined as the following separable Hilbert space:
X T = { v L 2 ( 0 , T ; H 0 1 ( Ω ; R 3 ) ) | ( v f , 0 v s ) V f 4 ( T ) × V s 4 ( T ) } , (4.1)
endowed with its natural Hilbert norm
v X T 2 = v L 2 ( 0 , T ; H 0 1 ( Ω ; R 3 ) ) 2
+ n = 0 3 [ t n v L 2 ( 0 , T ; H 4 n ( Ω 0 f ; R 3 ) ) 2 + t n 0 v L 2 ( 0 , T ; H 4 n ( Ω 0 s ; R 3 ) ) 2 ] .
We also need the space
Y T = { ( v , q ) X T × L 2 ( 0 , T ; H 3 ( Ω 0 f ; R ) ) | t n q L 2 ( 0 , T ; H 3 n ( Ω 0 f ; R ) ) ( n = 1 , 2 ) } ,
endowed with its natural Hilbert norm
( v , q ) Y T 2 = v X T 2 + n = 0 2 t n q L 2 ( 0 , T ; H 3 n ( Ω 0 f ; R ) ) 2 .
We shall also need L   -in-time control of certain norms of the velocity, which necessitates the use of the following closed subspace of X T   :
W T = { v X T | v t t t L ( 0 , T ; L 2 ( Ω ; R 3 ) ) ,
t n 0 v L ( 0 , T ; H 4 n ( Ω 0 s ; R 3 ) ) ( n = 0 , 1 , 2 , 3 ) } ,
endowed with the following norm
v W T 2 = v X T 2 + v t t t L ( 0 , T ; L 2 ( Ω ; R 3 ) ) 2 + n = 0 3 t n 0 v L ( 0 , T ; H 4 n ( Ω 0 s ; R 3 ) ) 2 .
Finally, we will also make use of the space
Z T = { ( v , q ) W T × L 2 ( 0 , T ; H 3 ( Ω 0 f ; R ) ) | t n q L 2 ( 0 , T ; H 3 n ( Ω 0 f ; R ) ) , ( n = 1 , 2 )
| q t t L ( 0 , T ; L 2 ( Ω f 0 ; R ) ) } ,
endowed with its natural norm
( v , q ) Z T 2 = v W T 2 + n = 0 2 t n q L 2 ( 0 , T ; H 3 n ( Ω 0 f ; R ) ) 2 + t 2 q L ( 0 , T ; L 2 ( Ω 0 f ; R ) ) 2 .
Remark 2. Note that our functional framework does not make use of the third time derivative of the pressure q t t t   , even though we do use the third time derivative of velocity w t t t   ; this functional framework is necessitated by the fact that the Dirichlet boundary condition together with the limited regularity of w t t t   does not allow us to obtain q t t t   with the appropriate regularity. Note also that we have added the L   -in-time control of q t t   in the definition of Z T   mostly for a more convenient way to prove our theorems, rather than absolute necessity.
Throughout the paper, we shall use C   to denote a generic constant, which may possibly depend on the coefficients ν   , λ   , μ   , or on the initial geometry given by Ω   and Ω 0 f   (such as a Sobolev constant or an elliptic constant). For the sake of notational convenience, we will also write u ( t )   for u ( t , )   .

5 A first theorem

We now state our first theorem. We impose greater regularity requirements on the initial data than is optimal so as to avoid technical difficulties associated with a particular type of initial data regularization that would otherwise be necessitated.
We consider the case of optimal regularity on the initial data in Theorem  12.1 .
Theorem 5.1. Let Ω R 3   be a bounded domain of class H 4   , and let Ω 0 s   be an open set (with a finite number 1   of connected components) of class H 4   such that Ω 0 s ¯ Ω   and such that the distance between two distinct connected components of Ω s 0   (if there are multiple solid components) is greater than zero. Let us denote Ω 0 f = Ω ( Ω 0 s ¯ ) c   . Let ν > 0   , λ > 0   , μ > 0   be given. Let
( f , f t , f t t , f t t t ) L 2 ( 0 , T ¯ ; H 3 ( Ω ; R 3 ) × H 2 ( Ω ; R 3 ) × H 1 ( Ω ; R 3 ) × L 2 ( Ω ; R 3 ) ) , (5.1a)
f ( 0 ) H 4 ( Ω ; R 3 ) , f t ( 0 ) H 4 ( Ω ; R 3 ) . (5.1b)
Assume that the initial data satisfies u 0 H 6 ( Ω 0 f ; R 3 ) H 6 ( Ω 0 s ; R 3 ) H 0 1 ( Ω ; R 3 ) L d i v , f 2 ,   as well as the compatibility conditions
[ u 0 f N ] tan = 0 on Γ 0 , w 1 = 0 = w 2 on Ω , ν u 0 f q 0 = 0 on Γ 0 , (5.2a)
[ ( ν [ ( a l k a l j ) w f , k i ] t ( 0 ) N j ) i = 1 3 ] tan [ ( q 0 a i j t ( 0 ) N j ) i = 1 3 ] tan
= [ c m j k l [ ( η s , m η s , j δ m j ) η s , k ] t ( 0 ) N l ] tan on Γ 0 , (5.2b)
[ ( ν [ ( a l k a l j ) w f , k i ] t t ( 0 ) N j ) i = 1 3 ] tan [ ( 2 q 1 a i j t ( 0 ) N j + q 0 a i j t t ( 0 ) N j ) i = 1 3 ] tan
= [ c m j k l [ ( η s , m η s , j δ m j ) η s , k ] t t ( 0 ) N l ] tan on Γ 0 , (5.2c)
ν w 1 f + ν ( ( a l j a l k ) t ( 0 ) u 0 f , k ) , j + F t ( 0 ) [ ( ( a i j ) t ( 0 ) q 0 ) , j + q 1 , i ] i = 1 3
= f t ( 0 ) + c m j k l [ [ ( η s , m η s , j δ m j ) η s , k ] , l ] t ( 0 ) on Γ 0 , (5.2d)
where the time derivatives appearing in these equations and in the following ones are computed from any w   satisfying w ( 0 ) = u 0   , t n w ( 0 ) = w n   ( n = 1 , 2   ), and from any q   satisfying t n q ( 0 ) = q n   ( n = 0 , 1 , 2   ), the quantities w n   and q n   being defined as follows. First, q 0 H 3 ( Ω 0 f ; R )   is defined by
q 0 = div f ( 0 ) + ( a i j ) t ( 0 ) u 0 i , j in Ω 0 f , (5.3a)
q 0 = ν [ u 0 f N ] N on Γ 0 , (5.3b)
q 0 N = f ( 0 ) N + ν u 0 f N on Ω , (5.3c)
and w 1 H 0 1 ( Ω ; R 3 ) H 4 ( Ω 0 s ; R 3 ) H 4 ( Ω 0 f ; R 3 )   by
w 1 = ν u 0 q 0 + f ( 0 ) in Ω 0 f (5.4a)
w 1 = f ( 0 ) in Ω 0 s . (5.4b)
Note that w 1 H 4 ( Ω 0 f ; R 3 )   since w 1 H 2 ( Ω f 0 ; R 3 )   and w 1 = 0   on Ω   , w 1 = f ( 0 )   on Γ 0   . We also have q 1 H 3 ( Ω 0 f ; R )   defined by
q 1 = div [ ν w 1 + F t ( 0 ) + [ ν ( ( a l j a l k ) t ( 0 ) u 0 i , k ) , j ( ( a i j ) t ( 0 ) q 0 ) , j ] i = 1 3 ]
+ 2 ( a i j ) t ( 0 ) w 1 i , j + ( a i j ) t t ( 0 ) u 0 i , j in Ω 0 f , (5.5a)
q 1 = ν [ w 1 f N N + ( a l k a l j ) t ( 0 ) u 0 f i , k N j N i ] q 0 a i j t ( 0 ) N j N i
c m j k l [ ( η s , m η s , j δ m j ) η s , k ] t ( 0 ) N l N on Γ 0 , (5.5b)
q 1 N = F t ( 0 ) N [ ( a i j ) t ( 0 ) q 0 ] , j N i + ν w 1 N + ν ( ( a l j a l k ) t ( 0 ) u 0 i , k ) , j N i on Ω , (5.5c)
and w 2 H 0 1 ( Ω ; R 3 ) H 4 ( Ω 0 s ; R 3 ) H 2 ( Ω 0 f ; R 3 )   by
w 2 i = ν w 1 i + ν ( ( a l j a l k ) t ( 0 ) u 0 i , k ) , j + F t i ( 0 ) ( ( a i j ) t ( 0 ) q 0 ) , j q 1 , i in Ω 0 f , (5.6a)
w 2 = f t ( 0 ) + c m j k l [ [ ( η , m η , j δ m j ) η , k ] , l ] t ( 0 ) in Ω 0 s , (5.6b)
Finally, q 2 H 1 ( Ω 0 f ; R )   is defined by
q 2 = div [ ( f η ) t t ( 0 ) + ν [ ( a l j a l k w , k , j ] t t ( 0 ) [ ( ( a i j ) t t ( 0 ) q 0 + 2 ( a i j ) t ( 0 ) q 1 ) , j ] i = 1 3 ]
+ 3 ( a i j ) t ( 0 ) w 2 i , j + 3 ( a i j ) t t ( 0 ) w 1 i , j + ( a i j ) t t t ( 0 ) u 0 i , j in Ω 0 f , (5.7a)
q 2 = ν [ ( a l k a l j ) w f i , k ] t t ( 0 ) N j N i c m j k l [ ( η s , m η s , j δ m j ) η s , k ] t t ( 0 ) N l N
q 0 a i j t t ( 0 ) N j N i 2 q 1 a i j t ( 0 ) N j N i on Γ 0 , (5.7b)
q 2 N = ( f η ) t t ( 0 ) N 2 [ ( a i j ) t ( 0 ) q 1 ] , j N i [ ( a i j ) t t ( 0 ) q 0 ] , j N i + ν w 2 N
+ 2 ν ( ( a l j a l k ) t ( 0 ) w 1 i , k ) , j N i + ν ( ( a l j a l k ) t t ( 0 ) u 0 i , k ) , j N i on Ω . (5.7c)
Then there exists T ( 0 , T ¯ )   depending on u 0   , f   , and Ω 0 f   , such that there exists a unique solution ( v , q ) Z T   of the problem ( 3.2 ). Furthermore, η C 0 ( [ 0 , T ] ; H 4 ( Ω 0 f ; R 3 ) H 4 ( Ω 0 s ; R 3 ) H 1 ( Ω ; R 3 ) )   .
Remark 3. The remarks appearing in [5at the end of Section 5 concerning the compatibility conditions and forcing functions for the linear elasticity case still hold in this setting with the necessary adjustments. In particular, we do not need the forcing functions to have the same regularity in both phases.

6 Preliminary result

In the remainder of the paper, we set L ( u ) i = [ c i j k l ( u k , l + u l , k ) ] , j .   In our limit process as the artificial viscosity tends to zero, we will make use in a crucial way of the basic following result:
Lemma 6.1. Let g C 0 ( [ 0 , T ] ; L 2 ( Ω s 0 ; R 3 ) )   and u   be such that u t L 2 ( 0 , T ; H 2 ( Ω s 0 ; R 3 ) )   and
ε L ( u t ) + L ( u ) = g on [ 0 , T ] × Ω s 0 . (6.1)
Then, independently of ε > 0   , L ( u ) L ( 0 , T ; L 2 ( Ω s 0 ; R 3 ) ) g L ( 0 , T ; L 2 ( Ω s 0 ; R 3 ) ) + L ( u 0 ) L 2 ( Ω s 0 ; R 3 ) .  
  • Proof. Since L ( u ) C 0 ( 0 , T ; L 2 ( Ω s 0 ; R 3 ) )   , let T [ 0 , T ]   be such that L ( u ( T ) ) L 2 ( Ω s 0 ; R 3 ) = sup [ 0 , T ] L ( u ) L 2 ( Ω s 0 ; R 3 ) .   If T = 0   , then the statement of the Lemma is satisfied. Now, let us assume that T ( 0 , T ]   . Let δ ( 0 , T )   be arbitrary. From ( 6.1 ), we infer that
    ε 2 T δ T L ( u t ) L 2 ( Ω s 0 ; R 3 ) 2 + T δ T L ( u ) L 2 ( Ω s 0 ; R 3 ) 2
    + ε [ L ( u ( T ) ) L 2 ( Ω s 0 ; R 3 ) 2 L ( u ( T δ ) ) L 2 ( Ω s 0 ; R 3 ) 2 ] = T δ T g L 2 ( Ω s 0 ; R 3 ) 2 .
    From the definition of T   we then infer that for any δ ( 0 , T )   , T δ T L ( u ) L 2 ( Ω s 0 ; R 3 ) 2 T δ T g L 2 ( Ω s 0 ; R 3 ) 2 ,   which after division by δ   gives at the limit δ 0   :
    L ( u ( T ) ) L 2 ( Ω s 0 ; R 3 ) 2 g ( T ) L 2 ( Ω s 0 ; R 3 ) 2 ,   which concludes the proof of the Lemma.
Remark 4. It should be clear that Lemma  6.1 applies to a more general class of linear operators than L   .

7 The smoothed problem and its basic linear problem

As we described in the introduction, we cannot find an appropriate linear problem whose fixed-point provides a solution to ( 3.2 ). We are thus lead to introduce introduce the following (parabolic) regularization of ( 3.2 ), with the artificial viscosity κ > 0   :
v t i ν ( a l j a l k v i , k ) , j + ( a i k q ) , k = f i η in ( 0 , T ) × Ω 0 f , (7.1a)
a i k v i , k = 0 in ( 0 , T ) × Ω 0 f , (7.1b)
v t i κ [ c i j k l v k , l ] , j + N ( η ) i = f i + κ h i in ( 0 , T ) × Ω 0 s , (7.1c)
κ c i j k l v k , l N j + G ( η ) i = ν v i , k a l k a l j N j q a i j N j + κ g i on ( 0 , T ) × Γ 0 , (7.1d)
v ( t , ) H 0 1 ( Ω ; R 3 ) a.e. in ( 0 , T ) , (7.1e)
v = u 0 on Ω × { t = 0 } , (7.1f )
where
N ( η ) = c m j k l [ ( η , m η , j δ m j ) η i , k ] , l in Ω s 0 , (7.2a)
G ( η ) = c m j k l [ ( η , m η , j δ m j ) η i , k ] N l on Γ 0 , (7.2b)
and
h i ( t , ) = [ c i j k l ( u 0 + t w 1 + t 2 2 w 2 ) k , l ] , j in Ω s 0 , (7.3a)
g i ( t , ) = [ c i j k l ( u 0 + t w 1 + t 2 2 w 2 ) k , l ] N j on Γ 0 . (7.3b)
Solutions to ( 3.2 ) will be obtained as the limit as κ 0   of solutions to ( 7.1 ).
Suppose that v W T   is given. Let η = I d + 0 v   and let a i j   be the quantity associated with η   through ( 3.1 ).
We are concerned with the following time-dependent linear problem, whose fixed-point w = v   provides a solution to ( 7.1 ):
w t i ν ( a l j a l k w i , k ) , j + ( a i k q ) , k = f i η in ( 0 , T ) × Ω 0 f , (7.4a)
a i k w i , k = 0 in ( 0 , T ) × Ω 0 f , (7.4b)
w t i κ [ c i j k l w k , l ] , j + N ( η ) i = f i + κ h i in ( 0 , T ) × Ω 0 s , (7.4c)
κ c i j k l w k , l N j + G ( η ) i = ν w i , k a l k a l j N j
q a i j N j + κ g i on ( 0 , T ) × Γ 0 , (7.4d)
w ( t , ) H 0 1 ( Ω ; R 3 ) a.e. in ( 0 , T ) , (7.4e)
w = u 0 on Ω × { t = 0 } . (7.4f )
Remark 5. The two forcing functions ( 7.3a ) are introduced for compatibility conditions at time t = 0   , allowing the solution of ( 7.4 ) to satisfy ( w t ( 0 ) , w t t ( 0 ) ) H 0 1 ( Ω ; R 3 ) 2   and even to satisfy the same initial conditions as solutions to ( 3.2 ) would.
In the following, for the sake of notational convenience, we will denote by N ( u 0 , ( w i ) i = 1 3 )   a generic smooth function depending only on i = 0 3 [ w 3 i H i ( Ω s 0 ; R 3 ) + w 3 i H i ( Ω f 0 ; R 3 ) ]   (with the convention that u 0 = w 0   ), by N ( ( q i ) i = 0 2 )   a generic smooth function depending only on i = 0 2 q 2 i H i ( Ω f 0 ; R 3 )   and by M ( f , κ g , κ h )   a generic smooth function depending only on f V f 3 ( T ) + f V s 3 ( T ) + κ [ u 0 H 4 ( Ω s 0 ; R 3 ) + w 1 H 4 ( Ω s 0 ; R 3 ) + w 2 H 4 ( Ω s 0 ; R 3 ) ]   . Then, let w 3 L 2 ( Ω ; R 3 )   be defined by
w 3 i = ν [ ( a l j a l k w i , k ) , j ] t t ( 0 ) + ( f η ) t t ( 0 ) [ a i j q , j ] t t ( 0 ) in Ω 0 f , (7.5a)
w 3 i = f t t i ( 0 ) + c m j k l [ [ ( η , m η , j δ m j ) η , k ] , l ] t t ( 0 ) in Ω 0 s , (7.5b)
where the time derivatives are computed with any η ( 0 , x ) = x   , w = η t ( 0 ) = u 0   , t n w ( 0 ) = w n , ( n = 1 , 2 )   , t n q ( 0 ) = q n , ( n = 0 , 1 , 2 )   .
Let us now define
b κ ( φ ) = κ ( c i j k l w 2 k , l , φ i , j ) L 2 ( Ω 0 s ; R ) , (7.6a)
c κ ( φ ) = κ ( c i j k l w 1 k , l , φ i , j ) L 2 ( Ω 0 s ; R ) , (7.6b)
d κ ( φ ) = κ ( c i j k l u 0 k , l , φ i , j ) L 2 ( Ω 0 s ; R ) . (7.6c)
By proceding as in [5, we can establish the existence of a fixed-point to the system ( 7.1 ). This follows the lines of [5by first approximating by a penalty scheme the divergence-free constraint in the fluid in our Lagrangian setting, and by performing a regularity analysis of the solution of ( 7.4 ) allowing the use of the Tychonoff fixed-point theorem. Given the estimates obtained in [5, no new difficulty arise, since the parabolic artificial viscosity in the solid controls the forcing coming from the quasilinear part on a short time which is a priori shrinking to zero, and for this reason the proof is omitted here.
This leads us to the following
Lemma 7.1. There exists T κ > 0   depending a priori on κ   and on a given expression of the type N 0 ( u 0 , ( w i ) i = 1 3 ) + N 0 ( ( q i ) i = 0 2 ) + M 0 ( f , κ g , κ h )   , so that there exists a unique solution ( w κ , q κ ) Z T κ   of the regularized problem ( 7.1 ). Moreover, w κ V s 4 ( T κ )   .
In the next section we will study the limit of these solutions of the smoothed problems as κ 0   (this being problematic since the solutions to these regularized problems are a priori defined on a time interval shrinking to zero as κ 0   ).
Moreover, the following variational equations (for n = 0 , 1 , 2 )   are satisfied for any test function φ L 2 ( 0 , T κ ; H 0 1 ( Ω ; R 3 ) )   :
0 T κ ( t n ( w κ ) t , φ ) L 2 ( Ω ; R 3 ) d t + ν 0 T κ ( t n ( a k r a k s w κ , r ) , φ , s ) L 2 ( Ω 0 f ; R 3 ) d t
+ κ 0 T κ ( c i j k l t n w κ k , l , φ i , j ) L 2 ( Ω 0 s ; R ) d t 0 T κ ( t n ( a k l q κ ) , φ k , l ) L 2 ( Ω 0 f ; R ) d t
+ 0 T κ ( c i j k l t n [ ( η , i η , j δ i j ) η , l ] , φ , k ) L 2 ( Ω 0 s ; R 3 ) d t
= 0 T κ ( t n F , φ ) L 2 ( Ω 0 f ; R 3 ) + ( t n f , φ ) L 2 ( Ω 0 s ; R 3 ) + t n [ t 2 2 ] b κ ( φ ) + t n [ t ] c κ ( φ ) + t n [ 1 ] d κ ( φ ) d t , (7.7)
together with the initial conditions w κ ( 0 ) = u 0   , ( w κ ) t ( 0 ) = w 1   , ( w κ ) t t ( 0 ) = w 2   and q κ ( 0 ) = q 0   , ( q κ ) t ( 0 ) = q 1   , ( q κ ) t t ( 0 ) = q 2   . Moreover for the third time differentiated problem in time, we also have that a.e. in ( 0 , T κ )   ,
1 2 ( w κ ) t t t ( t ) L 2 ( Ω ; R 3 ) 2 + ν 0 t ( ( a k r a k s w κ , r ) t t t , ( w κ ) t t t , s ) L 2 ( Ω 0 f ; R 3 )
+ κ 0 t ( c i j k l ( w κ ) t t t k , l ( t ) , ( w κ ) t t t i , j ( t ) ) L 2 ( Ω 0 s ; R )
0 t Ω 0 f ( q κ ) t t [ 3 ( a i j ) t t ( w κ ) t i , j + 3 ( a i j ) t ( w κ ) t t i , j + ( a i j ) t t t w κ i , j ] t
+ Ω 0 f ( q κ ) t t ( t ) [ 3 ( a i j ) t t ( w κ ) t i , j + 3 ( a i j ) t ( w κ ) t t i , j + ( a i j ) t t t w κ i , j ] ( t )
0 t Ω 0 f [ 3 ( a i j ) t t ( q κ ) t ( w κ ) t t t i , j + 3 ( a i j ) t ( q κ ) t t ( w κ ) t t t i , j + ( a i j ) t t t q κ ( w κ ) t t t i , j ]
+ 0 t Ω s 0 c i j k l [ ( η , i η , j δ i j ) η , l ] t t t ( w κ ) t t t , k
N ( u 0 , ( w i ) i = 1 3 ) + N ( ( q i ) i = 0 2 ) + 0 t [ ( F t t t , ( w κ ) t t t ) L 2 ( Ω 0 f ; R 3 ) + ( f t t t , ( w κ ) t t t ) L 2 ( Ω 0 s ; R 3 ) ] , (7.8)
where, recall that C   does not depend on the artificial viscosity κ   . The following result will be fundamental to our proof that the time interval of existence of solutions to ( 7.1 ) is in fact κ   -independent.
Lemma 7.2. The mapping γ : t ( w κ , q κ ) Z t   is continuous on [ 0 , T κ ]   .
  • Proof. The continuity with respect to t   of the terms of the type L 2 ( 0 , t ; H s )   is obvious, and since w κ V s 4 ( T κ )   (due to our artificial viscosity), so is the continuity of n = 0 3 t n η κ L ( 0 , t ; H 4 n ( Ω s 0 ; R 3 ) ) 2   . The only terms that remain are t 3 w κ L ( 0 , t ; L 2 ( Ω ; R 3 ) )   and t 2 q κ L ( 0 , t ; L 2 ( Ω 0 f ; R ) )   .
    In order to treat them, we will invoke the fact that due to our artificial viscosity in the solid, we in fact have t 4 w κ L 2 ( 0 , T κ ; L 2 ( Ω ; R 3 ) )   , which provides t 3 w κ C 0 ( [ 0 , T κ ] ; L 2 ( Ω ; R 3 ) )   . For the second time derivative of the pressure, we notice that from the variational form, true almost everywhere on [ 0 , T κ ]   for any φ H 0 1 ( Ω ; R 3 )   ,
    ( t 3 w κ , φ ) L 2 ( Ω ; R 3 ) + ν ( t 2 ( a k r a k s w κ , r ) , φ , s ) L 2 ( Ω 0 f ; R 3 ) + κ ( c i j k l t 2 w κ k , l , φ i , j ) L 2 ( Ω 0 s ; R )
    ( t 2 ( a k l q κ ) a k l q κ t t , φ k , l ) L 2 ( Ω 0 f ; R ) + ( c i j k l t 2 [ ( η , i η , j δ i j ) η , l ] , φ , k ) L 2 ( Ω 0 s ; R 3 )
    ( t 2 F , φ ) L 2 ( Ω 0 f ; R 3 ) ( t 2 f , φ ) L 2 ( Ω 0 s ; R 3 ) b κ ( φ ) = ( a k l q κ t t , φ k , l ) L 2 ( Ω 0 f ; R ) ,
    and the Lagrange multiplier Lemma 13 of [5associated to the continuity results previously established, we have the continuity of t q κ t t L ( 0 , t ; L 2 ( Ω f 0 ; R ) )   on [ 0 , T κ ]   .
    We now explain briefly why such a control on the fourth time derivative of w ~ κ   holds, and is possible only with the addition of the artificial viscosity in the solid. In particular, this norm cannot be controlled as κ 0   , which is not crucial for our purposes in any case. In order to understand the idea, we return to the level of the setting of the fixed-point argument, where we assume that v   in an appropriate convex set of V f 4 ( T ) × V s 4 ( T )   is given, and search for a solution w   of ( 7.4 ) by a Galerkin approximation on a penalized problem (for the pressure), in a way similar to [5. The penalization parameter ε > 0   is given, and we denote q ε n = n = 0 2 t n n ! q n 1 ε a i j ( w ε n ) i , j   , where w ε n   is solution of the Galerkin approximation at rank n   , and where a i j   is computed from η   associated to the given v   . Our interest will be with the first problem that appears in our methodology in [5; namely, the highest order time-differentiated problem is multiplied by t 4 w ε n   (which is permitted since it belongs to the appropriate finite dimensional space), and then integrate from 0   to t   . We obtain
    0 t t 4 w ε n L 2 ( Ω ; R 3 ) 2 + ν 0 t ( t 3 ( a k r a k s w ε n , r ) , t 4 ( w ε n ) , s ) L 2 ( Ω 0 f ; R 3 )
    + [ κ 2 ( c i j k l t 3 ( w ε n ) k , l , t 3 ( w ε n ) i , j ) L 2 ( Ω 0 s ; R ) ] 0 t 0 t ( t 3 ( a k l q ε n ) , t 4 ( w ε n ) k , l ) L 2 ( Ω 0 f ; R )
    0 t ( c i j k l t 4 [ ( η , i η , j δ i j ) η , l ] , t 3 ( w ε n ) , k ) L 2 ( Ω 0 s ; R 3 )
    + [ Ω s 0 c i j k l t 3 [ ( η , i η , j δ i j ) η , l ] t 3 ( w ε n ) , k ] 0 t
    = 0 t [ Ω f 0 t 3 F t 4 w ε n + Ω s 0 t 3 f t 4 w ε n ] ,
    leading us for a time small enough depending on the artificial viscosity κ   (but not on n   and ε   ) to an inequality of the type,
    0 t t 4 w ε n L 2 ( Ω ; R 3 ) 2 + sup [ 0 , t ] [ t 3 w ε n H 1 ( Ω 0 f ; R 3 ) 2 + κ t 3 w ε n H 1 ( Ω 0 s ; R 3 ) 2 + ε t 3 q ε n L 2 ( Ω 0 f ; R ) 2 ]
    C ε [ N ( u 0 , ( w i ) i = 1 3 ) + N ( ( q i ) i = 0 2 ) + N ( f ) ] ,
    where C ε   depends a priori on ε   . By proceding in a way inspired by our methodology in Section 9 of [5, we can then prove that we have control, independently of ε   , on the first three norms. Taking the limit first as   and then as ε 0   , indeed provides us with t 4 w κ L 2 ( 0 , T κ ; L 2 ( Ω f 0 ; R 3 ) )   as announced.
We note that this latter regularity property in the solid is only possible with the artificial viscosity κ > 0   .

8 Estimate for the solutions of ( 7.4 ) independently of κ  

In this section, we will denote ( w κ , q κ ) = ( w ~ , q ~ )   and denote the corresponding quantities a i j   by a ~ i j   . In what follows, δ > 0   is a given positive number to be made precise later when it will be chosen to be sufficiently small.

8.1 Energy estimate for w ~ t t t   independently of κ   .

We are now going to use the regularity result ( w ~ , q ~ ) Z T κ   in the energy inequality ( 7.8 ) (which was established independently of the artificial viscosity), this time by interpolating and using the energy properties of the nonlinear elasticity operator, in order to get an estimate independent of the artificial viscosity.
Step 1. Let I 1 = 0 t Ω s 0 c i j k l ( η ~ , i η ~ , j δ i j ) w ~ t t , l w ~ t t t , k   . An integration by parts in time shows that
I 1 = 1 2 0 t Ω s 0 c i j k l ( η ~ , i η ~ , j δ i j ) t w ~ t t , l w ~ t t , k
+ 1 2 Ω s 0 c i j k l ( η ~ , i η ~ , j δ i j ) w ~ t t , l w ~ t t , k ( t ) ,
and thus with the properties of the Bochner integral in H 2 ( Ω s 0 ; R )   ,
η ~ , i η ~ , j ( t ) δ i j = 0 t [ η ~ , i w ~ , j + w ~ , i η ~ , j ] ,
we deduce
| I 1 | C t w ~ t t L ( 0 , t ; H 1 ( Ω s 0 ; R 3 ) ) 2 w ~ L ( 0 , t ; H 3 ( Ω s 0 ; R 3 ) ) η ~ L ( 0 , t ; H 3 ( Ω s 0 ; R 3 ) )
C t w ~ W t 4 . (8.1)
Step 2. Let I 2 = 3 0 t Ω s 0 c i j k l ( η ~ , i η ~ , j δ i j ) t w ~ t , l w ~ t t t , k   . Similarly,
I 2 = 3 0 t Ω s 0 c i j k l [ ( η ~ , i η ~ , j δ i j ) t w ~ t t , l w ~ t t , k + ( η ~ , i η ~ , j δ i j ) t t w ~ t , l w ~ t t , k ]
+ 3 [ Ω s 0 c i j k l ( η ~ , i η ~ , j δ i j ) t w ~ t , l w ~ t t , k ( t ) ] 0 t .
By the same type of argument as for the previous step, we then get
| I 2 | C t w ~ t t L ( 0 , t ; H 1 ( Ω s 0 ; R 3 ) ) 2 w ~ L ( 0 , t ; H 3 ( Ω s 0 ; R 3 ) ) η ~ L ( 0 , t ; H 3 ( Ω s 0 ; R 3 ) )
+ C t w ~ t t L ( 0 , t ; H 1 ( Ω s 0 ; R 3 ) ) w ~ L ( 0 , t ; H 3 ( Ω s 0 ; R 3 ) ) 2 w ~ t L ( 0 , t ; H 1 ( Ω s 0 ; R 3 ) )
+ C t w ~ t t L ( 0 , t ; H 1 ( Ω s 0 ; R 3 ) ) w ~ t L ( 0 , t ; H 2 ( Ω s 0 ; R 3 ) ) 2 η ~ L ( 0 , t ; H 3 ( Ω s 0 ; R 3 ) )
+ C c i j k l [ ( u 0 i , j + u 0 j , i ) w 1 , l + 0 ( ( η ~ , i η ~ , j δ i j ) t w ~ t , l ) t ] L ( 0 , t ; L 2 ( Ω s 0 ; R 3 ) )
× w ~ , k t t L ( 0 , t ; L 2 ( Ω s 0 ; R 3 ) )
+ N ( u 0 , ( w i ) i = 1 3 ) ,
and thus,
| I 2 | δ w ~ W t 2 + C t w ~ W t 4 + C δ N ( u 0 , ( w i ) i = 1 3 ) . (8.2)
Step 3. Let I 3 = 3 0 t Ω s 0 c i j k l ( η ~ , i η ~ , j δ i j ) t t w ~ , l w ~ t t t , k   . By an integration by parts in time,
I 3 = 3 0 t Ω s 0 c i j k l [ ( η ~ , i η ~ , j δ i j ) t t w ~ t , l w ~ t t , k + ( η ~ , i η ~ , j δ i j ) t t t w ~ , l w ~ t t , k ]
+ [ Ω s 0 c i j k l ( η ~ , i η ~ , j δ i j ) t t w ~ , l w ~ t t , k ( t ) ] 0 t .
Similarly as before, we get
| I 3 | C t w ~ t t L ( 0 , T ; H 1 ( Ω s 0 ; R 3 ) ) w ~ L ( 0 , t ; H 3 ( Ω s 0 ; R 3 ) ) 2 w ~ t L ( 0 , t ; H 1 ( Ω s 0 ; R 3 ) )
+ C t w ~ t t L ( 0 , T ; H 1 ( Ω s 0 ; R 3 ) ) w ~ t L ( 0 , t ; H 2 ( Ω s 0 ; R 3 ) ) 2 η ~ L ( 0 , t ; H 3 ( Ω s 0 ; R 3 ) )
+ C t w ~ t t L ( 0 , T ; H 1 ( Ω s 0 ; R 3 ) ) 2 w ~ L ( 0 , t ; H 3 ( Ω s 0 ; R 3 ) ) η ~ L ( 0 , t ; H 3 ( Ω s 0 ; R 3 ) )
+ C t w ~ t t L ( 0 , T ; H 1 ( Ω s 0 ; R 3 ) ) w ~ t L ( 0 , t ; H 1 ( Ω s 0 ; R 3 ) ) w ~ L ( 0 , t ; H 3 ( Ω s 0 ; R 3 ) ) 2
+ C c i j k l ( ( η ~ , i η ~ , j ) t t ( 0 ) w 1 , l + 0 ( ( η ~ , i η ~ , j δ i j ) t t w ~ , l ) t ) L ( 0 , t ; L 2 ( Ω s 0 ; R 3 ) )
× w ~ t t , k L ( 0 , t ; L 2 ( Ω s 0 ; R 3 ) )
+ N ( u 0 , ( w i ) i = 1 3 ) ,
and therefore
| I 3 | δ w ~ W t 2 + C t w ~ W t 4 + C δ N ( u 0 , ( w i ) i = 1 3 ) . (8.3)
Step 4. Let I 4 = 0 t Ω s 0 c i j k l ( η ~ , i η ~ , j δ i j ) t t t η ~ , l w ~ t t t , k   . By symmetry of c   , we notice that
I 4 = 1 2 0 t Ω s 0 c i j k l ( η ~ , i η ~ , j ) t t t ( η ~ , l η ~ , k ) t t t t
0 t Ω s 0 c i j k l [ 4 ( η ~ , i η ~ , j ) t t t ( w ~ , l w ~ t t , k ) + 3 ( η ~ , i η ~ , j ) t t t ( w ~ t , l w ~ t , k ) ] ,
and thus,
| I 4 1 4 Ω s 0 c i j k l ( η ~ , i η ~ , j ) t t t ( η ~ , l η ~ , k ) t t t ( t ) |
C t w ~ t t L ( 0 , t ; L 2 ( Ω s 0 ; R 9 ) ) 2 w ~ L ( 0 , t ; H 2 ( Ω s 0 ; R 9 ) ) η ~ L ( 0 , t ; H 2 ( Ω s 0 ; R 9 ) )
+ C t w ~ t L ( 0 , t ; L 2 ( Ω s 0 ; R 9 ) ) w ~ t t L ( 0 , t ; L 2 ( Ω s 0 ; R 9 ) ) w ~ L ( 0 , t ; H 2 ( Ω s 0 ; R 9 ) ) 2
+ C t w ~ t L ( 0 , t ; H 1 ( Ω s 0 ; R 9 ) ) 2 w ~ t t L ( 0 , t ; L 2 ( Ω s 0 ; R 9 ) ) η ~ L ( 0 , t ; H 2 ( Ω s 0 ; R 9 ) )
+ C t w ~ L ( 0 , t ; H 2 ( Ω s 0 ; R 9 ) ) w ~ t L ( 0 , t ; H 1 ( Ω s 0 ; R 9 ) ) 3 + N ( u 0 , ( w i ) i = 1 3 )
N ( u 0 , ( w i ) i = 1 3 ) + C t w ~ W t 4 . (8.4)
Now, Let I = 1 4 Ω s 0 c i j k l ( η ~ , i η ~ , j ) t t t ( η ~ , l η ~ , k ) t t t ( t )   . By expanding the integrand with respect to the time derivatives and using the relation in H 3 ( Ω s 0 ; R 3 )   : η ~ ( t , ) = Id + 0 t w ~ ( t , ) d t   and estimates similar as in the previous steps, we find that
| I Ω s 0 c i j k l w ~ t t j , i w ~ t t k , l ( t ) | C δ N ( u 0 , ( w i ) i = 1 3 ) + δ w ~ W t 2 + C t w ~ W t 4 . (8.5)
Step 5. By using ( 8.4 ) and ( 8.5 ) we find that
| I 4 Ω s 0 c i j k l w ~ t t j , i w ~ t t k , l ( t ) | C δ N ( u 0 , ( w i ) i = 1 3 ) + δ w ~ W t 2 + C t w ~ W t 4 . (8.6)
Step 6. By proceding in a way similar to [5, except that we replace the constants C ( M )   appearing there by appropriate powers of ( w ~ , q ~ ) Z t   , we find that the integrals set in the fluid domain are bounded by
δ ( w ~ , q ~ ) Z t 2 + C δ [ N ( ( q i ) i = 0 2 ) + N ( u 0 , ( w i ) i = 1 3 ) + M ( f , κ g , κ h ) + t 1 4 ( w ~ , q ~ ) Z t 6 ] .
Step 7. Thus, from ( 7.8 ), and Steps 1-6, we then get on [ 0 , T κ ]   :
sup [ 0 , t ] w ~ t t t L 2 ( Ω ; R 3 ) 2 + 0 t [ w ~ t t t H 1 ( Ω 0 f ; R 3 ) 2 + κ w ~ t t t H 1 ( Ω 0 s ; R 3 ) 2 ] + sup [ 0 , t ] w ~ t t H 1 ( Ω 0 s ; R 3 ) 2
C δ [ N ( u 0 , ( w i ) i = 1 3 ) + M ( f , κ g , κ h ) + N ( ( q i ) i = 0 2 ) + t 1 4 ( w ~ , q ~ ) Z t 6 ] + C δ ( w ~ , q ~ ) Z t 2 . (8.7)
Step 8. The estimate of q ~ t t   in L ( L 2 )   , independently of κ   , will require some adjustments with respect to the methodology of [5. To this end, we notice that we can apply a Lagrange multiplier Lemma similar to Lemma 13 of [5, but corresponding to the case a i j = δ i j   , to the variational form true on [ 0 , T κ ]   : for all φ H 0 1 ( Ω ; R 3 )   ,
( w ~ t t t , φ ) L 2 ( Ω ; R 3 ) + ν ( ( a ~ k r a ~ k s w ~ , r ) t t , φ , s ) L 2 ( Ω 0 f ; R 3 ) + κ ( c i j k l w ~ t t k , l , φ i , j ) L 2 ( Ω 0 s ; R )
+ ( c i j k l [ ( η ~ , i η ~ , j δ i j ) η ~ , l ] t t , φ , k ) L 2 ( Ω 0 s ; R 3 ) ( ( a ~ k l δ k l ) q ~ t t , φ k , l ) L 2 ( Ω 0 f ; R )
( ( a ~ k l q ~ ) t t a ~ k l q ~ t t , φ k , l ) L 2 ( Ω 0 f ; R ) ( F t t , φ ) L 2 ( Ω 0 f ; R 3 ) ( f t t , φ ) L 2 ( Ω 0 s ; R 3 ) b κ ( φ )
= ( q ~ t t , div φ ) L 2 ( Ω 0 f ; R ) ,
which provides for any t [ 0 , T κ ]   :
q ~ t t L 2 ( Ω 0 f ; R ) C [ w ~ t t t L 2 ( Ω ; R 3 ) + ( a ~ k r a ~ k s w ~ , r ) t t L 2 ( Ω 0 f ; R 3 ) + κ w ~ t t H 1 ( Ω 0 s ; R )
+ [ ( η ~ , i η ~ , j δ i j ) η ~ , l ] t t L 2 ( Ω 0 s ; R 3 ) + ( a ~ k l q ~ ) t t a ~ k l q ~ t t L 2 ( Ω ; R )
+ a ~ Id H 2 ( Ω f 0 ; R 9 ) q ~ t t L 2 ( Ω f 0 ; R ) + N ( u 0 , ( w i ) i = 1 3 ) + M ( f , κ g , κ h ) ] .
By using ( 8.7 ) for the first four terms of the right-hand side of this inequality and remembering that the L ( 0 , t ; L 2 ( Ω f 0 ; R ) )   norm of q ~ t t   is part of the norm Z t   for the next two terms of this inequality, we get
sup [ 0 , t ] q ~ t t L 2 ( Ω 0 f ; R ) 2 C δ [ N ( u 0 , ( w i ) i = 1 3 ) + M ( f , κ g , κ h ) + N ( ( q i ) i = 0 2 ) ]
+ C δ t 1 4 ( w ~ , q ~ ) Z t 6 + C δ ( w ~ , q ~ ) Z t 2 . (8.8)

8.2 Estimate on w t t   and w t   .

From the previous estimates, and the arguments that we will see hereafter for the case of w ~   , we have
w ~ t t L 2 ( 0 , t ; H 2 ( Ω f 0 ; R 3 ) ) 2 + q ~ t t L 2 ( 0 , t ; H 1 ( Ω f 0 ; R ) ) 2 + w ~ t L ( 0 , t ; H 2 ( Ω s 0 ; R 3 ) ) 2
C δ [ N ( u 0 , ( w i ) i = 1 3 ) + M ( f , κ g , κ h ) + N ( ( q i ) i = 0 2 ) ]
+ C δ t 1 4 ( w ~ , q ~ ) Z t 6 + C δ ( w ~ , q ~ ) Z t 2 . (8.9)
Similarly, we infer from ( 8.9 ) that
w ~ t L 2 ( 0 , t ; H 3 ( Ω f 0 ; R 3 ) ) 2 + q ~ t L 2 ( 0 , t ; H 2 ( Ω f 0 ; R ) ) 2 + w ~ L ( 0 , t ; H 3 ( Ω s 0 ; R 3 ) ) 2
C δ [ N ( u 0 , ( w i ) i = 1 3 ) + M ( f , κ g , κ h ) + N ( ( q i ) i = 0 2 ) ]
+ C δ t 1 4 ( w ~ , q ~ ) Z t 6 + C δ ( w ~ , q ~ ) Z t 2 . (8.10)

8.3 Estimate on w ~  

We will denote R + 3 = { x R 3 | x 3 > 0 }   , R 3 = { x R 3 | x 3 < 0 }   , and B ( 0 , r ) = B ( 0 , r ) R 3   . We denote by Ψ   an H 4   diffeomorphism from B ( 0 , 1 )   into a neighborhood V   of a point x 0 Γ 0   such that Ψ ( B ( 0 , 1 ) R + 3 ) = V Ω f 0   , Ψ ( B ( 0 , 1 ) R 3 ) = V Ω s 0   , Ψ ( B ( 0 , 1 ) R 2 × { 0 } ) = V Γ 0   , with det Ψ = 1   . We consider a cut-off function ζ   compactly supported in B ( 0 , 1 )   , and equal to 1   in B ( 0 , 1 2 )   .
With the use of test functions φ p = [ ρ p ( ζ 2 w ~ Ψ ) ] , α 1 α 1 α 2 α 2 α 3 α 3 Ψ 1   (which is in L 2 ( 0 , T κ ; H 0 1 ( Ω ; R 3 ) )   ) in ( 7.7 ) for n = 0   , and by denoting W = w ~ Ψ   , Q = q ~ Ψ   , E = η ~ Ψ   , we get after integrating by parts appropriately and letting p   ,
1 2 ζ W , α 1 α 2 α 3 ( t ) L 2 ( R 3 ; R 3 ) 2 + 0 t ( W t , α 1 α 2 α 3 , [ ζ 2 W ] , α 1 α 2 α 3 ζ 2 W , α 1 α 2 α 3 ) L 2 ( R 3 ; R 3 )
+ ν 0 t ( [ b ~ k r b ~ k s W , r ] , α 1 α 2 α 3 , [ ζ 2 W ] , s α 1 α 2 α 3 ) L 2 ( R + 3 ; R 3 )
0 t R + 3 [ Q b ~ i j ] , α 1 α 2 α 3 [ ζ 2 W ] i , j α 1 α 2 α 3
+ κ 0 t ( [ C i j k l ( W , i Ψ , j ) Ψ , l ] , α 1 α 2 α 3 , [ ζ 2 W ] , k α 1 α 2 α 3 ) L 2 ( R 3 ; R 3 )
+ 0 t ( [ C i j k l ( E , i E , j Ψ , i Ψ , j ) E , l ] , α 1 α 2 α 3 , [ ζ 2 W ] , k α 1 α 2 α 3 ) L 2 ( R 3 ; R 3 )
C N ( u 0 , ( w i ) i = 1 3 ) + lim p 0 t [ ( F , φ p ) L 2 ( Ω 0 f ; R 3 ) + ( f , φ p ) L 2 ( Ω 0 s ; R 3 ) ]
+ lim p 0 t [ t 2 2 b κ ( φ p ) + t c κ ( φ p ) + d κ ( φ p ) ] , (8.11)
where C i j k l = c m n o p g m i g n j g o k g p l H 3 ( B ( 0 , 1 ) ; R )   , g = [ Ψ ] 1 H 3 ( B ( 0 , 1 ) ; R 9 )   , b ~ l j = a ~ l k ( Ψ ) g k j   .
Remark 6. Note that this limit process as p   for the nonlinear elastic energy is possible because t n w ~ L 2 ( 0 , T κ ; H 4 n ( Ω s 0 ; R 3 ) )   ( n = 0 , 1   ) due to our artificial viscosity in the solid. Whereas we could also use difference quotients, it appears that the product rules are less cumbersome with the use of horizontal derivatives instead, which is permitted since we already know at this stage the regularity of w ~   and q ~   . Also, the limits on the right-hand side of ( 8.11 ) do not present any difficulty, given the regularity of the forcing functions and three integrations by parts with respect to horizontal variables.
Remark 7. Since ζ   is compactly supported in B ( 0 , 1 )   , the integrals set on R 3   , R 3   , R + 3   do not depend on the extension that we chose for W   , E   or Q   , and simply represent a more convenient way to write these integrals.
Step 1. Let L 1 = κ 0 t ( [ C i j k l W , i Ψ , j Ψ , l ] , α 1 α 2 α 3 , [ ζ 2 W ] , k α 1 α 2 α 3 ) L 2 ( R 3 ; R 3 )   . By using the H 3   regularity of the coefficients C i j k l   ,
L 1 = κ 0 t ( C i j k l W , i α 1 α 2 α 3 Ψ , j Ψ , l , ζ 2 W , k α 1 α 2 α 3 ) L 2 ( R 3 ; R 3 )
+ κ 0 t ( [ C i j k l W , i Ψ , j Ψ , l ] , α 1 α 2 α 3 C i j k l W , i α 1 α 2 α 3 Ψ , j Ψ , l , ζ 2 W , k α 1 α 2 α 3 ) L 2 ( R 3 ; R 3 )
+ κ 0 t ( [ C i j k l W , i Ψ , j Ψ , l ] , α 1 α 2 α 3 , [ ζ 2 W , k ] , α 1 α 2 α 3 ζ 2 W , k α 1 α 2 α 3 ) L 2 ( R 3 ; R 3 )
C κ 0 t ( C i j k l W , i α 1 α 2 α 3 Ψ , j Ψ , l , ζ 2 W , k α 1 α 2 α 3 ) L 2 ( R 3 ; R 3 ) C κ 0 t w ~ H 3 ( Ω s 0 ; R 3 ) 2
C κ 0 t W , α 1 α 2 α 3 H 1 ( B ( 0 , 1 2 ) ; R 3 ) 2 C κ t w ~ W t 2 . (8.12)
Step 2. Let L 2 = 0 t ( [ C i j k l ( E , i E , j Ψ , i Ψ , j ) E , l ] , α 1 α 2 α 3 , [ ζ 2 E t ] , k α 1 α 2 α 3 ) L 2 ( R 3 ; R 3 ) .   Σ 3   denoting the set of permutations of { 1 , 2 , 3 }   ,
L 2 = 0 t ( C i j k l ( E , i E , j Ψ , i Ψ , j ) E , l α 1 α 2 α 3 , ζ 2 E t , k α 1 α 2 α 3 ) L 2 ( R 3 ; R 3 )
+ 2 0 t ( ζ 2 C i j k l ( E , i α 1 α 2 α 3 E , j ) E , l , E t , k α 1 α 2 α 3 ) L 2 ( R 3 ; R 3 )
+ 0 t ( ζ 2 [ [ C i j k l ( E , i E , j ) ] , α 1 α 2 α 3 2 C i j k l ( E , i α 1 α 2 α 3 E , j ) ] E , l , E t , k α 1 α 2 α 3 ) L 2 ( R 3 ; R 3 )
+ 0 t ( [ ζ 2 ( C i j k l Ψ , i Ψ , j ) , α 1 α 2 α 3 E , l ] t , E , k α 1 α 2 α 3 ) L 2 ( R 3 ; R 3 )
[ ( ζ 2 ( C i j k l Ψ , i Ψ , j ) , α 1 α 2 α 3 E , l , E , k α 1 α 2 α 3 ) L 2 ( R 3 ; R 3 ) ] 0 t
σ Σ 3 0 t ( [ ζ 2 [ C i j k l ( E , i E , j Ψ , i Ψ , j ) ] , α σ ( 1 ) E , α σ ( 2 ) α σ ( 3 ) l ] , α 1 , E t , α 2 α 3 k ) L 2 ( R 3 ; R 3 )
σ Σ 3 0 t ( [ ζ 2 [ C i j k l ( E , i E , j Ψ , i Ψ , j ) ] , α σ ( 1 ) α σ ( 2 ) E , α σ ( 3 ) l ] , α 1 , E t , α 2 α 3 k ) L 2 ( R 3 ; R 3 )
+ 0 t ( [ C i j k l ( E , i E , j Ψ , i Ψ , j ) E , l ] , α 1 α 2 α 3 , [ ζ 2 E t ] , α 1 α 2 α 3 k ζ 2 E t , α 1 α 2 α 3 k ) L 2 ( R 3 ; R 3 ) .
From the regularity of w ~   and the H 4   regularity of Ψ   , we then infer
L 2 = 0 t ( C i j k l ( E , i E , j Ψ , i Ψ , j ) E , α 1 α 2 α 3 l , ζ 2 E t , α 1 α 2 α 3 k ) L 2 ( R 3 ; R 3 )
+ 2 0 t ( C i j k l E , α 1 α 2 α 3 i E , j ζ 2 E , l , E t , α 1 α 2 α 3 k ) L 2 ( R 3 ; R 3 )
+ 0 t ( C , α 1 α 2 α 3 i j k l E , i E , j ζ 2 E , l , E t , α 1 α 2 α 3 k ) L 2 ( R 3 ; R 3 ) + L 2 r ,
with
| L 2 r | δ w ~ W t 2 + C t w ~ W t 4 + C δ N ( u 0 , ( w i ) i = 1 3 ) . (8.13)
By integrating by parts in time, we deduce
L 2 = 1 2 0 t ( C i j k l ( E , i E , j Ψ , i Ψ , j ) t E , α 1 α 2 α 3 l , ζ 2 E , α 1 α 2 α 3 k ) L 2 ( R 3 ; R 3 )
+ [ 1 2 ( C i j k l ( E , i E , j Ψ , i Ψ , j ) E , α 1 α 2 α 3 l , ζ 2 E , α 1 α 2 α 3 k ) L 2 ( R 3 ; R 3 ) ] 0 t
2 0 t ( C i j k l E , α 1 α 2 α 3 i E , j ζ 2 E t , l , E , α 1 α 2 α 3 k ) L 2 ( R 3 ; R 3 )
+ [ ( C i j k l E , α 1 α 2 α 3 i E , j ζ 2 E , l , E , α 1 α 2 α 3 k ) L 2 ( R 3 ; R 3 ) ] 0 t
0 t ( C , α 1 α 2 α 3 i j k l ( E , i E , j ζ 2 E , l ) t , E , α 1 α 2 α 3 k ) L 2 ( R 3 ; R 3 )
+ [ ( C , α 1 α 2 α 3 i j k l E , i E , j ζ 2 E , l , E , α 1 α 2 α 3 k ) L 2 ( R 3 ; R 3 ) ] 0 t + L 2 r ,
which implies in turn
| L 2 ( C i j k l ζ 2 E , α 1 α 2 α 3 i ( t ) Ψ , j Ψ , l , E , α 1 α 2 α 3 k ( t ) ) L 2 ( R 3 ; R 3 ) |
δ w ~ W t 2 + C δ N ( u 0 , ( w i ) i = 1 3 ) + C t w ~ W t 4 . (8.14)
With e k   ( k = 1 , 2 , 3   ) denoting the canonical vectors of R 3   , let P ( t ) = 2 c i j k l ( η ~ , m n i l e j ) e k [ c i j k l ( η ~ , i η ~ j δ i j ) η ~ , k ] , l m n L 2 ( Ω s 0 ; R 3 ) ( t ) ,   where m   and n   are arbitrarily fixed in { 1 , 2 , 3 }   . We then have P ( t ) P 1 ( t ) + P 2 ( t ) ,   with
P 1 ( t ) = 2 c i j k l ( η ~ , i l m n e , j ) e k 2 [ c i j k l ( η ~ , m n i η ~ , j ) η ~ , k ] , l L 2 ( Ω s 0 ; R 3 ) ( t ) ,
P 2 ( t ) = [ c i j k l ( η ~ , i η ~ , j δ i j ) η ~ , k ] , l m n + 2 [ c i j k l ( η ~ , m n i η ~ , j ) η ~ , k ] , l L 2 ( Ω s 0 ; R 3 ) ( t ) .
We first notice that
P 1 ( t ) 2 c i j k l [ ( η ~ , i l m n e j ) e k ( η ~ , i l m n η ~ , j ) η ~ , k ] L 2 ( Ω s 0 ; R 3 ) ( t )
+ 2 c i j k l [ η ~ , i m n ( η ~ , j ] η ~ , k ) , l L 2 ( Ω s 0 ; R 3 ) ( t ) .
Next, by writing η ~ ( t ) = Id + 0 t w ~   and [ η ~ , m n i ( η ~ , j ] η ~ , k ) , l ( t ) = [ η ~ , m n i ( η ~ , j ] η ~ , k ) , l ( 0 ) + 0 t [ [ η ~ , m n i ( η ~ , j ] η ~ , k ) , l ] t   respectively in H 3 ( Ω s 0 ; R 3 )   and L 2 ( Ω s 0 ; R 3 )   , we obtain
P 1 ( t ) C [ 0 t w ~ H 3 ( Ω s 0 ; R 3 ) ] sup [ 0 , t ] [ η ~ H 4 ( Ω s 0 ; R 3 ) 2 + η ~ H 4 ( Ω s 0 ; R 3 ) ]
+ N ( u 0 , ( w i ) i = 1 3 ) + C 0 t w ~ H 3 ( Ω s 0 ; R 3 ) sup [ 0 , t ] η ~ H 4 ( Ω s 0 ; R 3 ) 2
N ( u 0 , ( w i ) i = 1 3 ) + C t w ~ W t 3 .
Next, we see that
P 2 ( t ) c i j k l [ ( η ~ , i η ~ , j δ i j ) η ~ , k m n + ( η ~ , i η ~ , j ) , m η ~ , k n + ( η ~ , i η ~ , j ) , n η ~ , k m ] , l L 2 ( Ω s 0 ; R 3 ) ( t )
+ c i j k l [ ( η ~ , m i η ~ , n j + η ~ , i n η ~ , j m ) η ~ , k ] , l L 2 ( Ω s 0 ; R 3 ) ( t ) ,
and by the same type of arguments as for P 1 ( t )   ,
P 2 ( t ) N ( u 0 , ( w i ) i = 1 3 ) + C t w ~ W t 3 ,
implying
P ( t ) N ( u 0 , ( w i ) i = 1 3 ) + C t w ~ W t 3 . (8.15)
Now, from the definition of a solution to the smoothed problem ( 7.1 ),
κ c i j k l ( η ~ t , i l m n e j ) e k + [ c i j k l ( η ~ , i η ~ j δ i j ) η ~ , k ] , m n l L 2 ( Ω s 0 ; R 3 ) ( t )
= ( w ~ t f κ h ) , m n ( t ) L 2 ( Ω s 0 ; R 3 ) ,
which implies with ( 8.15 )
κ 2 L ( w ~ , n m ) + L ( η ~ , n m ) L 2 ( Ω s 0 ; R 3 ) ( t ) ( w ~ t f κ h ) , n m ( t ) L 2 ( Ω s 0 ; R 3 )
+ N ( u 0 , ( w i ) i = 1 3 ) + C t w ~ W t 3 .
Since this inequality also holds for any t ( 0 , t )   , Lemma  6.1 provides
L ( η ~ , n m ) L ( 0 , t ; L 2 ( Ω s 0 ; R 3 ) ) C w ~ t , m n L ( 0 , t ; L 2 ( Ω s 0 ; R 3 ) ) + C M ( f , κ g , κ h )
+ N ( u 0 , ( w i ) i = 1 3 ) + C t w ~ W t 3 ,
which with the estimate on w t   from the previous subsection leads to
L ( η ~ , n m ) L ( 0 , t ; L 2 ( Ω s 0 ; R 3 ) ) C δ [ N ( u 0 , ( w i ) i = 1 3 ) + M ( f , κ g , κ h ) + N ( ( q i ) i = 0 2 ) ]
+ C δ t 1 4 ( w ~ , q ~ ) Z t 6 + C δ ( w ~ , q ~ ) Z t 2 . (8.16)
Step 3. From the estimates on L 1 L 2   , and similar estimates that we could get in the fluid as in [5, but this time by replacing C ( M )   by appropriate powers of ( w ~ , q ~ ) Z t   , we then deduce that for all t [ 0 , T ˇ ]   ,
1 2 ζ W , α 1 α 2 α 3 ( t ) L 2 ( R + 3 ; R 3 ) 2 + ν 0 t ( ζ 2 b k r W , α 1 α 2 α 3 r , b k s W , α 1 α 2 α 3 s ) L 2 ( R + 3 ; R 3 )
C δ [ N ( u 0 , ( w i ) i = 1 3 ) + M ( f , κ g , κ h ) + N ( ( q i ) i = 0 2 ) ]
+ C δ t 1 4 ( w ~ , q ~ ) Z t 6 + C δ ( w ~ , q ~ ) Z t 2 .
By the trace theorem, we then get
0 t W H 3.5 ( S ; R 3 ) 2 C δ [ N ( u 0 , ( w i ) i = 1 3 ) + M ( f , κ g , κ h ) + N ( ( q i ) i = 0 2 ) ]
+ C δ t 1 4 ( w ~ , q ~ ) Z t 6 + C δ ( w ~ , q ~ ) Z t 2 ,
where S = { ( x 1 , x 2 , x 3 ) R 3 | | x 1 | 1 2 , | x 2 | 1 2 , x 3 = 0 }   . By a finite covering argument, we then get
0 t w ~ H 3.5 ( Γ 0 ; R 3 ) 2 C δ [ N ( u 0 , ( w i ) i = 1 3 ) + M ( f , κ g , κ h ) + N ( ( q i ) i = 0 2 ) ]
+ C δ t 1 4 ( w ~ , q ~ ) Z t 6 + C δ ( w ~ , q ~ ) Z t 2 . (8.17)
From the estimate ( 8.10 ) on w ~ t   and the trace estimate ( 8.17 ), we infer in a way similar to [5by elliptic regularity arguments that
w ~ L 2 ( 0 , t ; H 4 ( Ω f 0 ; R 3 ) ) 2 + q ~ L 2 ( 0 , t ; H 3 ( Ω f 0 ; R ) ) 2
C δ [ N ( u 0 , ( w i ) i = 1 3 ) + M ( f , κ g , κ h ) + N ( ( q i ) i = 0 2 ) ]
+ C δ t 1 4 ( w ~ , q ~ ) Z t 6 + C δ ( w ~ , q ~ ) Z t 2 . (8.18)
Similarly, from ( 8.16 ), and the trace estimate ( 8.17 ), elliptic regularity provides
η ~ L ( 0 , t ; H 4 ( Ω s 0 ; R 3 ) ) 2 C δ [ N ( u 0 , ( w i ) i = 1 3 ) + M ( f , κ g , κ h ) + N ( ( q i ) i = 0 2 ) ]
+ C δ t 1 4 ( w ~ , q ~ ) Z t 6 + C δ ( w ~ , q ~ ) Z t 2 . (8.19)

9 Time of existence independent of κ  

From ( 8.7 ), ( 8.8 ), ( 8.9 ), ( 8.10 ), ( 8.18 ) and ( 8.19 ), we then have for any t [ 0 , T κ ]   ,
( w ~ , q ~ ) Z t 2 C δ [ N 0 ( u 0 , ( w i ) i = 1 3 ) + M 0 ( f , κ g , κ h ) ) + N 0 ( ( q i ) i = 0 2 ) ]
+ C δ t 1 4 ( w ~ , q ~ ) Z t 6 + C 0 δ ( w ~ , q ~ ) Z t 2 .
The subscripts 0   in C 0   , N 0   , M 0   mean that we no longer consider generic constants from now on.
Now, let δ 0 > 0   be such that C 0 δ 0 = 1 2   . For κ > 0   small enough, for any t ( 0 , T κ )   we have
( w ~ , q ~ ) Z t 2 4 C δ 0 [ N 0 ( u 0 , ( w i ) i = 1 3 ) + M 0 ( f ) + N 0 ( ( q i ) i = 0 2 ) ] + 2 C δ 0 t 1 4 ( w ~ , q ~ ) Z T 6 , (9.1)
where M 0 ( f ) = M 0 ( f , 0 , 0 )   . For conciseness, we will denote C 1 = 2 C δ 0   and N 1 = 4 C δ 0 [ N 0 ( u 0 , ( w i ) i = 1 3 ) + M 0 ( f ) + N 0 ( ( q i ) i = 0 2 ) ]   .
Now for t ( 0 , T κ )   fixed, let α t ( x ) = x 3 x C 1 t 1 4 + N 1 C 1 t 1 4   , so that α t ( ( w ~ , q ~ ) Z t 2 ) 0 .   Now let t 1 = [ 2 27 C 1 N 1 2 ] 4 > 0   , which does not depend on κ   , and let T ˇ = min ( T κ , t 1 )   .
From now on, we assume that t ( 0 , T ˇ )   . We then have α t ( ( 3 C 1 t 1 4 ) 1 2 ) < 0   which implies that α t   has three real roots z 1   , z 2   , z 3   , with z 1 < ( 3 C 1 t 1 4 ) 1 2 < z 2 < ( 3 C 1 t 1 4 ) 1 2 < z 3   . From the product z 1 z 2 z 3 = N 1 C 1 t 1 4   and α t ( 3 N 1 ) < 0   , we infer that 0 < z 2 < 3 N 1 < z 3   . From ( 9.1 ) and the continuity of t ( w ~ , q ~ ) Z t   (established in Lemma  7.2 ) we then infer since ( w ~ , q ~ ) Z 0 2 N 1 < z 3   that we have
t ( 0 , T ˇ ] , ( w ~ , q ~ ) Z t 2 z 2 3 N 1 . (9.2)
This implies that η ~ ( T ˇ ) H 4 ( Ω f 0 ; R 3 ) H 4 ( Ω s 0 ; R 3 )   , w ~ ( T ˇ ) H 0 1 ( Ω ; R 3 ) H 3 ( Ω s 0 ; R 3 ) H 3 ( Ω s 0 ; R 3 )   , w ~ t ( T ˇ ) H 0 1 ( Ω ; R 3 ) H 2 ( Ω s 0 ; R 3 ) H 2 ( Ω s 0 ; R 3 )   , w ~ t t ( T ˇ ) H 0 1 ( Ω ; R 3 )   , w ~ t t t ( T ˇ ) L 2 ( Ω ; R 3 )   , q ~ ( T ˇ ) H 2 ( Ω 0 f ; R )   , q ~ t ( T ˇ ) H 1 ( Ω 0 f ; R )   , q ~ t t ( T ˇ ) L 2 ( Ω 0 f ; R )   , with a bound that depends only on the right-hand side of ( 9.2 ). The compatibility conditions for the smoothed problem ( 7.1 ) at T ˇ   are also satisfied by definition of a solution, which means that we do not have any new term of the type of b κ   , c κ   or d κ   associated to w ~ ( T ˇ )   to add to the already existing forcing terms coming from t = 0   .
We can thus build a solution of the smoothed problem ( 7.1 ) defined on [ T ˇ , T ˇ + δ T ]   , δ T   depending solely on the right-hand side of ( 9.2 ), that we will still denote ( w ~ , q ~ )   .
It is then readily seen that ( w ~ , q ~ ) Z T ˇ + δ T   and is a solution of the approximated problem ( 7.1 ) on [ 0 , T ˇ + δ T ]   . If T ˇ = t 1   , we have our solution defined on the κ   independent time interval [ 0 , t 1 ]   , with the κ   independent estimate ( 9.2 ). Otherwise, if T ˇ < t 1   , we can also assume that T ˇ + δ T t 1   , which implies, in the same fashion as we got ( 9.1 ),
t [ 0 , T ~ + δ T ] , ( w ~ , q ~ ) Z t 2 N 1 + C 1 t 1 4 ( w ~ , q ~ ) Z t 6 . (9.3)
This implies in turn that η ~ ( T ˇ + δ T )   , w ~ ( T ˇ + δ T ) , w ~ t ( T ˇ + δ T ) , w ~ t t ( T ˇ + δ T ) , w ~ t t t ( T ˇ + δ T )   , q ~ ( T ˇ + δ T ) , q ~ t ( T ˇ + δ T ) , q ~ t t ( T ˇ + δ T )   are in the same spaces as their respective counterparts at time T ˇ   , with the same bound as well, since we could from ( 9.3 ) repeat the same argument leading to ( 9.2 ), this time on [ 0 , T ˇ + δ T ]   . The compatibility conditions at T ˇ + δ T   being also automatically satisfied, we can thus build a solution of the approximated problem ( 7.1 ) defined on [ T ˇ + δ T , T ˇ + 2 δ T ]   , the time of existence being the same as starting from T ˇ   from the similarity of the bound that we obtain on η ~ ( T ˇ + δ T ) , t n w ~ ( T ˇ + δ T ) ( n = 0 , 1 , 2 , 3 ) , t n q ~ ( T ˇ + δ T ) ( n = 0 , 1 , 2 )   and their respective counterparts at time T ˇ   . We will still denote this solution ( w ~ , q ~ )   . It is then readily seen that ( w ~ , q ~ ) Z T ˇ + 2 δ T   and is a solution of the approximated problem on [ 0 , T ˇ + 2 δ T ]   . We then have in the same fashion as we got ( 9.1 ), t [ 0 , T ~ + 2 δ T ] , ( w ~ , q ~ ) Z t 2 N 1 + C 1 t 1 4 ( w ~ , q ~ ) Z t 6 .   By induction, we then see that we get a solution ( w ~ , q ~ )   defined on [ 0 , t 1 ]   ,satisfying the estimate
t [ 0 , t 1 ] , ( w ~ , q ~ ) Z t 3 N 1 = 12 C δ 0 [ N 0 ( u 0 , ( w i ) i = 1 3 ) + M 0 ( f ) + N 0 ( ( q i ) i = 0 2 ) ] , (9.4)
establishing the independence of the time of existence respectively to κ   , since t 1   does not depend on κ   . In the following we will note T = t 1   .

10 Existence for ( 3.2 )

  • Proof. We can here choose to take κ = 1 n   , and let n   . By the bound ( 9.4 ) independent of κ   on [ 0 , T ]   , we then have the existence of a weakly convergent subsequence of ( w ~ , q ~ )   in the reflexive Hilbert space Y T   , to a limit that we call ( v , q )   , which also belongs to Z T   and satisfies the estimate ( v , q ) Z T 3 N 1 = 12 C δ 0 [ N 0 ( u 0 , ( w i ) i = 1 3 ) + M 0 ( f ) + N ( ( q i ) i = 0 2 ) ] .   The usual compactness theorems ensure at this stage that ( v , q )   is a solution of ( 3.2 ) on [ 0 , T ]   . The smoothness of our solution ensures that the solids do not collide with each other (if there is more than one) or the boundary (for an eventually smaller time), which establishes the existence part of Theorem.  5.1 

11 Uniqueness for ( 3.2 )

  • Proof. Since we cannot use a contractive mapping scheme for our problem, we have to establish uniqueness separately. Let then ( v ¯ , q ¯ )   denote another solution of ( 3.2 ) in Z T   . Then, taking v v ¯   as a test function in the variational formulation of the difference between the systems ( 3.2 ) associated to each solution provides for t [ 0 , T ]   :
    1 2 ( v v ¯ ) ( t ) L 2 ( Ω ; R 3 ) 2 + ν 0 t ( a k r a k s v , r a ¯ k r a ¯ k s v ¯ , r , v , s v ¯ , s ) L 2 ( Ω 0 f ; R 3 )
    + 0 t ( c i j k l [ ( η , i η , j δ i j ) η , k ( η ¯ , i η ¯ , j δ i j ) η ¯ , k ] , v , l v ¯ , l ) L 2 ( Ω s 0 ; R 3 )
    0 t ( a i j q a ¯ i j q ¯ , v i , j v ¯ i , j ) L 2 ( Ω 0 f ; R ) = 0 t ( f η f η ¯ , v v ¯ ) L 2 ( Ω 0 f ; R 3 ) . (11.1)
    For the viscous term in the fluid, we write a k r a k s v , r a ¯ k r a ¯ k s v ¯ , r = a k r a k s ( v , r v ¯ , r ) + ( a k r a k s a ¯ k r a ¯ k s ) v ¯ , r ,   which with the L ( 0 , T ; H 3 ( Ω f 0 ; R 3 ) )   control of v ¯   and v   provides us with an estimate of the type (where C   denotes once again a generic constant)
    0 t ( a k r a k s v , r a ¯ k r a ¯ k s v ¯ , r , v , s v ¯ , s ) L 2 ( Ω 0 f ; R 3 ) C 0 t v v ¯ H 1 ( Ω 0 f ; R 3 ) 2
    C 0 t 0 t v v ¯ H 1 ( Ω 0 f ; R 3 ) 2 . (11.2)
    Concerning the forcing term in the fluid, we first notice that if we still denote E ( Ω ) ( f )   as f   , f ( t , η ¯ ( t , x ) ) f ( t , η ( t , x ) ) = 0 1 f , i ( t , ( η + t ( η ¯ η ) ) ( t , x ) ) d t ( η ¯ i ( t , x ) η i ( t , x ) ) ,   leading us to
    f ( t , η ¯ ( t , ) ) f ( t , η ( t , ) ) L 1.5 ( Ω f 0 ; R 3 )
    C η ¯ ( t , ) η ( t , ) L 6 ( Ω f 0 ; R 3 ) [ i = 1 3 0 1 Ω f 0 f , i 2 ( t , φ ( t , t , x ) ) d x d t ] 0.5 ,
    with φ ( t , t , x ) = η ( t , x ) + t ( η ¯ ( t , x ) η ( t , x ) )   . We have φ ( t , t , ) C 0 ( Ω ¯ ; R 3 ) C 1 ( Ω Γ 0 c ; R 3 )   . Moreover φ ( t , t , Ω ) = Ω   . We then have by invariance by homotopy of the Brouwer degree (for the parameter t   ) z Ω , deg ( φ ( t , t , ) , Ω , z ) = deg ( φ ( t , 0 , ) , Ω , z ) = deg ( Id , Ω , z ) = 1 ,   which together with the regularity of φ ( t , t , )   establishes that φ ( t , t , ) ( Ω ) = Ω   and that Card { φ 1 ( t , t , ) ( x ) } = 1   for almost all x Ω   . Thus, Ω f 0 f , i 2 ( t , φ ( t , t , x ) ) d x = φ ( t , t , Ω f 0 ) f , i 2 ( t , y ) | det φ ( t , t , φ 1 ( t , t , y ) ) | 1 d y ,   which with the L ( 0 , T ; H 4 ( Ω f 0 ; R 3 ) )   control of η   and η ¯   provides Ω f 0 f , i 2 ( t , φ ( t , t , x ) ) d x C Ω f , i 2 ( t , y ) d y .   Consequently,
    f ( t , η ¯ ( t , ) ) f ( t , η ( t , ) ) L 1.5 ( Ω f 0 ; R 3 ) C η ¯ ( t , ) η ( t , ) H 1 ( Ω f 0 ; R 3 ) f H 1 ( Ω ; R 3 ) ,
    implying
    | 0 t ( f η f η ¯ , v v ¯ ) L 2 ( Ω 0 f ; R 3 ) | C t f L 2 ( 0 , t ; H 1 ( Ω ; R 3 ) ) v v ¯ L 2 ( 0 , t ; H 1 ( Ω f 0 ; R 3 ) ) 2 . (11.3)
    Concerning the elastic term,
    0 t ( c i j k l [ ( η , i η , j δ i j ) η , k ( η ¯ , i η ¯ , j δ i j ) η ¯ , k ] , v , l v ¯ , l ) L 2 ( Ω s 0 ; R 3 ) = I 1 + I 2 + I 3 ,
    with
    I 1 = 0 t ( c i j k l ( η , i η , j δ i j ) ( η , k η ¯ , k ) , v , l v ¯ , l ) L 2 ( Ω s 0 ; R 3 )
    = 1 2 ( c i j k l ( η , i η , j δ i j ) ( η , k η ¯ , k ) , η , l η ¯ , l ) L 2 ( Ω s 0 ; R 3 ) ( t )
    1 2 0 t ( c i j k l ( η , i η , j δ i j ) t ( η , k η ¯ , k ) , η , l η ¯ , l ) L 2 ( Ω s 0 ; R 3 )
    C t η ( t ) η ¯ ( t ) H 1 ( Ω s 0 ; R 3 ) 2 C 0 t η η ¯ H 1 ( Ω s 0 ; R 3 ) 2 ,
    where we have used the L ( 0 , T ; H 3 ( Ω s 0 ; R 3 ) )   control of v   and v ¯   for the inequality.
    Next, for the same reasons,
    I 2 = 0 t ( c i j k l ( η , i η ¯ , i ) η , j η ¯ , k , v , l v ¯ , l ) L 2 ( Ω s 0 ; R 3 )
    = 0 t ( c i j k l ( η , i η ¯ , i ) ( η , j η ¯ , j ) η ¯ , k , v , l v ¯ , l ) L 2 ( Ω s 0 ; R 3 )
    + 0 t ( c i j k l ( η , i η ¯ , i ) η ¯ , j η ¯ , k , v , l v ¯ , l ) L 2 ( Ω s 0 ; R 3 )
    = 0 t ( c i j k l ( η , i η ¯ , i ) ( η , j η ¯ , j ) η ¯ , k , v , l v ¯ , l ) L 2 ( Ω s 0 ; R 3 )
    + 1 2 ( c i j k l ( η , i η ¯ , i ) η ¯ , j η ¯ , k , η , l η ¯ , l ) L 2 ( Ω s 0 ; R 3 ) ( t )
    0 t ( c i j k l ( η , i η ¯ , i ) η ¯ , j v ¯ , k , η , l η ¯ , l ) L 2 ( Ω s 0 ; R 3 ) .
    We then write for the second term on the right-hand side of the last equality η ¯ , i ( t , ) = e i + 0 t v ¯ , i   , to get by Korn's inequality
    I 2 C [ η ( t ) η ¯ ( t ) H 1 ( Ω s 0 ; R 3 ) 2 η ( t ) η ¯ ( t ) L 2 ( Ω s 0 ; R 3 ) 2 ] C t sup [ 0 , t ] η η ¯ H 1 ( Ω s 0 ; R 3 ) 2 .
    Similarly,
    I 3 = 0 t ( c i j k l ( η , j η ¯ , j ) η ¯ , i η ¯ , k , v , l v ¯ , l ) L 2 ( Ω s 0 ; R 3 )
    C [ η ( t ) η ¯ ( t ) H 1 ( Ω s 0 ; R 3 ) 2 η ( t ) η ¯ ( t ) L 2 ( Ω s 0 ; R 3 ) 2 ] C 0 t η η ¯ H 1 ( Ω s 0 ; R 3 ) 2 .
    Thus,
    0 t ( c i j k l [ ( η , i η , j δ i j ) η , k ( η ¯ , i η ¯ , j δ i j ) η ¯ , k ] , v , l v ¯ , l ) L 2 ( Ω s 0 ; R 3 )
    C [ η ( t ) η ¯ ( t ) H 1 ( Ω s 0 ; R 3 ) 2 η ( t ) η ¯ ( t ) L 2 ( Ω s 0 ; R 3 ) 2 ] C 0 t η η ¯ H 1 ( Ω s 0 ; R 3 ) 2 . (11.4)
    Concerning the pressure term, with a i j q a ¯ i j q ¯ = ( a i j a ¯ i j ) q + a ¯ i j ( q q ¯ )   and the L ( 0 , T ; H 2 ( Ω f 0 ; R ) )   control of the pressure, we get
    0 t ( a i j q a ¯ i j q ¯ , v i , j v ¯ i , j ) L 2 ( Ω 0 f ; R )
    C [ t q q ¯ L ( 0 , t ; L 2 ( Ω f 0 ; R ) ) v v ¯ L 2 ( 0 , t ; H 1 ( Ω f 0 ; R 3 ) ) + t v v ¯ L 2 ( 0 , t ; H 1 ( Ω f 0 ; R 3 ) ) 2 ] . (11.5)
    In order to get the estimate of q q ¯   in L 2 ( Ω f 0 ; R )   , we have to introduce the time differentiated problem. By taking v t v ¯ t   in the variational formulation associated to the difference between the time differentiated systems, we obtain
    1 2 ( v t v ¯ t ) ( t ) L 2 ( Ω ; R 3 ) 2 + ν 0 t ( [ a k r a k s v , r a ¯ k r a ¯ k s v ¯ , r ] t , [ v , s v ¯ , s ] t ) L 2 ( Ω 0 f ; R 3 )
    + 0 t ( c i j k l [ ( η , i η , j δ i j ) η , k ( η ¯ , i η ¯ , j δ i j ) η ¯ , k ] t , [ v , l v ¯ , l ] t ) L 2 ( Ω s 0 ; R 3 )
    0 t ( [ a i j q a ¯ i j q ¯ ] t , [ v i , j v ¯ i , j ] t ) L 2 ( Ω 0 f ; R ) = 0 t ( [ f η f η ¯ ] t , v t v ¯ t ) L 2 ( Ω 0 f ; R 3 ) . (11.6)
    For the fluid viscous term, we easily find with the L 2 ( 0 , T ; H 3 ( Ω f 0 ; R 3 ) )   control of the first time derivative of the velocity that
    0 t ( [ a k r a k s v , r a ¯ k r a ¯ k s v ¯ , r ] t , v t , s v ¯ t , s ) L 2 ( Ω 0 f ; R 3 ) C ( 1 t ) 0 t v t v ¯ t H 1 ( Ω 0 f ; R 3 ) 2 . (11.7)
    Concerning the forcing term in the fluid, since ( f η ) t = ( f t + v i f , i ) ( η )   (with a similar formula for v ¯   ), we then deduce in a way similar to the steps leading to ( 11.3 ) that
    | 0 t ( [ f η f η ¯ ] t , [ v v ¯ ] t ) L 2 ( Ω 0 f ; R 3 ) |
    C t [ f t L 2 ( 0 , t ; H 1 ( Ω ; R 3 ) ) + f L 2 ( 0 , t ; H 2 ( Ω ; R 3 ) ) ] v t v ¯ t L 2 ( 0 , t ; H 1 ( Ω f 0 ; R 3 ) ) 2 . (11.8)
    For the elastic term, we can also essentially reproduce the arguments leading to ( 11.4 ), leading us to
    0 t ( c i j k l [ ( η , i η , j δ i j ) η , k ( η ¯ , i η ¯ , j δ i j ) η ¯ , k ] t , [ v , l v ¯ , l ] t ) L 2 ( Ω s 0 ; R 3 )
    C [ v ( t ) v ¯ ( t ) H 1 ( Ω s 0 ; R 3 ) 2 v ( t ) v ¯ ( t ) L 2 ( Ω s 0 ; R 3 ) 2 ] C t sup [ 0 , t ] v v ¯ H 1 ( Ω s 0 ; R 3 ) 2 . (11.9)
    The pressure term will require more care since we want to avoid the introduction of q t q ¯ t   that the most direct method would lead to. To do so, we notice that
    0 t ( [ a i j q a ¯ i j q ¯ ] t , [ v i , j v ¯ i , j ] t ) L 2 ( Ω 0 f ; R ) = I 4 + I 5 + I 6 ,
    with
    I 4 = 0 t ( [ ( a i j ) t q ( a ¯ i j ) t q ¯ ] , [ v i , j v ¯ i , j ] t ) L 2 ( Ω 0 f ; R ) ,
    I 5 = 0 t ( a i j ( q t q ¯ t ) , [ v i , j v ¯ i , j ] t ) L 2 ( Ω 0 f ; R ) ,
    I 6 = 0 t ( [ a i j a ¯ i j ] q ¯ t , [ v i , j v ¯ i , j ] t ) L 2 ( Ω 0 f ; R ) .
    For I 4   , we have in a way similar to ( 11.5 ),
    | I 4 | C [ t q q ¯ L ( 0 , t ; L 2 ( Ω f 0 ; R ) ) v t v ¯ t L 2 ( 0 , t ; H 1 ( Ω f 0 ; R 3 ) )
    + t v t v ¯ t L 2 ( 0 , t ; H 1 ( Ω f 0 ; R 3 ) ) 2 ] .
    For I 6   , the L 2 ( 0 , T ; H 2 ( Ω f 0 ; R ) )   control of q ¯ t   provides us with | I 6 | C t v t v ¯ t L 2 ( 0 , t ; H 1 ( Ω f 0 ; R 3 ) ) 2 .   For I 5   we have:
    I 5 = 0 t ( q t q ¯ t , a i j v t i , j a ¯ i j v ¯ t i , j ) L 2 ( Ω 0 f ; R ) 0 t ( q t q ¯ t , ( a i j a ¯ i j ) v ¯ t i , j ) L 2 ( Ω 0 f ; R )
    = 0 t ( q ¯ t q t , ( a i j ) t v i , j ( a ¯ i j ) t v ¯ i , j ) L 2 ( Ω 0 f ; R ) 0 t ( q t q ¯ t , ( a i j a ¯ i j ) v ¯ t i , j ) L 2 ( Ω 0 f ; R ) ,
    where we have used the relations a i j v i , j = 0 = a ¯ i j v ¯ i , j   in Ω f 0   for the first integral.
    By integrating by parts in time,
    I 5 = 0 t ( q q ¯ , [ ( a i j ) t v i , j ( a ¯ i j ) t v ¯ i , j ] t ) L 2 ( Ω 0 f ; R ) + 0 t ( q q ¯ , [ ( a i j a ¯ i j ) v ¯ t i , j ] t ) L 2 ( Ω 0 f ; R )
    + ( q ¯ q , ( a i j ) t v i , j ( a ¯ i j ) t v ¯ i , j ) L 2 ( Ω 0 f ; R ) ( t ) + ( q ¯ q , ( a i j a ¯ i j ) v ¯ t i , j ) L 2 ( Ω 0 f ; R ) ( t ) .
    With the L 2 ( 0 , T ; H 3 ( Ω f 0 ; R 3 ) )   control of v t   we have
    | 0 t ( q q ¯ , [ ( a i j ) t v i , j ( a ¯ i j ) t v ¯ i , j ] t ) L 2 ( Ω 0 f ; R ) | + | 0 t ( q q ¯ , [ ( a i j ) t ( a ¯ i j ) t ] v ¯ t i , j ) L 2 ( Ω 0 f ; R ) |
    C t q q ¯ L ( 0 , t ; L 2 ( Ω f 0 ; R ) ) v t v ¯ t L 2 ( 0 , t ; H 1 ( Ω f 0 ; R 3 ) ) ,
    | ( q q ¯ , ( a i j ) t v i , j ( a ¯ i j ) t v ¯ i , j ) L 2 ( Ω 0 f ; R ) ( t ) |
    C t q ( t ) q ¯ ( t ) L 2 ( Ω f 0 ; R ) v t v ¯ t L 2 ( 0 , t ; H 1 ( Ω f 0 ; R 3 ) ) .
    The remaining terms are more delicate. We first have
    | 0 t ( q q ¯ , ( a i j a ¯ i j ) v ¯ t t i , j ) L 2 ( Ω 0 f ; R ) | + | ( q q ¯ , ( a i j a ¯ i j ) v ¯ t i , j ) L 2 ( Ω 0 f ; R ) ( t ) |
    C 0 t q q ¯ L 2 ( Ω f 0 ; R ) a a ¯ L 4 ( Ω f 0 ; R 9 ) v ¯ t t L 4 ( Ω f 0 ; R 9 )
    + q ( t ) q ¯ ( t ) L 2 ( Ω f 0 ; R ) a ( t ) a ¯ ( t ) L 4 ( Ω f 0 ; R 9 ) v ¯ t ( t ) L 4 ( Ω f 0 ; R 9 ) . (11.10)
    The apparent problem here is that a a ¯   is estimated in L 2 ( Ω f 0 ; R 9 )   in terms of v v ¯   in H 1 ( Ω f 0 ; R 3 )   . Now, a bound of this quantity in L 4 ( Ω f 0 ; R 9 )   will require a bound of v v ¯   in H 2 ( Ω f 0 ; R 3 )   . In order to get such an estimate, we will bound v v ¯   in H 2 ( Ω f 0 ; R 3 )   by lower order terms in v v ¯   . To do so, let us first estimate the trace of v v ¯   on Γ 0   by using the test function [ ζ 2 ( v v ¯ ) Ψ ] , α α Ψ 1   in the difference between the variational problems satisfied by v   and v ¯   . By proceeding as in Section  10 , we would then get an estimate of the type, where δ > 0   is given:
    0 t ζ [ ( v v ¯ ) Ψ ] , α L 2 ( R + 3 ; R 9 ) 2 + ζ [ ( η η ¯ ) Ψ ] , α ( t ) L 2 ( R 3 ; R 9 ) 2
    C [ t + δ ] 0 t [ v v ¯ H 2 ( Ω f 0 ; R 3 ) 2 + q q ¯ H 1 ( Ω f 0 ; R ) 2 ] + C δ 0 t v t v ¯ t L 2 ( Ω ; R 3 ) 2
    + C 0 t η η ¯ H 2 ( Ω s 0 ; R 3 ) 2 + C 0 t v v ¯ H 1 ( Ω ; R 3 ) 2
    which by patching all the charts defining Γ 0   leads to an estimate of v v ¯   in L 2 ( 0 , t ; H 1.5 ( Γ 0 ; R 3 ) )   yielding by elliptic regularity:
    0 t [ v v ¯ H 2 ( Ω f 0 ; R 3 ) 2 + q q ¯ H 1 ( Ω f 0 ; R ) 2 ] + η ( t ) η ¯ ( t ) H 2 ( Ω s 0 ; R 3 ) 2
    C [ t + δ ] 0 t [ v v ¯ H 2 ( Ω f 0 ; R 3 ) 2 + q q ¯ H 1 ( Ω f 0 ; R ) 2 ] + C δ 0 t v t v ¯ t L 2 ( Ω ; R 3 ) 2
    + C 0 t η η ¯ H 2 ( Ω s 0 ; R 3 ) 2 + C 0 t v v ¯ H 1 ( Ω ; R 3 ) 2 .
    Thus, with a choice of δ > 0   small enough, we have for t   small enough and the use of Gronwall's inequality,
    η ( t ) η ¯ ( t ) H 2 ( Ω s 0 ; R 3 ) 2 + 0 t [ v v ¯ H 2 ( Ω f 0 ; R 3 ) 2 + q q ¯ H 1 ( Ω f 0 ; R ) 2 ]
    C 0 t v t v ¯ t L 2 ( Ω ; R 3 ) 2 + C 0 t v v ¯ H 1 ( Ω ; R 3 ) 2 .
    By using this estimate in ( 11.10 ), we then get for a time small enough
    | 0 t ( q q ¯ , ( a i j a ¯ i j ) v ¯ t t i , j ) L 2 ( Ω 0 f ; R ) | + | ( q q ¯ , ( a i j a ¯ i j ) v ¯ t i , j ) L 2 ( Ω 0 f ; R ) ( t ) |
    C t [ 0 t v t v ¯ t L 2 ( Ω ; R 3 ) 2 + 0 t v v ¯ H 1 ( Ω ; R 3 ) 2 + q q ¯ L ( 0 , t ; L 2 ( Ω f 0 ; R ) ) 2 ] .
    By putting together the estimates on I 4   , I 5   and I 6   , we have
    | 0 t ( [ a i j q a ¯ i j q ¯ ] t , [ v i , j v ¯ i , j ] t ) L 2 ( Ω 0 f ; R ) |
    C t [ 0 t v t v ¯ t L 2 ( Ω ; R 3 ) 2 + 0 t v v ¯ H 1 ( Ω ; R 3 ) 2 + q q ¯ L ( 0 , t ; L 2 ( Ω f 0 ; R ) ) 2 ] . (11.11)
    Now, by considering the difference between the two variational forms satisfied repectively by ( v , q )   and ( v ¯ , q ¯ )   , and writing the difference between the pressure terms as
    Ω f 0 ( a i j q a ¯ i j q ¯ ) φ i , j = Ω f 0 a i j ( q q ¯ ) φ i , j + Ω f 0 ( a i j a ¯ i j ) q ¯ φ i , j ,
    the Lagrange multiplier Lemma 13 of [5provides for all t [ 0 , T ]   :
    q ( t ) q ¯ ( t ) L 2 ( Ω f 0 ; R ) C [ ( v t v ¯ t ) ( t ) L 2 ( Ω ; R 3 ) + ( v v ¯ ) ( t ) H 1 ( Ω 0 f ; R 3 )
    + ( η η ¯ ) ( t ) H 1 ( Ω s 0 ; R 3 ) + t v v ¯ L 2 ( 0 , t ; H 1 ( Ω f 0 ; R 3 ) ) ] . (11.12)
    By putting together the estimates ( 11.1 )-( 11.12 ), we then obtain for t u > 0   small enough an inequality of the type:
    v t v ¯ t L ( 0 , t u ; L 2 ( Ω ; R 3 ) ) 2 + 0 t u v t v ¯ t H 1 ( Ω f 0 ; R 3 ) 2 + v v ¯ L ( 0 , t u ; H 1 ( Ω 0 s ; R 3 ) ) 2 0 ,
    which shows that ( v , q ) = ( v ¯ , q ¯ )   on [ 0 , t u ]   . Let T u = sup { t [ 0 , T ] | ( v , q ) = ( v ¯ , q ¯ ) on [ 0 , t ] } .   If T u < T   , we can repeat the same procedure with T u   replacing 0   , which would lead to uniqueness for [ T u , T u + δ t )   as well. Thus, we have T u = T   , which concludes the proof of the theorem.

12 Optimal regularity on the initial data

We first remind some extensions and regularization results on domains:
Lemma 12.1. Let Ω   be a domain of class H 4   . Then, there exists a linear and continuous operator E ( Ω )   from H m ( Ω ; R 3 )   into H m ( R 3 ; R 3 )   (for each 0 m 4   ) such that E ( Ω ) ( u ) = u   in Ω   . Also, if the H 4   norms of a family of domains stay bounded, the norms of the corresponding linear operators also stay bounded.
Lemma 12.2. Since Ω 0 s   is of class H 4   , let ψ m H 4 ( B ( 0 , 1 ) ; R 3 )   ( m = 1 , . . . , N   ) be a collection of charts defining a neighborhood of its boundary. We note Ω f 0 H 4 = m = 1 N ψ m H 4 ( B ( 0 , 1 ) ; R 3 ) .   Then, there exists a sequence of domains ( Ω 0 s , n   ) of class C   , so that Ω s 0 Ω 0 s , n   , and which are defined with a collection of charts ψ m , n H 4 ( B ( 0 , 1 ) ; R 3 )   ( m = 1 , . . . , N   ) so that m = 1 N ψ m ψ m , n H 4 ( B ( 0 , 1 ) ; R 3 ) 0   as n   . We then denote the complementary of Ω ¯ 0 s , n   in Ω   by Ω 0 f , n   and Γ 0 n = Ω 0 s , n   . We also assume n   large enough so that the different connected components of Ω 0 s , n   (if there is more than one solid) do not intersect each other or the boundary of Ω   . We denote α n = Ω s , n H 6   .
We now state the optimal regularity assumptions needed in our analysis, and explain the adjustements required to the previous proofs.
Theorem 12.1. With the same assumptions as in Theorem  5.1 , except for the following concerning the regularity of the initial data:
u 0 H 6 ( Ω 0 f ; R 3 ) H 3 ( Ω 0 s ; R 3 ) H 0 1 ( Ω ; R 3 ) L d i v , f 2 , (12.1a)
f s ( 0 ) H 2 ( Ω 0 s ; R 3 ) H 3.5 ( Γ 0 ; R 3 ) , ( f s ) t ( 0 ) H 1 ( Ω 0 s ; R 3 ) , ( f s ) t t ( 0 ) L 2 ( Ω ; R 3 ) , (12.1b)
the conclusion of Theorem  5.1 holds.
Remark 8. We have chosen here to take different forcings for the fluid, that we still note f   with the same assumptions as in Theorem  5.1 , and the solid, in order to stress out that the higher order regularity required indeed comes from the hyperbolic scaling of the Navier-Stokes equations. The somewhat not so natural condition f s ( 0 ) H 3.5 ( Γ 0 ; R 3 )   is made in order to get w 1 H 4 ( Ω f 0 ; R 3 )   associated to the condition w 3 L 2 ( Ω f 0 ; R 3 )   .
  • Proof. The idea is to first regularize the domains, initial data and modify the forcings in an appropriate way, and then pass to the limit.
    Given 0 ρ D ( B ( 0 , 1 ) )   with B ( 0 , 1 ) ρ = 1   , we define as usual ρ n ( x ) = n 3 ρ ( n x )   .
    We first notice that u 0   , w 1   , q 0   and q 1   still have the same regularity in Ω f 0   as in Theorem  5.1 . We first define in Ω 0 f , n   , u 0 n = u 0   and w 1 n = w 1   , q 0 n = q 0   , q 1 n = q 1   (which is permitted since Ω 0 f , n Ω 0 f   . We next define w 2 n   in Ω 0 f , n   ,
    ν w 2 n + q 2 n = ρ n E ( Ω 0 f ) ( ν w 2 + q 2 ) in Ω 0 f , n , (12.2a)
    div w 2 n = [ ( a i j ) t t ( 0 ) u 0 i , j + 2 ( a i j ) t ( 0 ) w 1 i , j ] in Ω 0 f , n , (12.2b)
    w 2 n = 0 on Ω , (12.2c)
    ν w 2 n N n q 2 n N n = ν N n ρ n E ( Ω 0 f ) ( w 2 ) ρ n E ( Ω 0 f ) ( q 2 ) N n on Γ 0 n , (12.2d)
    where N n   denotes the unit normal exterior to Ω 0 f , n   , and finally w 3 n L 2 ( Ω 0 f , n ; R 3 )   by w 3 n = [ ν ( a l j a l k u , k ) , j ( a i j q , j ) i = 1 3 + F ] t t ( 0 ) in Ω 0 f , n ,   where the time derivatives on the right-hand side are computed with the usual rules from u ( 0 ) = u 0 n   , t p u ( 0 ) = w p n   ( p = 1 , 2 )   , t p q ( 0 ) = q p n   ( p = 0 , 1 , 2 )   .
    We next define in the solid u 0 n   by
    L 2 u 0 n = L 2 [ ρ n E ( Ω s 0 ) ( u 0 ) ] in Ω 0 s , n , (12.3a)
    u 0 n = ( u 0 n ) f on Γ 0 n , (12.3b)
    L ( u 0 n ) + ρ n E ( Ω s 0 ) ( ( f s ) t ( 0 ) ) = ( w 2 n ) f on Γ 0 n , (12.3c)
    where the right-hand sides of the previous boundary conditions come from the fluid regularization previously carried. Note also that
    L u 0 n H 4 ( Ω 0 s , n ; R 3 ) , (12.4)
    (with an estimate that may blow up as n   ) since
    L ( L u 0 n ) = L 2 [ ρ n E ( Ω s 0 ) ( u 0 ) ] in Ω 0 s , n ,
    L ( u 0 n ) = ρ n E ( Ω s 0 ) ( ( f s ) t ( 0 ) ) + ( w 2 n ) f on Γ 0 n .
    We can then define f 0 n   in Ω 0 s , n   by
    L 2 f 0 n = L 2 [ ρ n E ( Ω s 0 ) ( f s ( 0 ) ) ] in Ω 0 s , n ,
    f 0 n = ( w 1 n ) f on Γ 0 n ,
    c m j k i ( f 0 n , m Id , j + f 0 n , j Id , m ) N k n = 2 c m j k i ( u 0 n , m Id , j + u 0 n , j Id , m ) u 0 n , k i
    + ν [ v i , k a l k a l j ] t t ( 0 ) N j n [ q a i j ] t t ( 0 ) N j n on Γ 0 n ,
    with the same conventions as for the previous system for the time derivatives evaluated from Ω 0 f , n   and c m j k l ( u 0 n , m Id , j + u 0 n , j Id , m ) u 0 n , k i N l n   evaluated from Ω 0 s , n   .
    We then define in Ω 0 s , n   ,
    w 1 n = f 0 n ,
    w 2 n = [ [ c m j k l ( η , m η , j δ m j ) η , k ] , l ] t ( 0 ) + ρ n E ( Ω s 0 ) ( ( f s ) t ( 0 ) ) ,
    = L ( u 0 n ) + ρ n E ( Ω s 0 ) ( ( f s ) t ( 0 ) ) ,
    w 3 n = [ [ c m j k l ( η , m η , j δ m j ) η , k ] , l ] t t ( 0 ) + ρ n E ( Ω s 0 ) ( ( f s ) t t ( 0 ) ) ,
    where the time derivatives on the right-hand side are evaluated with v ( 0 ) = u 0 n   , v t ( 0 ) = w 1 n   . We also define the regularized forcing in the solid
    f n ( t ) = ρ n E ( Ω s 0 ) ( f s ( t ) f s ( 0 ) ) + f 0 n in Ω 0 s , n .
    We then have u 0 n   , w 1 n   , w 2 n   in H 0 1 ( Ω ; R 3 ) H 4 ( Ω 0 f , n ; R 3 ) H 4 ( Ω 0 s , n ; R 3 )   and div u 0 n = 0   in Ω 0 f , n   , w 3 n L 2 ( Ω ; R 3 )   , with
    E ( Ω 0 f , n ) ( u 0 n ) u 0 H 4 ( Ω f 0 ; R 3 ) + E ( Ω 0 s , n ) ( u 0 n ) u 0 H 3 ( Ω s 0 ; R 3 ) 0 as n , (12.5a)
    E ( Ω 0 f , n ) ( w 1 n ) w 1 H 4 ( Ω f 0 ; R 3 ) + E ( Ω 0 s , n ) ( w 1 n ) w 1 H 2 ( Ω s 0 ; R 3 )
    + w 2 n w 2 H 1 ( Ω ; R 3 ) + E ( Ω 0 f , n ) ( w 2 n ) w 2 H 2 ( Ω 0 f ; R 3 ) 0 as n , (12.5b)
    u 0 n H 6 ( Ω 0 s , n ; R 3 ) β n , w 1 n H 4 ( Ω 0 s , n ; R 3 ) β n , w 2 n H 4 ( Ω 0 s , n ; R 3 ) β n , (12.5c)
    w 3 n w 3 L 2 ( Ω ; R 3 ) as n , (12.5d)
    where β n   is a given polynomial expression of α n   and n   . We briefly explain how those constants appear. For instance, for the first estimate of ( 12.5c ), we have by elliptic regularity on ( 12.3 ) that u 0 n H 6 ( Ω 0 s , n ; R 3 )   is bounded by a sum of terms, one of which being P ( Ω 0 s , n H 6 ) ( w 2 n ) f H 4 ( Ω 0 f , n ; R 3 )   , P   being a polynomial which does not depend on n   . Next, still by elliptic regularity on ( 12.2 ), we have that ( w 2 n ) f H 4 ( Ω 0 f , n ; R 3 )   is bounded by a sum of terms such as ρ n E ( Ω f 0 ) ( w 2 ) H 2 ( Ω 0 f , n ; R 3 )   .
    This particular term, by the properties of the convolution, is in turn bounded by n 3 E ( Ω 0 f , n ) ( w 2 ) H 1 ( R 3 ; R 3 )   . This shows that a term of the type P ( α n ) n 3 w 2 H 1 ( Ω f 0 ; R 3 )   appears in the sum of all terms bounding u 0 n H 6 ( Ω 0 s , n ; R 3 )   . Since the other terms in the sum can be dealt with similarly, this explains our estimate ( 12.5c ).
    For the pressures, we have
    E ( Ω 0 f , n ) ( q 0 n ) q 0 H 3 ( Ω f 0 ; R ) + E ( Ω 0 f , n ) ( q 1 n ) q 1 H 3 ( Ω f 0 ; R )
    + E ( Ω 0 f , n ) ( q 2 n ) q 2 H 1 ( Ω f 0 ; R ) 0 as n . (12.6)
    Since the initial data u 0 n   and forcings f n ( 0 )   , f t n ( 0 )   , f t t n ( 0 )   are smooth enough to ensure the regularity properties ( 12.5 ), we then deduce that we have similarly as for theorem ( 5.1 ) the existence of a solution w n   of a system similar to ( 7.7 ) with f   , u 0   , Ω f 0   , Ω s 0   being replaced by their counterparts with an exponent n   , and b κ   , c κ   , d κ   being replaced by b n   , c n   , d n   (with the choice κ = 1 n ( β n + 1 )   ) given by
    b n ( φ ) = 1 n ( β n + 1 ) ( c i j k l w 2 n , l k , φ i , j ) L 2 ( Ω 0 s , n ; R ) ,
    c n ( φ ) = 1 n ( β n + 1 ) ( c i j k l w 1 n , l k , φ i , j ) L 2 ( Ω 0 s , n ; R ) + ( w 2 n w 2 , φ ) L 2 ( Ω 0 f , n ; R )
    + ( [ ( a i j q ) t ( 0 ) N j n ] i = 1 3 + [ ( a l j a l k ) u , k ] t ( 0 ) N j n , φ ) L 2 ( Γ 0 n ; R 3 )
    ( c i j k l [ ( η , i η , j δ i j ) η k ] t ( 0 ) N l n , φ ) L 2 ( Γ 0 n ; R 3 ) ,
    d n ( φ ) = 1 n ( β n + 1 ) ( c i j k l u 0 n , l k , φ i , j ) L 2 ( Ω 0 s , n ; R )
    + ( [ ( a i j q ) ( 0 ) N j n ] i = 1 3 + [ ( a l j a l k ) u , k ] ( 0 ) N j n , φ ) L 2 ( Γ 0 n ; R 3 ) ,
    where the time derivatives are computed with a velocity satisfying u ( 0 ) = u 0 n   , u t ( 0 ) = w 1 n   and a pressure such that q ( 0 ) = q 0 n   , q t ( 0 ) = q 1 n   . Note that by construction, the solutions w n   to these problems in Ω   satisfy w n ( 0 ) = u 0 n   , w n t ( 0 ) = w 1 n   , w n t t ( 0 ) = w 2 n   , w n t t t ( 0 ) = w 3 n   . Next, we proceed to energy estimates similar to Section  8 . The bounds obtained are similar, except that this time the terms associated to b n   , c n   and d n   tend to zero as n   . This is clear from the convergence results ( 12.5 ), ( 12.6 ) for the integral terms associated to the fluid.
    The terms associated to the solid asymptotically tend to zero by properties of the convolution. For instance, with the notations of Section  8 , for φ p = [ ρ p ( ζ 2 w n Ψ ) ] , α 1 α 1 α 2 α 2 α 3 α 3 Ψ 1   , we get after change of variables, an integration by parts in time, three integrations by parts in space:
    | 0 t t 2 2 n ( β n + 1 ) ( c i j k l w 2 n , l k , φ p i , j ) L 2 ( Ω 0 s , n ; R ) d t |
    C n ( β n + 1 ) w 2 n H 4 ( Ω 0 s , n ; R 3 ) η n L ( 0 , t ; H 4 ( Ω 0 s , n ; R 3 ) ) ,
    and thus with our estimate ( 12.5c ), we have
    | 0 t t 2 2 n ( β n + 1 ) ( c i j k l w 2 n , l k , φ p i , j ) L 2 ( Ω 0 s , n ; R ) d t | C n ( w n , q n ) Z t n ,
    where Z t n   denotes the same type of space as Z t   with Ω 0 s   and Ω 0 f   being replaced by their counterparts with an exponent n   . This type of estimates thus shows that this term does not change the energy inequalities in Section  8 . We can thus reproduce the arguments of Section  9 , establishing that ( w n , q n )   can be defined over a time T   independently of n   , and that its norm in Z T n   depends solely on N ( u 0 n , ( w i n ) i = 1 3 ) + N ( ( q i ) i = 0 2 ) + M ( f n , 0 , 0 )   and thus, thanks to the estimates ( 12.5a ), ( 12.5b ), ( 12.5d ), solely on N ( u 0 , ( w i ) i = 1 3 ) + N ( ( q i ) i = 0 2 ) + M ( f , 0 , 0 )   . We can then consider the sequence ( E ( Ω ) ( w n ) , E ( Ω 0 f , n ) ( q n ) )   which is bounded in a space similar as Z T   , but defined on R 3   and extract (with respect to n   ) a weakly convergent sequence in a space modified from Y T   by replacing the condition u H 0 1 ( Ω ; R 3 )   by u H 1 ( Ω ; R 3 )   . By the classical compactness results, we next see that the weak limit ( v , q ) Z T   and is a solution of ( 3.2 ) with f   as forcing and v ( 0 ) = u 0   . This solution is also unique in Z T   .

13 The case of incompressible elasticity

In this Section, we explain how to treat the supplementary difficulties appearing when the incompressibility constraint is added in the solid. This leads to the same system as ( 3.2 ), with the addition of the condition det η = 1   a.e. in Ω s 0   and the addition of [ ( a i k q ) , k ] i = 1 3   on the left-hand side of ( 3.2d ) and the addition of q a i j N j   (the trace of q   being from the solid phase in this new term) on the left-hand side of ( 3.2e ). We now state our result and explain how to overcome the additional difficulties related to this constraint.
We first update our functional frameworks. While X T   and W T   do not change, Y T   and Z T   become respectively
Y T = { ( v , q ) X T × L 2 ( 0 , T ; L 2 ( Ω ; R ) ) | t n q L 2 ( 0 , T ; H 3 n ( Ω 0 f ; R ) ) ,
t n q L 2 ( 0 , T ; H 3 n ( Ω 0 s ; R ) ) ( n = 0 , 1 , 2 ) } ,
Z T = { ( v , q ) W T × L 2 ( 0 , T ; L 2 ( Ω ; R ) ) | t n q L 2 ( 0 , T ; H 3 n ( Ω 0 f ; R ) ) ,
t n q L 2 ( 0 , T ; H 3 n ( Ω 0 s ; R ) ) ( n = 0 , 1 , 2 ) | q t t L ( 0 , T ; L 2 ( Ω ; R ) ) } .
Remark 9. Whereas the pressure in the solid satisfies t n q L ( 0 , T ; H 3 n ( Ω 0 s ; R ) )   ( n = 0 , 1 , 2 )   , it appears that the limit pressures q κ   are controlled uniformly in the norm of Z T   and seemingly not in these norms. Note also that whereas the velocity field is smoother in the fluid phase for the solution of our next theorem, the pressure field is actually smoother in the solid phase. Whereas our artificial viscosity smoothes the velocity field in the solid, it also interestingly makes the pressure in the solid for the regularized system less smooth than the one associated to the solution of the constrained problem, which is source of difficulties that we shall describe later.
We now state our result:
Theorem 13.1. With the same regularity assumptions as in Theorem  12.1 and assuming that the compatibility conditions associated to our new system at t = 0   hold (for the sake of conciseness we do not state them here), the conclusion of Theorem  5.1 holds for the case where the incompressibility constraint is added to the solid part.
Furthermore, t n q L ( 0 , T ; H 3 n ( Ω 0 s ; R ) )   ( n = 0 , 1 , 2 )   .
  • Proof. The extra regularity (with respect to the norm of Z T   ) on the pressure in the solid simply comes from the equation v t c m j k l [ ( η , m η , j δ i j ) η , k ] , l + a i j q , j = f in ( 0 , T ) × Ω 0 s ,   which once the regularity for the solution w W T   is known provides immediately the result. We now explain how to obtain a solution in Z T   .
    The beginning of the proof follows the same lines as for the compressible elasticity case. We first assume that the initial data satisfies the regularity assumptions of Theorem  5.1 , and define the same smoothed problem as ( 7.1 ) with the corresponding updates for the incompressibility constraint. We then define the same fixed point linear problem as ( 7.4 ) where the condition a i k w i , k = 0   in Ω s 0   is added (the a i k   being computed from the given v   ) and add a i k q , k   on the left-hand side of ( 7.4c ) and q a i j N j   (the traces being taken from Ω s 0   ) on the left-hand side of ( 7.4d ).
    We then proceed as in [5to construct a solution to this system by a penalty method (the penalty term being this time defined over Ω   ) and get the same type of regularity result. This provides us with a solution ( w κ , q κ )   , that we also denote by ( w ~ , q ~ )   , for the incompressible version of ( 7.1 ) on a time T κ   shrinking to zero.
    As for the compressible case, ( w κ , q κ )   is in Z T κ   , and since our smoothed problem has a parabolic artificial viscosity, we also have for the velocity in the solid the regularity t n w L 2 ( 0 , T κ ; H 4 n ( Ω s 0 ; R 3 ) )   ( n = 0 , 1 , 2 , 3   ) (with estimates that blow up as κ 0   ). Thus, ( w κ , q κ ) Z ~ T κ   with Z ~ t = { ( w , q ) Z t | t n w L 2 ( 0 , t ; H 4 n ( Ω s 0 ; R 3 ) ) ( n = 0 , 1 , 2 ) } ,   endowed with the norm ( w , q ) Z ~ t 2 = ( w , q ) Z t 2 + κ 2 n = 0 2 t n w L 2 ( 0 , t ; H 4 n ( Ω s 0 ; R 3 ) ) 2 .   We next proceed as in Section  8 to get energy estimates, that will be carried this time for the κ   dependent norm of Z ~ t   , independently of κ   on [ 0 , T κ ]   , and for such a purpose it is important to keep the κ 2   factor in the definition of the norm. We could extend the sum to n = 3   , whereas it is not necessary.
    As before, the first set of estimates has to be carried on the highest order time derivative. Our energy inequality ( 7.8 ) has the same form, except that the integrals over Ω 0 f   where q ~   appears has to be taken this time on Ω   . The part over Ω 0 f   is estimated as before. We now explain how to deal with the integrals set on Ω s 0   for the pressure, which indeed needs some justifications given that the velocity in the solid is not controlled uniformly in κ   in a space as smooth as the velocity in the fluid, while the pressure is controlled in the same type of spaces in both phases.
  • Estimates on w ~ t t t   Here t   denotes any time in ( 0 , T κ )   . The most difficult integrals set in [ 0 , t ] × Ω s 0   and associated to the incompressibility constraint in the solid are K 1 = 0 t Ω s 0 q ~ t t ( a ~ i j ) t w ~ t t t i , j   and K 2 = 0 t Ω s 0 q ~ ( a ~ i j ) t t t w ~ t t t i , j   , the others being either less difficult or similar to estimate.
    Step 1. For K 1   , if we denote N s = N   , we have
    | K 1 | = | 0 t Ω s 0 ( q ~ t t ) , j ( a ~ i j ) t w ~ t t t i + 0 t Γ 0 q ~ t t ( a ~ i j ) t w ~ t t t i N j s |
    C [ 0 t q ~ t t H 1 ( Ω s 0 ; R ) η ~ H 3 ( Ω s 0 ; R 3 ) w ~ H 3 ( Ω s 0 ; R 3 ) w ~ t t t L 2 ( Ω s 0 ; R 3 )
    + 0 t q ~ t t H 1 2 ( Ω s 0 ; R ) η ~ H 3 ( Ω s 0 ; R 3 ) w ~ H 3 ( Ω s 0 ; R 3 ) w ~ t t t H 1 2 ( Ω f 0 ; R 3 ) ]
    C t ( w ~ , q ~ ) Z t 4 + C t 1 2 sup [ 0 , t ] [ q ~ t t L 2 ( Ω s 0 ; R ) 1 2 w ~ t t t L 2 ( Ω f 0 ; R 3 ) 1 2 ] ( w ~ , q ~ ) Z t 3 (13.1)
    C t 1 4 ( w ~ , q ~ ) Z t 4 , (13.2)
    where we have used the continuity of w ~ t t t   in the sense of traces along Γ 0   to bound the L 2 ( Γ 0 ; R 3 )   norm of w ~ t t t   by means of the H 1 2 ( Ω f 0 ; R 3 )   norm. Note that we have also used the fact that the L ( L 2 )   norm of q ~ t t   is in the definition of the norm of Z t   . In order to get an estimate on this norm, we would proceed in a way similar as to get ( 8.8 ) in Section  9 .
    Step 2. Concerning K 2   , we have by integrating by parts in space
    K 2 = 0 t Ω s 0 q ~ , j ( a ~ i j ) t t t w ~ t t t i + 0 t Γ 0 q ~ ( a ~ i j ) t t t w ~ t t t i N j s ,
    since our artificial viscosity provides the regularity w ~ t t L 2 ( 0 , T κ ; H 2 ( Ω s 0 ; R 3 ) )   and w ~ t t t L 2 ( 0 , T κ ; H 1 ( Ω s 0 ; R 3 ) )   (with estimates that may blow up as κ 0   ).
    The difficulty here comes from the second integral. Whereas as for K 1   we can estimate the trace of w ~ t t t   on Γ 0   from the fluid, we have to take the norm of w ~ t t   in H 0.5 ( Γ 0 ; R 9 )   , which is problematic given that the norm Z t   contains only its L 2 ( Ω s 0 ; R 9 )   norm. In order to circumvent this difficulty, we notice that the same formula holds if we replace w ~ t t t   by E ( Ω f 0 ) ( w ~ t t t f )   (the extension to R 3   of the velocity in the fluid). Since w ~ t t t = w ~ t t t f   on Γ 0   , we have w ~ t t t = E ( Ω f 0 ) ( w ~ t t t f )   on Γ 0   , which implies:
    K 2 = 0 t Ω s 0 q ~ , j ( a ~ i j ) t t t w ~ t t t i + 0 t Ω s 0 q ~ ( a ~ i j ) t t t E ( Ω f 0 ) ( w ~ t t t f ) , j i
    + 0 t Ω s 0 q ~ , j ( a ~ i j ) t t t E ( Ω f 0 ) ( w ~ t t t f ) i ,
    and thus,
    K 2 C 0 t q ~ H 3 ( Ω s 0 ; R ) w ~ t t H 1 ( Ω s 0 ; R 3 ) η ~ H 3 ( Ω s 0 ; R 3 ) w ~ t t t L 2 ( Ω ; R 3 )
    + C 0 t q ~ H 3 ( Ω s 0 ; R ) w ~ t H 1 ( Ω s 0 ; R 3 ) w ~ H 3 ( Ω s 0 ; R 3 ) w ~ t t t L 2 ( Ω ; R 3 )
    + C 0 t q 0 + 0 q ~ t H 2 ( Ω s 0 ; R ) w ~ t t H 1 ( Ω s 0 ; R 3 ) η ~ H 3 ( Ω s 0 ; R 3 ) w ~ t t t H 1 ( Ω f 0 ; R 3 )
    + C 0 t q 0 + 0 q ~ t H 2 ( Ω s 0 ; R ) w ~ t H 1 ( Ω s 0 ; R 3 ) w ~ H 3 ( Ω s 0 ; R 3 ) w ~ t t t H 1 ( Ω f 0 ; R 3 )
    C ( w ~ , q ~ ) Z t 3 0 t q ~ H 3 ( Ω s 0 ; R ) + C ( w ~ , q ~ ) Z t 2 q 0 H 2 ( Ω s 0 ; R ) 0 t w ~ t t t H 1 ( Ω f 0 ; R 3 )
    + C t ( w ~ , q ~ ) Z t 2 q ~ t L 2 ( 0 , t ; H 2 ( Ω s 0 ; R ) ) 0 t w ~ t t t H 1 ( Ω f 0 ; R 3 )
    C t [ ( w ~ , q ~ ) Z t 4 + N ( ( q i ) i = 0 2 ) ] .
    The most difficult integral set at time t   on Ω s 0   and containing q ~   is K 3 = Ω s 0 q ~ t t ( a ~ i j ) t w ~ t t i , j ,   for which we apparently just have an estimate of the type | I 3 | C ( w ~ , q ~ ) Z t 2   (without any small parameter in front). We now explain how to treat this difficulty.
    Step 3. We first notice that
    K 3 = Ω s 0 q ~ t t , j ( a ~ i j ) t w ~ t t i + Γ 0 q ~ t t ( a ~ i j ) t w ~ t t i N j s .
    If we could say that q ~ t t   is L ( H 1 )   controlled, the L ( L 2 )   control of w ~ t t   would give us a suitable bound for K 3   . Whereas we have seen in the statement of our theorem that q t t   for the limit solution is indeed in L ( H 1 )   , we cannot seemingly get such a bound on the approximate pressures q ~ t t   . In order to get around this, we introduce similarly as in the previous step the extension to the solid domain of the velocity in the fluid. Since a similar integration by parts formula holds when we replace w ~ t t   by E ( Ω f 0 ) ( w ~ t t f )   , we deduce
    K 3 = Ω s 0 q ~ t t , j ( a ~ i j ) t w ~ t t i + Ω s 0 q ~ t t ( a ~ i j ) t E ( Ω f 0 ) ( w ~ t t f ) i , j + Ω s 0 q ~ t t , j ( a ~ i j ) t E ( Ω f 0 ) ( w ~ t t f ) i . (13.3)
    The easier term to estimate is K 3 2 = Ω s 0 q ~ t t ( a ~ i j ) t E ( Ω f 0 ) ( w ~ t t f ) i , j   , for which we have for an arbitrary δ > 0   :
    | K 3 2 | C q ~ t t L 2 ( Ω s 0 ; R ) Id + 0 t w ~ H 3 ( Ω s 0 ; R 3 ) u 0 + 0 t w ~ t H 2 ( Ω s 0 ; R 3 ) 1 4 w ~ H 3 ( Ω s 0 ; R 3 ) 3 4
    × [ w 2 H 1 ( Ω f 0 ; R 3 ) + t w ~ t t t L 2 ( 0 , t ; H 1 ( Ω f 0 ; R 3 ) ) ]
    C q ~ t t L 2 ( Ω s 0 ; R ) [ 1 + t ( w ~ , q ~ ) Z t ] [ N ( u 0 , ( w i ) i = 1 3 ) + t 1 4 ( w ~ , q ~ ) Z t 1 4 ] ( w ~ , q ~ ) Z t 3 4
    × [ N ( u 0 , ( w i ) i = 1 3 ) + t 1 2 ( w ~ , q ~ ) Z t ]
    C ( w ~ , q ~ ) Z t [ N ( u 0 , ( w i ) i = 1 3 ) + t 1 4 ( w ~ , q ~ ) Z t 1 4 ] ( w ~ , q ~ ) Z t 3 4
    × [ N ( u 0 , ( w i ) i = 1 3 ) + t 1 2 ( w ~ , q ~ ) Z t ] 2
    δ ( w ~ , q ~ ) Z t 2 + C δ N ( u 0 , ( w i ) i = 1 3 ) + C δ t 1 4 ( w ~ , q ~ ) Z t 4 . (13.4)
    For the first integral, the nonlinear elastodynamics equation in Ω s 0   provides
    q ~ t t = a ~ 1 [ w ~ t t t + κ L w ~ t t + c i j k l [ ( η ~ , i η ~ , j δ i j ) η ~ , k ] t t , l 2 a ~ t q ~ t a ~ t t q ~ + f t t + κ h ] ,
    leading us for K 3 1 = Ω s 0 q ~ t t , j ( a ~ i j ) t w ~ t t i   to (since a ~ 1 = η ~   in virtue of det η ~ = 1   ),
    K 3 1 = Ω s 0 [ η ~ [ w ~ t t t + c i j k l [ ( η ~ , i η ~ , j δ i j ) η ~ , k ] t t , l 2 a ~ t q ~ t a ~ t t q ~ + f t t + κ h ] ] j ( a ~ i j ) t w ~ t t i
    + κ Ω s 0 [ η ~ [ L w ~ t t ] ] j ( a ~ i j ) t w ~ t t i . (13.5)
    The integrals on the first line of this equality do not give any trouble and can be estimated in the same fashion. For instance, we have for K 3 3 = Ω s 0 [ η ~ [ c i j k l [ ( η ~ , i η ~ , j δ i j ) w ~ t , l k ] ] j ( a ~ i j ) t w ~ t t i ,  
    | K 3 3 | C Id + 0 t w ~ H 3 ( Ω s 0 ; R 3 ) 4 w ~ t H 2 ( Ω s 0 ; R 3 ) w 2 + 0 t w ~ t t t L 2 ( Ω s 0 ; R 3 )
    × u 0 + 0 t w ~ t H 2 ( Ω s 0 ; R 3 ) 1 4 w ~ H 3 ( Ω s 0 ; R 3 ) 3 4
    δ ( w ~ , q ~ ) Z t 2 + C δ t 1 4 ( w ~ , q ~ ) Z t 7 + C δ N ( u 0 , ( w i ) i = 1 3 ) ) . (13.6)
    Now, the difficult term to handle is K 3 4 = κ Ω s 0 [ η ~ [ L w ~ t t ] ] j ( a ~ i j ) t w ~ t t i   . We first write the divergence form L w ~ t t p = σ , m m p ( w ~ t t )   , and integrate by parts:
    K 3 4 = κ Ω s 0 [ η ~ , m [ σ m p ( w ~ t t ) ] p = 1 3 ] j ( a ~ i j ) t w ~ t t i
    κ Ω s 0 [ η ~ [ σ m p ( w ~ t t ) ] p = 1 3 ] j [ ( a ~ i j ) t w ~ t t i ] , m
    + κ Γ 0 [ η ~ [ σ m p ( w ~ t t ) ] p = 1 3 ] j ( a ~ i j ) t w ~ t t i N m s ,
    leading us to
    | K 3 4 κ Γ 0 [ η ~ [ σ m p ( w ~ t t ) ] p = 1 3 ] j ( a ~ i j ) t w ~ t t i N m s | C κ ( w ~ , q ~ ) Z t 5 , (13.7)
    and thus by putting together ( 13.5 ), ( 13.6 ) and ( 13.7 ),
    | Ω s 0 q ~ t t , j ( a ~ i j ) t w ~ t t i κ Γ 0 [ η ~ [ σ m p ( w ~ t t ) ] p = 1 3 ] j ( a ~ i j ) t w ~ t t i N m s |
    C κ ( w ~ , q ~ ) Z t 5 + C δ t 1 4 ( w ~ , q ~ ) Z t 7 + δ ( w ~ , q ~ ) Z t 2 (13.8)
    + C δ [ N ( u 0 , ( w i ) i = 1 3 ) + N ( ( q i ) i = 0 2 ) + M ( f , κ g , κ h ) ] . (13.9)
    Now, the apparent problem comes from the term σ m p ( w ~ t t )   on Γ 0   that should be taken in H 0.5 ( Γ 0 ; R )   , which is troublesome since the norm in Z t   appropriate for our limit process only contains its L ( 0 , t ; L 2 ( Ω s 0 ; R ) )   norm. In order to circumvent this, we notice that we also have, since E ( Ω f 0 ) ( w ~ t t )   is at least as smooth as w ~ t t   in Ω s 0   ,
    | Ω s 0 q ~ t t , j ( a ~ i j ) t E ( Ω f 0 ) ( w ~ t t ) i κ Γ 0 [ η ~ [ σ m p ( w ~ t t ) ] p = 1 3 ] j ( a ~ i j ) t E ( Ω f 0 ) ( w ~ t t f ) i N m |
    C κ ( w ~ , q ~ ) Z t 5 + C δ t 1 4 ( w ~ , q ~ ) Z t 7 + δ ( w ~ , q ~ ) Z t 2
    + C δ [ N ( u 0 , ( w i ) i = 1 3 ) + N ( ( q i ) i = 0 2 ) + M ( f , κ g , κ h ) ] , (13.10)
    leading us, since w ~ = E ( Ω f 0 ) ( w ~ f )   on Γ 0   , to
    | Ω s 0 q ~ t t , j ( a ~ i j ) t w ~ t t i Ω s 0 q ~ t t , j ( a ~ i j ) t E ( Ω f 0 ) ( w ~ t t ) i |
    C κ ( w ~ , q ~ ) Z t 5 + C δ t 1 4 ( w ~ , q ~ ) Z t 7 + δ ( w ~ , q ~ ) Z t 2
    + C δ [ N ( u 0 , ( w i ) i = 1 3 ) + N ( ( q i ) i = 0 2 ) + M ( f , κ g , κ h ) ] . (13.11)
    Thus, by using ( 13.3 ), ( 13.4 ) and ( 13.11 ), we have
    | I 3 | ( C κ + C δ t 1 4 ) ( w ~ , q ~ ) Z t 7 + δ ( w ~ , q ~ ) Z t 2
    + C δ [ N ( u 0 , ( w i ) i = 1 3 ) + N ( ( q i ) i = 0 2 ) + M ( f , κ g , κ h ) ] . (13.12)
    Thus, we finally arrive to estimates analogous to ( 8.7 ) and ( 8.8 ), with the right-hand side being of the same type as in ( 13.12 ).
  • Estimate on w ~ t t   and w ~ t   . With the same arguments as in the next subsection, we have for n = 2 , 1   :
    t n w ~ L 2 ( 0 , t ; H 4 n ( Ω f 0 ; R 3 ) ) 2 + t n q ~ L 2 ( 0 , t ; H 3 n ( Ω f 0 ; R 3 ) ) 2 + t n η ~ L ( 0 , T ; H 4 n ( Ω s 0 ; R 3 ) ) 2
    + κ 2 t n w ~ L 2 ( 0 , t ; H 4 n ( Ω s 0 ; R 3 ) ) 2 + t n q ~ L 2 ( 0 , t ; H 3 n ( Ω s 0 ; R 3 ) ) 2
    C δ [ N ( u 0 , ( w i ) i = 1 3 ) + M ( f , κ g , κ h ) + N ( ( q i ) i = 0 2 ) ]
    + ( C κ + C δ t 1 4 ) ( w ~ , q ~ ) Z t 7 + C δ ( w ~ , q ~ ) Z t 2 . (13.13)
    We now explain on the case of the highest space derivative how to obtain elliptic estimates independent of κ   , since the addition of the pressure term does not allow to use Lemma  6.1 directly in the present case.
  • Estimate on η ~   in Ω s 0   .
  • Regularity of the trace of η ~   . First, by proceeding as in Section  8 , and as for the case of the highest order time derivative, we get an estimate for the trace similar to ( 8.17 ), with a majorant of the same type as in ( 13.12 ). We explain hereafter how to handle the estimates related to the pressure in the solid in order to get this trace estimate since difficulties different than in the higher order time derivative problem appear in the higher order space derivative problem.
    Step 1. Let Q 1 = 0 t R 3 [ Q b ~ i j ] , α 1 α 2 α 3 [ ζ 2 W i ] , α 1 α 2 α 3 j   .
    Then,
    Q 1 = Q 2 + Q 3 + Q 4 ,
    with
    Q 2 = 0 t R 3 Q b ~ i j , α 1 α 2 α 3 [ ζ 2 W i ] , α 1 α 2 α 3 j ,
    Q 3 = 0 t R 3 [ [ Q b ~ i j ] , α 1 α 2 α 3 Q , α 1 α 2 α 3 b ~ i j Q b ~ i j , α 1 α 2 α 3 ] [ ζ 2 W i ] , α 1 α 2 α 3 j ,
    Q 4 = 0 t R 3 Q , α 1 α 2 α 3 b ~ i j [ ζ 2 W i ] , α 1 α 2 α 3 j .
    For Q 2   , we first notice that for θ = η ~ Ψ   , if ε i j k   is the sign of the permutation between { i , j , k }   and { 1 , 2 , 3 }   if i , j , k   are distinct, and is set to zero otherwise,
    b ~ i j , α 1 α 2 α 3 W i , α 1 α 2 α 3 j = 1 2 ε m n i ε p q j [ θ , p m θ , q n ] , α 1 α 2 α 3 W , j α 1 α 2 α 3 i
    = ε m n i ε p q j θ , p α 1 α 2 α 3 m θ , q n W , j α 1 α 2 α 3 i
    + 1 2 σ Σ 3 ε m n i ε p q j θ , p α σ ( 1 ) m θ , q α σ ( 2 ) α σ ( 3 ) n W , j α 1 α 2 α 3 i
    + 1 2 σ Σ 3 ε m n i ε p q j θ , p α σ ( 1 ) α σ ( 2 ) m θ , q α σ ( 3 ) n W , j α 1 α 2 α 3 i
    = 1 2 ε m n i ε p q j [ θ , p α 1 α 2 α 3 m θ , j α 1 α 2 α 3 i ] t θ , q n
    + 1 2 σ Σ 3 ε m n i ε p q j θ , p α σ ( 1 ) m θ , q α σ ( 2 ) α σ ( 3 ) n W , j α 1 α 2 α 3 i
    + 1 2 σ Σ 3 ε m n i ε p q j θ , p α σ ( 1 ) α σ ( 2 ) m θ , q α σ ( 3 ) n W , j α 1 α 2 α 3 i ,
    where we have use ε m n i ε p q j = ε n m i ε q p j   on the second equality and ε m n i ε p q j = ε i n m ε j q p   on the third one. Thus,
    Q 2 = 0 t R 3 Q b ~ i j , α 1 α 2 α 3 [ [ ζ 2 W ] i , α 1 α 2 α 3 j ζ 2 W , j α 1 α 2 α 3 i ]
    1 2 0 t R 3 σ Σ 3 ε m n i ε p q j [ Q θ , p α σ ( 1 ) m θ , q α σ ( 2 ) α σ ( 3 ) n ζ 2 ] , α 1 W , j α 2 α 3 i
    1 2 0 t R 3 σ Σ 3 ε m n i ε p q j [ Q θ , p α σ ( 1 ) α σ ( 2 ) m θ , q α σ ( 3 ) n ζ 2 ] , α 1 W , j α 2 α 3 i
    1 2 0 t R 3 ε m n i ε p q j ζ 2 θ , p α 1 α 2 α 3 m θ , j α 1 α 2 α 3 i [ Q θ , q n ] t
    + 1 2 [ R 3 ε m n i ε p q j ζ 2 θ , p α 1 α 2 α 3 m θ , j α 1 α 2 α 3 i Q θ , q n ] 0 t ,
    showing
    | Q 2 | C t ( w ~ , q ~ ) Z t 4 + | R 3 ε m n i ε p q j ζ 2 θ , p α 1 α 2 α 3 m θ , q α 1 α 2 α 3 n Q θ , j i | ( t ) + N ( ( q i ) i = 0 2 ) .
    In order to estimate the remaining term, we notice by integrating by parts twice for Q 5 = R 3 ε m n i ε p q j ζ 2 θ , p α 1 α 2 α 3 m θ , q α 1 α 2 α 3 n Q θ , j i   that
    Q 5 = ε m n i ε p q j R 3 ζ 2 θ , q α 1 α 2 α 3 m θ , p α 1 α 2 α 3 n Q θ , j i
    ε m n i ε p q j R 3 [ ( ζ 2 Q θ , j i ) , p θ , α 1 α 2 α 3 m θ , q α 1 α 2 α 3 n ( ζ 2 Q θ , j i ) , q θ , α 1 α 2 α 3 m θ , p α 1 α 2 α 3 n ]
    + ε m n i ε p q j x 3 = 0 ζ 2 Q θ , α 1 α 2 α 3 m [ θ , q α 1 α 2 α 3 n θ , j i ( e 3 ) p θ , p α 1 α 2 α 3 n θ , j i ( e 3 ) q ] .
    Since ε p q j = ε q p j   , we then infer
    2 Q 5 = ε m n i ε p q j R 3 θ , α 1 α 2 α 3 m [ ( ζ 2 Q θ , j i ) , p θ , q α 1 α 2 α 3 n ( ζ 2 Q θ , j i ) , q θ , p α 1 α 2 α 3 n ]
    + ε m n i ε p q j x 3 = 0 ζ 2 Q θ , α 1 α 2 α 3 m [ θ , q α 1 α 2 α 3 n θ , j i ( e 3 ) p θ , p α 1 α 2 α 3 n θ , j i ( e 3 ) q ] .
    Now, if we note θ f = E ( Ω f 0 ) ( η ~ f ) Ψ   , we also have for R 3 ε m n i ε p q j ζ 2 θ f , p α 1 α 2 α 3 m θ , q α 1 α 2 α 3 n Q θ , j i   a similar formula. Since θ , α 1 α 2 α 3 m = θ f , α 1 α 2 α 3 m   on { x 3 = 0 }   , we then have
    2 Q 5 = ε m n i ε p q j R 3 ( θ θ f ) , α 1 α 2 α 3 m [ ( ζ 2 Q θ , j i ) , p θ , q α 1 α 2 α 3 n ( ζ 2 Q θ , j i ) , q θ , p α 1 α 2 α 3 n ]
    + 2 R 3 ε m n i ε p q j ζ 2 θ f , p α 1 α 2 α 3 m θ , q α 1 α 2 α 3 n Q θ , j i ,
    leading us to
    | Q 5 | C Id + 0 t w ~ H 3 ( Ω s 0 ; R 3 ) 1 4 η ~ H 4 ( Ω s 0 ; R 3 ) 7 4 q 0 + 0 t q ~ t H 2 ( Ω s 0 ; R ) Id + 0 t w ~ H 3 ( Ω s 0 ; R 3 )
    + C Id + 0 t w ~ H 3 ( Ω f 0 ; R 3 ) 1 4 η ~ H 4 ( Ω s 0 ; R 3 ) 7 4 q 0 + 0 t q ~ t H 2 ( Ω s 0 ; R ) Id + 0 t w ~ H 3 ( Ω s 0 ; R 3 )
    + C Id + 0 t w ~ f H 4 ( Ω f 0 ; R 3 ) η ~ H 4 ( Ω s 0 ; R 3 ) q 0 + 0 t q ~ t H 2 ( Ω s 0 ; R ) Id + 0 t w ~ H 3 ( Ω s 0 ; R 3 )
    δ ( w ~ , q ~ ) Z t 2 + C t ( w ~ , q ~ ) Z t 4 + C δ [ N ( u 0 , ( w i ) i = 1 3 ) + N ( ( q i ) i = 0 2 ) ]
    Step 2. We see by integrating by parts with respect to the direction α 1   that we have
    | Q 3 | C t ( w ~ , q ~ ) Z t 4 .
    Step 3. Next, Q 4 = Q 6 + Q 7   , where
    Q 6 = 0 t R 3 Q , α 1 α 2 α 3 b ~ i j ζ 2 W i , α 1 α 2 α 3 j
    Q 7 = 0 t R 3 Q , α 1 α 2 α 3 b ~ i j [ [ ζ 2 W ] i , α 1 α 2 α 3 j ζ 2 W i , α 1 α 2 α 3 j ] .
    We first have
    | Q 7 | C 0 t [ q ~ H 3 ( Ω s 0 ; R ) w ~ H 3 ( Ω s 0 ; R 3 ) η ~ H 3 ( Ω s 0 ; R 3 ) 2 ] C t ( w ~ , q ~ ) Z t 4 .
    For Q 6   the divergence condition b ~ i j W i , j = 0   on S u p p ζ   implies
    Q 6 = 0 t R 3 Q , α 1 α 2 α 3 ζ 2 [ b ~ i j W i , α 1 α 2 α 3 j ( b ~ i j W i , j ) , α 1 α 2 α 3 ] ,
    which in turn provides,
    | Q 6 | C 0 t [ q ~ H 3 ( Ω s 0 ; R ) w ~ H 3 ( Ω s 0 ; R 3 ) η ~ H 4 ( Ω s 0 ; R 3 ) 2 ] C t ( w ~ , q ~ ) Z t 4 ,
    which concludes the estimates on the pressure terms in the solid, justifying why we obtain a trace estimate similar as ( 8.17 ) with a majorant of the type of the right-hand side of ( 13.12 ). Now, we turn our attention to the recovery of the regularity in the solid, which will need some justifications since we cannot directly apply Lemma  6.1 .
  • Regularity in the incompressible solid First, with the introduction of F ~ = w ~ t + c i j k l [ ( η ~ , i η ~ , j δ i j ) η ~ , k ] , l L η ~ + f + κ h in Ω s 0 .   and of r ~   , the solution in Ω s 0   of
    κ 2 r ~ t + r ~ = q ~ ,
    r ~ ( 0 ) = q 0 ,
    we have for the nonlinear elastodynamics κ 2 η ~ L w ~ η ~ L η ~ + [ κ 2 r ~ t + r ~ ] = η ~ F ~ in Ω s 0 ,   i.e.,
    κ 2 [ η ~ L η ~ + r ~ ] t η ~ L η ~ + r ~ = η ~ F ~ κ 2 w ~ L η ~ in Ω s 0 . (13.14)
    We now apply Lemma  6.1 to this equation, leading us to
    sup [ 0 , t ] η ~ L ( η ~ ) + r ~ H 2 ( Ω s 0 ; R 3 ) sup [ 0 , t ] η ~ F ~ κ 2 w ~ L η ~ H 2 ( Ω s 0 ; R 3 )
    + L ( Id ) + q ~ 0 H 2 ( Ω s 0 ; R 3 ) ,
    and, with H ~ = L ( η ~ ) + r ~   , to
    sup [ 0 , t ] H ~ H 2 ( Ω s 0 ; R 3 )
    sup [ 0 , t ] [ η ~ F ~ κ 2 w ~ L η ~ H 2 ( Ω s 0 ; R 3 ) + ( η ~ Id ) L ( η ~ ) H 2 ( Ω s 0 ; R 3 ) ] + N ( ( q i ) i = 0 2 ) . (13.15)
    We then want to use elliptic regularity on the system:
    L η ~ + r ~ = H ~ , in Ω s 0 , (13.16a)
    div η ~ = ( a ~ i j + δ i j ) η ~ i , j + 3 , in Ω s 0 , (13.16b)
    η ~ = η ~ | Γ 0 on Γ 0 , (13.16c)
    where the trace on Γ 0   is estimated as we explained in the previous subsection.
    Now, for the divergence condition in Ω s 0   , we notice that:
    [ ( a ~ i j δ i j ) η ~ i , j ] , i 1 i 2 i 3 = a ~ i j , i 1 i 2 i 3 η ~ i , j + ( a ~ i j δ i j ) η ~ i , j i 1 i 2 i 3
    + σ Σ 3 [ a ~ i j , i σ ( 1 ) η ~ i , j i σ ( 2 ) i σ ( 3 ) + a ~ i j , i σ ( 1 ) i σ ( 2 ) η ~ i , j i σ ( 3 ) ] .
    For the apparently problematic first term on the right-hand side, we first notice that
    a ~ i j , i 1 η ~ i , j = 1 2 ε m n i ε p q j ( η ~ , p m η ~ , q n ) , i 1 η ~ , j i = ε m n i ε p q j η ~ , p i 1 m η ~ , q n η ~ , j i
    = ε i n m ε j q p η ~ , j i 1 i η ~ , q n η ~ , p m = 2 a ~ i j η ~ i , j i 1 ,
    which with the condition a ~ i j η ~ i , j = 3   , provides 0 = a ~ i j , i 1 η ~ i , j ,   and thus
    a ~ i j , i 1 i 2 i 3 η ~ i , j = a ~ i j , i 1 i 2 η ~ i , j i 3 a ~ i j , i 1 i 3 η ~ i , j i 2 a ~ i j , i 1 η ~ i , j i 2 i 3 .
    We then deduce that
    ( a ~ i j δ i j ) η ~ i , j H 3 ( Ω s 0 ; R ) ( t ) δ ( w ~ , q ~ ) Z t 2 + C δ N ( u 0 , ( w i ) i = 1 3 ) + C t ( w ~ , q ~ ) Z t 3 . (13.17)
    Now, with ( 13.15 ) and ( 13.17 ), elliptic regularity on ( 13.16 ) provides for η ~ L ( 0 , t ; H 4 ( Ω s 0 ; R 3 ) ) 2 + r ~ 1 | Ω s 0 | Ω s 0 r ~ L ( 0 , t ; H 3 ( Ω s 0 ; R ) ) 2   a bound of the same type as the right-hand side of ( 13.12 ), with however the norms in Z t   being replaced by the norms in Z ~ t   , due to the term κ w ~ L η ~ H 2 ( Ω s 0 ; R 3 )   appearing on the right-hand side of ( 13.15 ), that we bound by
    C κ w ~ ( t ) H 3 ( Ω s 0 ; R 3 ) L ( Id ) + 0 t L w ~ H 2 ( Ω s 0 ; R 3 ) C κ w ~ ( t ) H 3 ( Ω s 0 ; R 3 ) t w ~ L 2 ( 0 , t ; H 4 ( Ω s 0 ; R 3 ) )
    C t ( w ~ , q ~ ) Z ~ t 2 .
    We now turn our attention to the pressure, that we just need to control in L 2 ( 0 , t ; H 3 ( Ω s 0 ; R ) )   . In order to do so, we notice from ( 13.14 ) that we have for K ~ = κ 2 [ L w ~ + r ~ t ]   :
    K ~ = κ 2 [ η ~ Id ] L w ~ + η ~ L η ~ r ~ + η ~ F ~ κ 2 w ~ L η ~ ,
    which with the previous estimate on η ~   and r ~   shows that we have a bound on K ~ L 2 ( 0 , t ; H 2 ( Ω s 0 ; R 3 ) ) 2   of the same type as the right-hand side of ( 13.12 ), but where the norms in Z t   are replaced by norms in Z ~ t   due to the estimate in L 2 ( H 2 )   of κ [ η ~ Id ] L w ~   . Now, elliptic regularity on the system:
    L κ w ~ + κ r t = 2 K ~ , in Ω s 0 ,
    div κ w ~ = κ ( a ~ i j + δ i j ) w ~ i , j , in Ω s 0 ,
    κ w ~ = κ w ~ | Γ 0 f on Γ 0 ,
    provides after integrating in time an estimate for κ 2 [ w ~ L 2 ( 0 , t ; H 4 ( Ω s 0 ; R 3 ) ) 2 + r ~ t 1 | Ω s 0 | Ω s 0 r ~ t L 2 ( 0 , t ; H 3 ( Ω s 0 ; R ) ) 2 ]   with a bound similar as in ( 13.12 ), still with the norms in Z t   being replaced by norms in Z ~ t   .
    Thus, we obtain for q ~ 1 | Ω s 0 | Ω s 0 q ~ L 2 ( 0 , t ; H 3 ( Ω s 0 ; R ) ) 2   the same type of estimate as well. Given our estimate on q ~ t t   , this also implies the same type of majoration for q ~ L 2 ( 0 , t ; H 3 ( Ω s 0 ; R ) ) 2   .
    Thus, we are lead to
    ( w ~ , q ~ ) Z ~ t 2 ( C κ + C δ t 1 4 ) ( w ~ , q ~ ) Z ~ t 8 + δ ( w ~ , q ~ ) Z ~ t 2
    + C δ [ N ( u 0 , ( w i ) i = 1 3 ) + N ( ( q i ) i = 0 2 ) + M ( f , κ g , κ h ) ] ,
    which leads as in Section  9 to the introduction of a polynomial, this time of degree 4, which does not bring any substantial change with respect to Section  9 .
    Note that the addition of C κ ( w ~ , q ~ ) Z t 8   does not create any difficulty since a small κ 1   is chosen at the same stage as t 1   , and the conclusion is similar as in Section  9 from the continuity of ( w ~ , q ~ ) Z ~ t   on [ 0 , T κ ]   which is established in the same way as the continuity of ( w ~ , q ~ ) Z t   . We then infer that there is a time of existence of κ   for our smoothed problems, with a bound on ( w ~ , q ~ ) Z ~ T   and thus on ( w ~ , q ~ ) Z T   independent of κ   . Existence follows then by weak convergence in Y T   and uniqueness can be established similarly as for the compressible case in Section  11 .

Acknowledgments

DC was partially supported by the National Science Foundation under grant NSF ITR-0313370. SS was partially supported by the National Science Foundation under grants DMS-0105004 and NSF ITR-0313370. References

  1. J.T. Beale, Large-time regularity of viscous surface waves, Arch. Rational Mech. Anal., 84 (1983/84), 307–352.
  2. A. Chambolle, B. Desjardins, M.J. Esteban, C. Grandmont, Existence of weak solutions for an unsteady fluid-plate interaction problem, Preprint.
  3. P.G. Ciarlet, Mathematical Elasticity, Volume I: Three-Dimensional Elasticity, North-Holland, Amsterdam, 1986.
  4. C. Conca, J. San Martin, M. Tucsnak, Existence of solutions for the equations modelling the motion of a rigid body in a viscous fluid, Comm. Partial Differential Equations, 25 (2000), 1019-1042.
  5. D. Coutand, S. Shkoller, On the motion of an elastic solid inside of an incompressible viscous fluid, to appear in Arch. Rational Mech. Anal., http://arxiv.org/abs/math.AP/0407365.
  6. C.M. Dafermos, W.J. Hrusa, Energy methods for quasilinear hyperbolic initial boundary value problems. Applications to elastodynamics, Arch. Rational Mech. Anal., 87 (1985), 267-292.
  7. B. Desjardins, M.J. Esteban, Existence of weak solutions for the motion of rigid bodies in a viscous fluid, Arch. Rational Mech. Anal., 146 (1999), 59–71.
  8. B. Desjardins, M.J. Esteban, C. Grandmont, P. Le Tallec, Weak solutions for a fluid-structure interaction problem, Rev. Mat. Complut., 14 (2001), 523–538.
  9. D.G. Ebin, S.R. Simanca, Deformations of incompressible bodies with free boundaries, Arch. Rational Mech. Anal., 103 (1992), 61-97.
  10. F. Flori, P. Orenga, Fluid-structure interaction: analysis of a 3-D compressible model, Ann. Inst. H. Poincaré Anal. Non Linéaire, 17 (2000), 753-777.
  11. C. Grandmont, Existence et unicite de solutions d'un probleme de couplage fluide-structure bidimensionnel stationnaire, C. R. Acad. Sci. Paris Sér. I., 326 (1998), 651–656.
  12. C. Grandmont, Y. Maday, Existence for unsteady fluid-structure interaction problem, Math. Model. Numer. Anal., 34 (2000), 609–636.
  13. M.D. Gunzburger, H-C. Lee, G.A. Seregin, Global existence of weak solutions for viscous incompressible flows around a moving rigid body in three dimensions,, J. Math. Fluid Mech., 2 (2000), 219–266.
  14. M. Heil, Stokes flow in an elastic tube – a large-displacement fluid-structure interaction problem, Int. J. Numer. Meth. Fluids, 28 (1998), 243-265.
  15. M. Hillairet, D. Serre, Chute stationnaire d'un solide dans un fluide visqueux incompressible le long d'un plan incline, Ann. Inst. H. Poincaré Anal. Non Linéaire, 20 (2003), 779-803.
  16. T.J.R. Hughes, T. Kato, J.E. Marsden, Well-posed quasi-linear hyperbolic systems with applications to non-linear elastodynamics and general relativity, Arch. Rational Mech. Anal., 63 (1977), 273-294.
  17. W.J. Hrusa, M. Renardy, An existence theorem for the Dirichlet problem in the elastodynamics of incompressible materials, Arch. Rational Mech. Anal., 102 (1988), 95-117.
  18. C. Liu, N.J. Walkington, An Eulerian description of fluids containing visco-elastic particles, Arch. Rational Mech. Anal., 159 (2001), 229-252.
  19. M. Sablé-Tougeron, Existence pour un probleme de l'elastodynamique Neumann non lineaire en dimension 2, Arch. Rational Mech. Anal., 101 (1988), 261-292.
  20. D. Serre, Chute libre d'un solide dans un fluide visqueux incompressible: Existence, Japan J. Appl. Math., 4 (1987), 33–73.
  21. V.A. Solonnikov, Solvability of the problem of evolution of a viscous incompressible fluid bounded by a free surface on a finite time interval, St. Petersburg Math. J., 3 (1992), 189–220.
  22. H.F. Weinberger, Variational properties of steady fall in Stokes flow, J. Fluid Mech., 52 1972, 321–344.

E-mail address : coutand@math.ucdavis.edu E-mail address : shkoller@math.ucdavis.edu