• Aucun résultat trouvé

An improved result for the full justification of asymptotic models for the propagation of internal waves

N/A
N/A
Protected

Academic year: 2021

Partager "An improved result for the full justification of asymptotic models for the propagation of internal waves"

Copied!
28
0
0

Texte intégral

(1)

HAL Id: hal-01094516

https://hal.archives-ouvertes.fr/hal-01094516

Preprint submitted on 12 Dec 2014

HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

Samer Israwi, Ralph Lteif, Raafat Talhouk

To cite this version:

Samer Israwi, Ralph Lteif, Raafat Talhouk. An improved result for the full justification of asymptotic

models for the propagation of internal waves. 2014. �hal-01094516�

(2)

for the propagation of internal waves

Samer Israwi Ralph Lteif ∗† Raafat Talhouk December 12, 2014

Abstract

We consider here asymptotic models that describe the propagation of one-dimensional internal waves at the interface between two layers of immiscible fluids of different densities, under the rigid lid assumption and with uneven bottoms. The aim of this paper is to show that the full justification result of the model obtained by Duchˆene, Israwi and Talhouk [to appear in SIAM J. Math. Anal , (arXiv:1304.4554v2)], in the sense that it is consistent, well-posed, and that its solutions remain close to exact solutions of the full Euler system with corresponding initial data, can be improved in two directions. The first direction is taking into account medium amplitude topography variations and the second direction is allowing strong nonlinearity using a new pseudo-symmetrizer, thus canceling out the smallness assumption of the Camassa-Holm regime for the existence and uniqueness results.

1 Introduction

1.1 Presentation of the problem

In this work, we are interested in the propagation of internal waves in a two-fluid system, which consists in two layers of immiscible, homogeneous, ideal, incompressible fluids of different densities, under the only influence of gravity. The domain of the two layers is infinite in the horizontal space variable (assumed to be of dimension d = 1). We assume medium amplitude topography variations (non-flat bottom) and that the surface is confined by a flat rigid lid.

The derivation of the governing equations of such a system is not new: see [2], [5] and [16]. Under the aforementioned configuration, the governing equations describing the evolution of the flow may be reduced to a system of two coupled evolution equations located at the interface between the two layers (see [12, 32] for the water-wave configuration, and [5] for the bi-fluidic case), named full Euler system. In particular, the well-posedness in Sobolev spaces of the Cauchy problem for bi-fluidic full Euler system has been answered satisfactorily in the presence of a small amount of surface tension, see [22] (that is, with an existence of solutions on a time scale consistent with physical observations). However, the theoretical study of this system is extremely challenging.

Because of the complexity of these equations their solutions are very difficult to describe, this explains why a great deal of interests has been drawn to asymptotic models, in order to predict accurately the main behavior of the system, provided some parameters describing the domain and nature of the flow are small. Parameters of interests include

µ = d 2 1

λ 2 , ǫ = a

d 1 , β = a b

d 1 , δ = d 1

d 2 , γ = ρ 1

ρ 2 , Bo = g(ρ 2 − ρ 1 )λ 2

σ ,

where a(resp. a b ) is the maximal vertical deformation of the interface(resp. bottom) with respect to its rest position; λ is a characteristic horizontal length; d 1 (resp. d 2 ) is the depth of the upper (resp. lower)

Laboratory of Mathematics-EDST and Faculty of Sciences I, Lebanese University, Beirut, Lebanon.

(s_israwi83@hotmail.com)(rtalhouk@ul.edu.lb)

LAMA, UMR 5127 CNRS, Universit´ e de Savoie, 73376 Le Bourget du lac cedex, France. (ralphlteif_90@hotmail.com)

1

(3)

layer; and ρ 1 (resp. ρ 2 ) is the density of the upper (resp. lower) layer, g the gravitational acceleration, σ the interfacial tension coefficient and Bo the classical Bond number, which measures the ratio of gravity forces over capillary forces. In the following we use bo = µBo instead of the classical Bond number, Bo. Mathematically speaking, µ and ǫ measure respectively the amount of dispersion and nonlinearity which will contribute to the evolution of internal waves. Let us introduce some earlier results directly related to the present paper.

Bona, Lannes and Saut [5] followed a strategy initiated in [3, 4] in the water-wave setting (one layer of fluid, with free surface) to derive a large class of models for different regimes, under the rigid-lid assumption, neglecting surface tension effects and with flat bottom, (see also [2] where a topography and surface tension is added to the system, and [14] where the rigid-lid assumption is removed). Shallow water (µ ≪ 1) asymptotic models for uni-dimensional internal waves have been derived and studied in the pioneer works of [25, 26, 27]. More recently, weakly (ǫ = O (µ)) and strongly (ǫ ∼ 1) nonlinear models in two-dimensions have been derived by Camassa and Choi in [9, 10]. They obtain bi-fluidic extensions of the classical shallow water (or Saint-Venant [29]), Boussinesq [6, 7] and Green-Naghdi [19, 30] models.

Similar systems have been derived in [28] (with the additional assumption of γ ≈ 1) and in [11] (using a different approach, i.e. making use of the Hamiltonian structure of the full Euler equations). The models derived in these papers are systematically justified by a consistency result: roughly speaking, sufficiently smooth solutions of the full Euler system satisfy the equations of the asymptotic model, up to a small remainder.

Contrarily to the water-wave case, large amplitude internal waves in a bi-fluidic system are known to generate Kelvin-Helmholtz instabilities that appear at high frequencies , so that surface tension is necessary in order to regularize the flow. However, when adding a small amount of surface tension, Lannes [22] proved that, thanks to a stability criterion, the problem becomes well-posed with a time of existence that does not vanish as the surface tension goes to zero and thus is consistent with the observations. Therefore, adding a small amount of surface tension at the interface in the Euler system guarantees the well-posedness of the system and does not change our asymptotic models. The study of Lannes focuses on the two-layer fluid system with a flat bottom (β = 0). However, we believe that the theory in the uneven bottom case does not differ much from the one in the flat bottom configuration.

In [20], the well-posedness and stability results have been proved for bi-fluidic shallow-water system , and in [15] for a class of Boussinesq-type systems, neglecting surface tension and under reasonable assump- tions on the flow (typically, the shear velocity must be sufficiently small). However, the well-posedness of the Green-Naghdi model in the bi-fluidic case is not clear, and similar systems have been proved to be ill-posed in [24], which has led to various propositions in order to overcome this difficulty; see [8, 13]

and references therein. Green-Naghdi models consist in higher order extensions, which has since been widely used in coastal oceanography, as it takes into account the dispersive effects neglected by the shallow-water model, thus are consistent with precision O (µ 2 ) instead of O (µ), and allows waves of greater amplitude (whereas Boussinesq models are limited to the long wave regime: ǫ = O (µ)). Finally, we mention the recent work of Xu [31], which studies and rigorously justify the so-called intermediate long wave system, obtained in a regime similar to ours: ǫ ∼ √ µ, but δ ∼ √ µ.

In this work, we present a Green-Naghdi type model in the Camassa-Holm (or medium amplitude) regime ǫ = O ( √ µ) and we assume medium amplitude topography variations. More precisely, we assume that there exists β max < ∞ such that

β = O ( √ µ) with β ∈ [0, β max ].

We improve in this paper the existence and uniqueness results obtained in [17] using a new pseudo-

symmetrizer, thus canceling out the smallness assumption ǫ = O ( √ µ) and we prove that our Green-

Naghdi model is fully justified as an asymptotic model for a set of dimensionless parameters limited to

the so-called Camassa-Holm regime, that we describe precisely below.

(4)

We first consider the so-called shallow water regime for two layers of comparable depths:

P SW ≡ n

(µ, ǫ, δ, γ, β, bo) : 0 < µ ≤ µ max , 0 ≤ ǫ ≤ 1, δ ∈ (δ min , δ max ),

0 ≤ γ < 1, 0 ≤ β ≤ β max , bo min ≤ bo ≤ ∞ o , (1.1) with given 0 ≤ µ max , δ min

1 , δ max , bo

min 1 , β max < ∞ .

The two additional key restrictions for the validity of the model (3.9) are as follows:

P CH

(µ, ǫ, δ, γ, β, bo) ∈ P SW : ǫ ≤ M √ µ, β ≤ M √ µ and ν ≡ 1 + γδ 3δ(γ + δ) − 1

bo ≥ ν 0

, (1.2) with given 0 ≤ M, ν 0

1 < ∞ .

We denote for convenience M SW ≡ max

µ max , δ min

1 , δ max , bo

min 1 , β max , M CH ≡ max

M SW , M, ν

0 1 .

We prove that the full Euler system is consistent with our model, and that our system is well-posed (in the sense of Hadamard) in Sobolev spaces, and stable with respect to perturbations of the equations.

1.2 Organization of the paper

We start by introducing in Section 2 the non-dimensionalized full Euler system and the Green-Naghdi model.

In Section 3, we present our new model where the asymptotic model is precisely derived and motivated.

Sections 4 and 5 contain the necessary ingredients for the proof of our results.

In Section 6, we explain the full justification of asymptotic models and we state its main ingredients.

We conclude this section with an inventory of the notations used throughout the present paper.

Notations In the following, C 0 denotes any nonnegative constant whose exact expression is of no importance.

The notation a . b means that a ≤ C 0 b and we write A = O (B) if A ≤ C 0 B.

We denote by C(λ 1 , λ 2 , . . . ) a nonnegative constant depending on the parameters λ 1 , λ 2 ,. . . and whose dependence on the λ j is always assumed to be nondecreasing.

We use the condensed notation

A s = B s + h C s i s>s , to express that A s = B s if s ≤ s and A s = B s + C s if s > s.

Let p be any constant with 1 ≤ p < ∞ and denote L p = L p ( R ) the space of all Lebesgue-measurable functions f with the standard norm

| f | L

p

= Z

R

| f (x) | p dx 1/p

< ∞ .

The real inner product of any functions f 1 and f 2 in the Hilbert space L 2 ( R ) is denoted by (f 1 , f 2 ) =

Z

R

f 1 (x)f 2 (x)dx.

The space L

= L

( R ) consists of all essentially bounded, Lebesgue-measurable functions f with the norm

| f | L

= ess sup | f (x) | < ∞ .

(5)

Let k ∈ N , we denote by W k,

= W k,

( R ) = { f ∈ L

, | f | W

k,

< ∞} , where | f | W

k,

= X

α

∈N

k

| ∂ x α f | L

. For any real constant s ≥ 0, H s = H s ( R ) denotes the Sobolev space of all tempered distributions f with the norm | f | H

s

= | Λ s f | L

2

< ∞ , where Λ is the pseudo-differential operator Λ = (1 − ∂ x 2 ) 1/2 . For a given µ > 0, we denote by H µ s+1 ( R ) the space H s+1 ( R ) endowed with the norm

·

2

H

µs+1

≡ ·

2

H

s

+ µ ·

2 H

s+1

.

For any function u = u(t, x) and v(t, x) defined on [0, T ) × R with T > 0, we denote the inner product, the L p -norm and especially the L 2 -norm, as well as the Sobolev norm, with respect to the spatial variable x, by (u, v) = (u(t, · ), v(t, · )), | u | L

p

= | u(t, · ) | L

p

, | u | L

2

= | u(t, · ) | L

2

, and | u | H

s

= | u(t, · ) | H

s

, respectively.

We denote L

([0, T ); H s ( R )) the space of functions such that u(t, · ) is controlled in H s , uniformly for t ∈ [0, T ):

u

L

([0,T);H

s

(

R

)) = ess sup

t

[0,T) | u(t, · ) | H

s

< ∞ . Finally, C k ( R ) denote the space of k-times continuously differentiable functions.

For any closed operator T defined on a Banach space X of functions, the commutator [T, f] is defined by [T, f ]g = T (f g) − f T (g) with f , g and f g belonging to the domain of T . The same notation is used for f an operator mapping the domain of T into itself.

2 Previously obtained models

2.1 The full Euler system

The equations governing the evolution of the aforedescribed system in the introduction read (using non-dimensionalized variables and the Zakharov/Craig-Sulem formulation) [12, 32]:

 

 

 

 

 

 

∂ t ζ − 1

µ G µ ψ = 0,

∂ t

H µ,δ ψ − γ∂ x ψ

+ (γ + δ)∂ x ζ + ǫ 2 ∂ x

| H µ,δ ψ | 2 − γ | ∂ x ψ | 2

= µǫ∂ x N µ,δ − µ γ+δ bo

x

k(ǫ

µζ)

ǫ

µ ,

(2.1)

where we denote

N µ,δ

1

µ G µ ψ + ǫ(∂ x ζ)H µ,δ ψ 2

− γ µ 1 G µ ψ + ǫ(∂ x ζ)(∂ x ψ) 2

2(1 + µ | ǫ∂ x ζ | 2 ) .

ζ(t, x) represent the deformation of the interface between the two layers and b(x) represent the defor- mation of the bottom, ψ is the trace of the velocity potential of the upper-fluid at the interface.

The function k(ζ) = − ∂ x

√ 1

1+

|

x

ζ

|2

∂ x ζ

denotes the mean curvature of the interface and σ the surface (or interfacial) tension coefficient.

We will refer to (2.1) as the full Euler system, and solutions of this system will be exact solutions of the problem.

Finally, G µ and H µ,δ are the so-called Dirichlet-Neumann operators, defined as follows:

Definition 2.2 (Dirichlet-Neumann operators) Let ζ ∈ H t

0

+1 ( R ), t 0 > 1/2, such that there exists h 0 > 0 with h 1 ≡ 1 − ǫζ ≥ h 0 > 0 and h 2 ≡ 1 δ +ǫζ − βb ≥ h 0 > 0, and let ψ ∈ L 2 loc ( R ), ∂ x ψ ∈ H 1/2 ( R ).

Then we define

G µ ψ ≡ G µ [ǫζ]ψ ≡ p

1 + µ | ǫ∂ x ζ | 2 ∂ n φ 1

| z=ǫζ = − µǫ(∂ x ζ)(∂ x φ 1 ) | z=ǫζ + (∂ z φ 1 ) | z=ǫζ , H µ,δ ψ ≡ H µ,δ [ǫζ, βb]ψ ≡ ∂ x φ 2 | z=ǫζ

= (∂ x φ 2 ) | z=ǫζ + ǫ(∂ x ζ)(∂ z φ 2 ) | z=ǫζ ,

(6)

where φ 1 and φ 2 are uniquely defined (up to a constant for φ 2 ) as the solutions in H 2 ( R ) of the Laplace’s problems:

µ∂ x 2 + ∂ z 2

φ 1 = 0 in Ω 1 ≡ { (x, z) ∈ R 2 , ǫζ(x) < z < 1 } ,

∂ z φ 1 = 0 on Γ t ≡ { (x, z) ∈ R 2 , z = 1 } , φ 1 = ψ on Γ ≡ { (x, z) ∈ R 2 , z = ǫζ } ,

(2.3)

µ∂ x 2 + ∂ z 2

φ 2 = 0 in Ω 2 ≡ { (x, z) ∈ R 2 , − 1 δ + βb(x) < z < ǫζ } ,

∂ n φ 2 = ∂ n φ 1 on Γ,

∂ n φ 2 = 0 on Γ b ≡ { (x, z) ∈ R 2 , z = − 1 δ + βb(x) } .

(2.4)

The existence and uniqueness of a solution to (2.3)-(2.4), and therefore the well-posedness of the Dirichlet-Neumann operators follow from classical arguments detailed, for example, in [23].

2.2 The Green-Naghdi model

The key ingredient for constructing shallow water asymptotic models lies in the expansion given in [14, 18]

of the Dirichlet-Neumann operators, with respect to the shallowness parameter, µ. Thanks to such an expansion, one is able to obtain the so-called Green-Naghdi model (for internal waves). This model has been introduced in [16] (with a flat bottom) and generalized in [18]. It is also convenient to introduce a new velocity variable, namely the shear mean velocity v is equivalently defined as

v ≡ u 2 − γu 1 (2.5)

where u 1 and u 2 are the horizontal velocities integrated across each layer:

u 1 (t, x) = h 1

1

(t,x)

R 1

ǫζ(t,x) ∂ x φ 1 (t, x, z) dz and u 2 (t, x) = h 1

2

(t,x)

R ǫζ (t,x)

1δ

+βb(x) ∂ x φ 2 (t, x, z) dz, where φ 1 and φ 2 are the solutions to the Laplace’s problems (2.3)-(2.4).

The expansions of the Dirichlet-Neumann operators may be written in terms of the new variable v.

Plugging the expansions given in [18, Proposition 7] into the full Euler system (2.1), and withdraw- ing all O (µ 2 ) terms yields ( in the unidimensional case d = 1),

 

 

 

 

 

 

 

 

∂ t ζ + ∂ x

h 1 h 2

h 1 + γh 2 v

= 0,

∂ t v + µ Q [h 1 , h 2 ]v

!

+ (γ + δ)∂ x ζ + ǫ 2 ∂ x

h 2 1 − γh 2 2 (h 1 + γh 2 ) 2 v 2

= µǫ∂ x R [h 1 , h 2 ]v

+ µ γ + δ bo ∂ x 3 ζ,

(2.6)

where we denote h 1 = 1 − ǫζ and h 2 = δ

1 + ǫζ − βb, as well as

Q [h 1 , h 2 ]v = T [h 2 , βb] h 1 v h 1 + γh 2

− γ T [h 1 , 0] − h 2 v h 1 + γh 2

,

= − 1

3h 2

∂ x

h 3 2 ∂ x

h 1 v h 1 + γh 2

+ 1

2h 2

β h

∂ x

h 2 2 (∂ x b) h 1 v h 1 + γh 2

− h 2 2 (∂ x b)∂ x

h 1 v h 1 + γh 2

i

+ β 2 (∂ x b) 2 h 1 v h 1 + γh 2

− γ h 1 3h 1

∂ x

h 3 1 ∂ x

h 2 v h 1 + γh 2

i .

R [h 1 , h 2 ]v = 1 2

− h 2 ∂ x ( h 1 v h 1 + γh 2

) + β(∂ x b)( h 1 v h 1 + γh 2

) 2

− γ 2

h 1 ∂ x ( − h 2 v h 1 + γh 2

) 2

− ( h 1 v

h 1 + γh 2 ) T [h 2 , βb] h 1 v h 1 + γh 2

+ γ( − h 2 v

h 1 + γh 2 ) T [h 1 , 0] − h 2 v h 1 + γh 2

,

(7)

with,

T [h, b]V ≡ − 1

3h ∂ x (h 3 ∂ x V ) + 1

2h [∂ x (h 2 (∂ x b)V ) − h 2 (∂ x b)(∂ x V )] + (∂ x b) 2 V.

If additionally, one assume medium amplitude topography variations, then the above system may be simplified. More precisely, we assume that there exists β max < ∞ such that

β = O ( √ µ) with β ∈ [0, β max ]

Withdrawing again O (µ 2 ) in (2.6) terms, one obtains

 

 

 

 

 

 

 

 

∂ t ζ + ∂ x

h 1 h 2

h 1 + γh 2 v

= 0,

∂ t v + µ

Q [h 1 , h 2f ]v + β P [h 1 , h 2f ]v

!

+ (γ + δ)∂ x ζ + ǫ 2 ∂ x

h 2 1 − γh 2 2 (h 1 + γh 2 ) 2 v 2

= µǫ∂ x

R [h 1 , h 2f ]v + β S [h 1 , h 2f ]v

+ µ γ + δ bo ∂ x 3 ζ,

(2.7)

where we denote h 2f = δ

1 + ǫζ (f corresponds to flat topography) with

Q [h 1 , h 2f ]v = T [h 2f , 0] h 1 v h 1 + γh 2f

− γ T [h 1 , 0] − h 2f v h 1 + γh 2f

,

and

P [h 1 , h 2f ]v = 1 3h 2f

∂ x

3(h 2f ) 2 b∂ x

h 1 v h 1 + γh 2f

− 1 3h 2f

∂ x

(h 2f ) 3 ∂ x

h 1 (γb)v (h 1 + γh 2f ) 2

− b

3(h 2f ) 2 ∂ x

(h 2f ) 3 ∂ x

h 1 v h 1 + γh 2f

+ 1

2h 2f

h ∂ x

(h 2f ) 2 (∂ x b) h 1 v h 1 + γh 2f

− h 2 2f (∂ x b)∂ x

h 1 v h 1 + γh 2f

i

− γ h 1 3h 1

∂ x

h 3 1 ∂ x

γbh 2f v (h 1 + γh 2f ) 2

i

− γ h 1 3h 1

∂ x

h 3 1 ∂ x − bv h 1 + γh 2f

i ,

and

R [h 1 , h 2f ]v = 1 2

h 2f ∂ x

h 1 v h 1 + γh 2f

2

− γ 2

h 1 ∂ x − h 2f v h 1 + γh 2f

2

− h 1 v h 1 + γh 2f

h

− 1 3h 2f

∂ x

h 3 2f ∂ x

h 1 v h 1 + γh 2f

i

+ γ − h 2f v h 1 + γh 2f

h 1 3h 1

∂ x

h 3 1 ∂ x

h 2f v h 1 + γh 2f

i

.

(8)

and

S [h 1 , h 2f ]v = h 2f ∂ x

h 1 v h 1 + γh 2f

h h 2f ∂ x

h 1 γbv (h 1 + γh 2f ) 2

− b∂ x

h 1 v h 1 + γh 2f

i

− h 2f (∂ x b) h 1 v h 1 + γh 2f

∂ x

h 1 v h 1 + γh 2f

+ γ

h 1 ∂ x − h 2f v h 1 + γh 2f

h h 1 ∂ x

h 2f γbv (h 1 + γh 2f ) 2

− h 1 ∂ x

bv h 1 + γh 2f

i

− h 1 v h 1 + γh 2f

h − 1 3h 2f

∂ x

h 3 2f ∂ x

h 1 γbv (h 1 + γh 2f ) 2

+ 1

3h 2f

∂ x

3h 2 2f b∂ x

h 1 v h 1 + γh 2f

− b

3h 2 2f ∂ x

h 3 2f b∂ x

h 1 v h 1 + γh 2f

i

+ 1

2h 2f

h ∂ x

h 2 2f (∂ x b) h 1 v h 1 + γh 2f

− h 2 2f (∂ x b)∂ x

h 1 v h 1 + γh 2f

i

− γh 1 b (h 1 + γh 2f ) 2

h − 1 3h 2f ∂ x

h 3 2f ∂ x h 1 v h 1 + γh 2f

i

+ γ − h 2f v h 1 + γh 2f

h 1 3h 1

∂ x h 3 1 ∂ x

γbh 2f v

(h 1 + γh 2f ) 2 − bv h 1 + γh 2f

i , + γ − γbh 2f v

(h 1 + γh 2f ) 2 + bv h 1 + γh 2f

h 1 3h 1

∂ x

h 3 1 ∂ x h 2f v h 1 + γh 2f

i

Remark 2.8 The following approximation formally hold using that:

1

1 − X = 1 + X + X 2 + O (X 3 ) with X << 1, 1

h 1 + γh 2

= 1

h 1 + γh 2f − γβb

= 1

(h 1 + γh 2f )

1 − γβb h 1 + γh 2f

= 1

(h 1 + γh 2f )

1 + γβb h 1 + γh 2f

+ (γβb) 2

(h 1 + γh 2f ) 2 + O (β 3 ) .

Proposition 2.9 (Consistency) For p = (µ, ǫ, δ, γ, β, bo) ∈ P SW , let U

p

= (ζ

p

, ψ

p

)

be a family of solutions of the full Euler system (2.1) such that there exists T > 0, s ≥ s 0 + 1, s 0 > 1/2 for which (ζ

p

, ∂ x ψ

p

)

and (∂ t ζ

p

, ∂ t ∂ x ψ

p

)

are bounded in L

([0, T ); H s+N ) 2 (N sufficiently large), uniformly with respect to p ∈ P SW . Moreover, assume that b ∈ H s+N and there exists h 01 > 0 such that

h 1 = 1 − ǫζ

p

≥ h 01 > 0 , h 2f = 1/δ + ǫζ

p

≥ h 01 > 0 , h 2 = 1/δ + ǫζ

p

− βb ≥ h 01 > 0.

Define v

p

as in (2.5). Then (ζ

p

, v

p

)

satisfies (2.7) up to a remainder term, R = (0, r) T , bounded by k r k L

([0,T );H

s

) ≤ µ 2 C 1 ,

with C 1 = C(h

01 1 , M SW , | b | H

s+N

, k (ζ

p

, ∂ x ψ

p

) T k L

([0,T);H

s+N

)

2

, k (∂ t ζ

p

, ∂ t ∂ x ψ

p

) T k L

([0,T );H

s+N

)

2

), uni- formly with respect to p ∈ P SW .

Proof.

This results has been proved in [18, Proposition 8] for the system (2.6). So the proof is straightforwardly adapted to the simplified system (2.7), using in particular the following estimates, valid for s > 1/2:

| 1

h 1 + γh 2 − 1

h 1 + γh 2f | H

s

≤ βC(h

01 1 , | ζ

p

| H

s

) | b | H

s

;

| 1

h 1 + γh 2 − 1

h 1 + γh 2f − γβb

(h 1 + γh 2f ) 2 | H

s

≤ β 2 C(h

01 1 , | ζ

p

| H

s

) | b 2 | H

s

;

see, for example, [23, Proposition B.4].

(9)

3 Construction of the new model

The present work is limited to the so-called Camassa-Holm regime, that is using additional assumption ǫ = O ( √ µ). In this section, we manipulate the Green-Naghdi system (2.7), systematically withdrawing O (µ 2 , µǫ 2 , µǫβ) terms, in order to recover our model.

One can check that the following approximations formally hold:

Q [h 1 , h 2f ]v + β P [h 1 , h 2f ]v = − λ∂ x 2 v − ǫ γ + δ

3 (θ − α)v∂ x 2 ζ + (α + 2θ)∂ x (ζ∂ x v) − θζ∂ x 2 v + β γ + δ

3 ( α 1

2 + θ 1 )v∂ x 2 b + (α 1 + 2θ 1 )∂ x (b∂ x v) − θ 1 b∂ 2 x v + O (ǫ 2 , ǫβ),

R [h 1 , h 2f ]v + β S [h 1 , h 2f ]v = α 1

2 (∂ x v) 2 + 1 3 v∂ x 2 v

+ O (ǫ, β).

with

λ = 1 + γδ

3δ(γ + δ) , α = 1 − γ

(γ + δ) 2 and θ = (1 + γδ)(δ 2 − γ)

δ(γ + δ) 3 , (3.1)

and

α 1 = 1

(γ + δ) 2 and θ 1 = δ(1 + γδ)

(γ + δ) 3 . (3.2)

Plugging these expansions into system (2.7) yields a simplified model, with the same order of preci- sion of the original model (that is O (µ 2 )) in the Camassa-Holm regime. However, we will use several additional transformations, in order to produce an equivalent model (again, in the sense of consistency), which possess a structure similar to symmetrizable quasilinear systems, and allows the study of the subsequent sections.

Using the same techniques as in [17, Section 4.2] but with a different symmetric operator T[ǫζ, βb]

defined below, we obtain the following equation:

T[ǫζ, βb](∂ t v + ǫςv∂ x v) − q 1 (ǫζ, βb)∂ t

v + µ Q [h 1 , h 2f ]v + β P [h 1 , h 2f ]v + q 1 (ǫζ, βb)µ γ + δ

bo ∂ x 3 ζ + µǫq 1 (ǫζ, βb)∂ x

R [h 1 , h 2f ]v + β S [h 1 , h 2f ]

= ǫςq 1 (ǫζ, βb)v∂ x v − µǫ 2α

3 ∂ x (∂ x v) 2

+ µβω(∂ x ζ)(∂ 2 x b) + O (µ 2 , µǫ 2 , µǫβ) (3.3)

where we denote ω = (γ + δ) 2 3

α 1

2 + θ 1 and

T[ǫζ, βb]V = q 1 (ǫζ, βb)V − µν∂ x

q 2 (ǫζ, βb)∂ x V

, (3.4)

with q i (X, Y ) ≡ 1 + κ i X + ω i Y (i = 1, 2) and ν, κ 1 , κ 2 , ω 1 , ω 2 , ς are defined as follow:

ν = λ − 1

bo = 1 + γδ 3δ(γ + δ) − 1

bo , (3.5)

(λ − 1

bo )κ 1 = γ + δ

3 (2θ − α), (λ − 1

bo )κ 2 = (γ + δ)θ, (3.6)

(λ − 1

bo )ω 1 = − θ 1

γ + δ

3 , (λ − 1

bo )ω 2 = − (γ + δ)

3 (α 1 + 2θ 1 ), (3.7) (λ − 1

bo )ς = 2α − θ

3 − 1

bo δ 2 − γ

(δ + γ) 2 . (3.8)

(10)

When plugging the estimate (3.3) in (2.7), and after multiplying the second equation by q 1 (ǫζ, βb), we obtain the following system of equations:

 

 

 

 

 

 

∂ t ζ + ∂ x

h 1 h 2

h 1 + γh 2

v

= 0,

T[ǫζ, βb] (∂ t v + ǫςv∂ x v) + (γ + δ)q 1 (ǫζ, βb)∂ x ζ + ǫ 2 q 1 (ǫζ, βb)∂ x

h

2 1−

γh

22

(h

1

+γh

2

)

2

| v | 2 − ς | v | 2

= − µǫ 2 3 α∂ x (∂ x v) 2

+ µβω(∂ x ζ)(∂ x 2 b),

(3.9)

Proposition 3.10 (Consistency) For p = (µ, ǫ, δ, γ, β, bo) ∈ P SW , let U

p

= (ζ

p

, ψ

p

)

be a family of solutions of the full Euler system (2.1) such that there exists T > 0, s ≥ s 0 + 1, s 0 > 1/2 for which (ζ

p

, ∂ x ψ

p

)

and (∂ t ζ

p

, ∂ t ∂ x ψ

p

)

are bounded in L

([0, T ); H s+N ) 2 with sufficiently large N, and uniformly with respect to p ∈ P SW . Moreover assume that b ∈ H s+N and there exists h 01 > 0 such that h 1 = 1 − ǫζ

p

≥ h 01 > 0, h 2f = 1/δ + ǫζ

p

≥ h 01 > 0, h 2 = 1/δ + ǫζ

p

− βb ≥ h 01 > 0. (H1) Define v

p

as in (2.5). Then (ζ

p

, v

p

)

satisfies (3.9) up to a remainder term, R = (0, r) T , bounded by

k r k L

([0,T );H

s

) ≤ (µ 2 + µǫ 2 + µǫβ)C,

with C = C(h

01 1 , M SW , | b | H

s+N

, k (ζ

p

, ∂ x ψ

p

) T k L

([0,T);H

s+N

)

2

, k (∂ t ζ

p

, ∂ t ∂ x ψ

p

) T k L

([0,T );H

s+N

)

2

).

Proof.

Let U = (ζ, ψ)

satisfy the hypothesis above( withdrawing the explicit dependence with respect to parameters p for the sake of readability). We know from Proposition 2.9 that (ζ, v)

satisfies the system (2.7) up to a remainder R 0 = (0, r 0 )

bounded by,

k r 0 k L

([0,T );H

s

) ≤ µ 2 C 1 ,

with C 1 = C(h

01 1 , M SW , | b | H

s+N

, k (ζ

p

, ∂ x ψ

p

) T k L

([0,T);H

s+N

)

2

, k (∂ t ζ

p

, ∂ t ∂ x ψ

p

) T k L

([0,T);H

s+N

)

2

), uni- formly with respect to p ∈ P SW .

The proof now consists in checking that all terms neglected in the above calculations can be rigorously estimated in the same way. We do not develop each estimate, but rather provide the precise bound on the various remainder terms. One has

∂ t Q [h 1 , h 2f ]v + β P [h 1 , h 2f ]v

− λ∂ x 2 ∂ t v − ǫ γ + δ

3 ∂ t (β − α)v∂ 2 x ζ + (α + 2β )∂ x (ζ∂ x v) − βζ∂ x 2 v + β γ + δ

3 ∂ t

( α 1

2 + θ 1 )v∂ x 2 b + (α 1 + 2θ 1 )∂ x (b∂ x v) − θ 1 b∂ x 2 v H

s

≤ (ǫ 2 + ǫβ)C(s + 3), with C(s + 3) ≡ C

M SW , h

01 1 , ζ

H

s+3

,

∂ t ζ H

s+2

,

v H

s+3

,

∂ t v H

s+2

,

b H

s+3

, and

∂ x R [h 1 , h 2f ]v + β S [h 1 , h 2f ]v

− ∂ x

α 1

2 (∂ x v) 2 + 1

3 v∂ x 2 v

H

s

≤ (ǫ + β)C(s + 3).

It follows that (3.3) is valid up to a remainder R 1 , bounded by

| R 1 | H

s

≤ (µ 2 + µǫ 2 + µǫβ)C(s + 3) Finally, (ζ, v) satisfies (3.9) up to a remainder R, bounded by

| R | H

s

≤ | R 1 | H

s

+ | R 0 | H

s

≤ (µ 2 + µǫ 2 + µǫβ)C.

where we use that

v

H

s+3

+ ∂ t v

H

s+2

≤ C.

The estimate on v follows directly from the identity ∂ x

h

1

h

2

h

1

+γh

2

v

= − µ 1 G µ,ǫ ψ = ∂ t ζ. The estimate on ∂ t v can be proved, for example, following [14, Prop. 2.12]. This concludes the proof of Proposition 3.10.

(11)

4 Preliminary results

In this section, we recall the operator T[ǫζ, β], defined in (3.4):

T[ǫζ, βb]V = q 1 (ǫζ, βb)V − µν∂ x

q 2 (ǫζ, βb)∂ x V

. (4.1)

with ν, κ 1 , κ 2 , ω 1 , ω 2 are constants and ν = 1 + γδ 3δ(γ + δ) − 1

bo ≥ ν 0 > 0.

The operator T[ǫζ, βb], has exactly the same structure as the one introduced in [17]. In the following, we seek sufficient conditions to ensure the strong ellipticity of the operator T which will yield to the well-posedness and continuity of the inverse T

1 .

As a matter of fact, this condition, namely (H2) (and similarly the classical non-zero depth condi- tion, (H1)) simply consists in assuming that the deformation of the interface is not too large as given in [17] but here we have to take into account the topographic variation that plays a role in (H1) and in (H2). For fixed ζ ∈ L

and b ∈ L

, the restriction reduces to an estimate on ǫ max

ζ

L

+ β max | b | L

with ǫ max = min(M √ µ max , 1), and (H1)-(H2) hold uniformly with respect to (µ, ǫ, δ, γ, β, bo) ∈ P CH . Let us shortly detail the argument. Recall the non-zero depth condition

∃ h 01 > 0, such that inf

x

∈R

h 1 ≥ h 01 > 0, inf

x

∈R

h 2f ≥ h 01 > 0, inf

x

∈R

h 2 ≥ h 01 > 0. (H1) where h 1 = 1 − ǫζ and h 2 = 1/δ + ǫζ − βb are the depth of respectively the upper and the lower layer of the fluid and h 2f = 1/δ + ǫζ the depth of the lower layer of the fluid when the bottom is flat.

It is straightforward to check that, since for all (µ, ǫ, δ, γ, β, bo) ∈ P CH , the following condition ǫ max | ζ | L

+ β max | b | L

< min(1, 1

δ max

)

is sufficient to define h 01 > 0 such that (H1) is valid independently of (µ, ǫ, δ, γ, β, bo) ∈ P CH . In the same way, we introduce the condition

∃ h 02 > 0, such that inf

x

∈R

(1 + ǫκ 1 ζ + βω 1 b) ≥ h 02 > 0 ; inf

x

∈R

(1 + ǫκ 2 ζ + βω 2 b) ≥ h 02 > 0.

(H2) In what follows, we will always assume that (H1) and (H2) are satisfied. It is a consequence of our work that such assumption may be imposed only on the initial data, and then is automatically satisfied over the relevant time scale.

Now the preliminary results proved in [17, Section 5] remain true for the operator T[ǫζ, βb]. Let us recall these results,

Lemma 4.2 Let ζ ∈ L

, b ∈ L

and ǫ max = min(M √ µ max , 1) be such that there exists h 0 > 0 with max( | κ 1 | , | κ 2 | , 1, δ max )ǫ max

ζ

L

+ max( | ω 1 | , | ω 2 | , δ max )β max | b | L

≤ 1 − h 0 < 1.

Then there exists h 01 , h 02 > 0 such that (H1)-(H2) hold for any (µ, ǫ, δ, γ, β, bo) ∈ P CH ..

Before asserting the strong ellipticity of the operator T, let us first recall the quantity | · | H

1µ

, which is defined as

∀ v ∈ H 1 ( R ), | v | 2 H

µ1

= | v | 2 L

2

+ µ | ∂ x v | 2 L

2

,

and is equivalent to the H 1 ( R )-norm but not uniformly with respect to µ. We define by H µ 1 ( R ) the

space H 1 ( R ) endowed with this norm.

(12)

Lemma 4.3 Let (µ, ǫ, δ, γ, β, bo) ∈ P CH and ζ ∈ L

( R ), b ∈ L

( R ) such that (H2) is satisfied. Then the operator

T[ǫζ, βb] : H µ 1 ( R ) −→ (H µ 1 ( R ))

is uniformly continuous and coercive. More precisely, there exists c 0 > 0 such that

(Tu, v) ≤ c 0 | u | H

µ1

| v | H

µ1

; (4.4) (Tu, u) ≥ 1

c 0 | u | 2 H

µ1

(4.5)

with c 0 = C(M CH , h

02 1 , ǫ ζ

L

, β b

L

).

Moreover, the following estimates hold:

(i) Let s 0 > 1 2 and s ≥ 0. If ζ, b ∈ H s

0

( R ) ∩ H s ( R ) and u ∈ H s+1 ( R ) and v ∈ H 1 ( R ), then:

Λ s T[ǫζ, βb]u, v ≤ C 0

(1 + ǫ ζ

H

s0

+ β b

H

s0

) u

H

s+1

µ

+

(ǫ ζ

H

s

+ β b

H

s

) u

H

s0 +1 µ

s>s

0

v

H

1 µ

, (4.6) (ii) Let s 0 > 1 2 and s ≥ 0. If ζ, b ∈ H s

0

+1 ∩ H s ( R ), u ∈ H s ( R ) and v ∈ H 1 ( R ), then:

Λ s , T[ǫζ, βb]

u, v

≤ C 0

(ǫ ζ

H

s0 +1

+ β b

H

s0 +1

) u

H

s

µ

+ (ǫ

ζ H

s

+ β

b H

s

)

u H

s0 +1

µ

s>s

0

+1

v

H

1 µ

≤ max(ǫ, β)C 0

(

ζ H

s0 +1

+

b H

s0 +1

)

u H

µs

+

( ζ

H

s

+ b

H

s

) u

H

s0 +1µ

s>s

0

+1

v

H

µ1

(4.7) where C 0 = C(M CH , h

02 1 ).

The following lemma offers an important invertibility result on T.

Lemma 4.8 Let (µ, ǫ, δ, γ, β, bo) ∈ P CH and ζ ∈ L

( R ), b ∈ L

( R ) such that (H2) is satisfied. Then the operator

T[ǫζ, βb] : H 2 ( R ) −→ L 2 ( R ) is one-to-one and onto. Moreover, one has the following estimates:

(i) (T[ǫζ, βb])

1 : L 2 → H µ 1 ( R ) is continuous. More precisely, one has k T

1 k L

2

(

R

)

H

1µ

(

R

) ≤ c 0 , with c 0 = C(M CH , h

02 1 , ǫ

ζ L

, β

b L

).

(ii) Additionally, if ζ, b ∈ H s

0

+1 ( R ) with s 0 > 1 2 , then one has for any 0 < s ≤ s 0 + 1, k T

1 k H

s

(

R

)

H

µs+1

(

R

) ≤ c s

0

+1 .

(iii) If ζ, b ∈ H s ( R ) with s ≥ s 0 + 1, s 0 > 1 2 , then one has

k T

1 k H

s

(

R

)

H

µs+1

(

R

) ≤ c s

where c ¯ s = C(M CH , h

02 1 , ǫ | ζ | H

s¯

, β | b | H

s¯

), thus uniform with respect to (µ, ǫ, δ, γ, β, bo) ∈ P CH .

Finally, let us introduce the following technical estimate, which is used several times in the subsequent sections.

Corollary 4.9 Let (µ, ǫ, δ, γ, β, bo) ∈ P CH and ζ, b ∈ H s ( R ) with s ≥ s 0 + 1, s 0 > 1 2 , such that (H2) is satisfied. Assume moreover that u ∈ H s

1 ( R ) and that v ∈ H 1 ( R ). Then one has

Λ s , T

1 [ǫζ, βb]

u , T[ǫζ, βb]v

=

Λ s , T[ǫζ, βb]

T

1 u , v

≤ max(ǫ, β) C(M CH , h

02 1 , ζ

H

s

, b

H

s

) u

H

s−1

v H

1

µ

(4.10)

(13)

5 Linear analysis

Let us recall the system (3.9).

 

 

 

 

 

 

∂ t ζ + ∂ x

h 1 h 2

h 1 + γh 2

v

= 0,

T[ǫζ, βb] (∂ t v + ǫςv∂ x v) + (γ + δ)q 1 (ǫζ, βb)∂ x ζ + ǫ 2 q 1 (ǫζ, βb)∂ x

h

2 1−

γh

22

(h

1

+γh

2

)

2

| v | 2 − ς | v | 2

= − µǫ 2 3 α∂ x (∂ x v) 2

+ µβω(∂ x ζ)(∂ x 2 b),

(5.1)

with h 1 = 1 − ǫζ , h 2 = 1/δ + ǫζ − βb , q i (X, Y ) = 1 + κ i X + ω i Y (i = 1, 2) , κ i , ω i , ς defined in (3.6),(3.7),(3.8), and

T[ǫζ, βb]V = q 1 (ǫζ, βb)V − µν∂ x (q 2 (ǫζ, βb)∂ x V ) . In order to ease the reading, we define the function

f : X → (1 − X )(δ

1 + X − βb) 1 − X + γ(δ

1 + X − βb) , and

g : X → (1 − X )

1 − X + γ(δ

1 + X − βb) 2

. One can easily check that

f (ǫζ) = h 1 h 2

h 1 + γh 2 , f

(ǫζ) = h 2 1 − γh 2 2

(h 1 + γh 2 ) 2 and g(ǫζ) = h 1

h 1 + γh 2

2

. Additionally, let us denote

κ = 2 3 α = 2

3 1 − γ

(δ + γ) 2 and q 3 (ǫζ) = 1 2

h 2 1 − γh 2 2 (h 1 + γh 2 ) 2 − ς

, so that one can rewrite,

 

 

∂ t ζ + f (ǫζ)∂ x v + ǫ∂ x ζf

(ǫζ)v − β∂ x bg(ǫζ)v = 0, T

∂ t v + ǫ

2 ς∂ x (v 2 )

+ (γ + δ)q 1 (ǫζ, βb)∂ x ζ + ǫq 1 (ǫζ, βb)∂ x (q 3 (ǫζ)v 2 ) = − µǫκ∂ x (∂ x v) 2

+ µβω(∂ x ζ)(∂ x 2 b).

(5.2) with ∂ x (q 3 (ǫζ)) = − γǫ∂ x ζ(h 1 + h 2 ) 2 + γβ∂ x bh 1 (h 1 + h 2 )

(h 1 + γh 2 ) 3 .

The equations can be written after applying T

1 to the second equation in (5.2) as

∂ t U + A 0 [U]∂ x U + A 1 [U ]∂ x U + B[U ] = 0, (5.3) with

A 0 [U ] =

ǫf

(ǫζ)v f (ǫζ)

T

1 (Q 0 (ǫζ, βb) · ) ǫT

1 (Q[ǫζ, βb, v] · )

, A 1 [U ] =

0 0 ǫ 2 T

1 (Q 1 (ǫζ, βb, v) · ) ǫςv

, (5.4)

B[U ] =

− β∂ x bg(ǫζ)v ǫT

1 γβq 1 (ǫζ, βb)h 1 (h 1 + h 2 )v 2

(h 1 + γh 2 ) 3 ∂ x b

 (5.5)

where Q 0 (ǫζ, βb), Q 1 (ǫζ, βb, v) are defined as

Q 0 (ǫζ, βb) = (γ + δ)q 1 (ǫζ, βb) − µβω∂ x 2 b, Q 1 (ǫζ, βb, v) = − γq 1 (ǫζ, βb) (h 1 + h 2 ) 2

(h 1 + γh 2 ) 3 v 2 (5.6)

(14)

and the operator Q[ǫ, βb, v] defined by

Q[ǫζ, βb, v]f ≡ 2q 1 (ǫζ, βb)q 3 (ǫζ)vf + µκ∂ x (f ∂ x v). (5.7) Following the classical theory of quasilinear hyperbolic systems, the well-posedness of the initial value problem of the above system will rely on a precise study of the properties, and in particular energy estimates, for the linearized system around some reference state U = (ζ, v)

:

∂ t U + A 0 [U ]∂ x U + A 1 [U ]∂ x U + B[U ] = 0;

U

|t=0

= U 0 . (5.8)

5.1 Energy space

Let us first remark that by construction, one has a pseudo-symmetrizer of the system that allows to cancel the smallness assumption of the Cammassa-Holm regime ǫ = O ( √ µ) for the existence and uniqueness results, given by

Z[U ] =

Q

0

(ǫζ,βb)+ǫ

2

Q

1

(ǫζ,βb,v)

f(ǫζ) 0

0 T[ǫζ, βb]

!

(5.9) One should add an additional assumption in order to ensure that our pseudo-symmetrizer is defined and positive which is:

∃ h 03 > 0 such that Q 0 (ǫζ, βb) + ǫ 2 Q 1 (ǫζ, βb, v) ≥ h 03 > 0. (H3) Let us now define our energy space.

Definition 5.10 For given s ≥ 0 and µ, T > 0, we denote by X s the vector space H s ( R ) × H µ s+1 ( R ) endowed with the norm

∀ U = (ζ, v) ∈ X s , | U | 2 X

s

≡ | ζ | 2 H

s

+ | v | 2 H

s

+ µ | ∂ x v | 2 H

s

,

while X T s stands for the space of U = (ζ, v) such that U ∈ C 0 ([0, max(ǫ,β) T ]; X s ) and ∂ t U ∈ L

([0, max(ǫ,β) T ] × R ), endowed with the canonical norm

k U k X

Ts

≡ sup

t

[0,T / max(ǫ,β)] | U (t, · ) | X

s

+ ess sup

t

[0,T / max(ǫ,β)],x

∈R

| ∂ t U (t, x) | . A natural energy for the initial value problem (5.8) is now given by:

E s (U ) 2 = (Λ s U, Z[U ]Λ s U ) = (Λ s ζ, Q 0 (ǫζ, βb) + ǫ 2 Q 1 (ǫζ, βb, v)

f (ǫζ) Λ s ζ) + Λ s v, T[ǫζ, βb]Λ s v

. (5.11) In order to ensure the equivalency of X s with the energy of the pseudo-symmetrizer it requires to add the additional assumption given in (H3).

Lemma 5.12 Let p = (µ, ǫ, δ, γ, β, bo) ∈ P CH , s ≥ 0 , U ∈ L

( R ) and b ∈ W 2,

( R ), satisfying (H1), (H2), and (H3). Then E s (U ) is equivalent to the | · | X

s

-norm.

More precisely, there exists c 0 = C(M CH , h

01 1 , h

02 1 , h

03 1 , ǫ | U | L

, β | b | W

2,

) > 0 such that 1

c 0

E s (U ) ≤ U

X

s

≤ c 0 E s (U ).

Proof.

This is a straightforward application of Lemma 4.3, and that for Q 0 (ǫζ, βb) + ǫ 2 Q 1 (ǫζ, βb, v) ≥ h 03 > 0 and f (ǫζ) > 0,

x inf

∈R

Q 0 (ǫζ, βb) + ǫ 2 Q 1 (ǫζ, βb, v)

f (ǫζ) ≥ inf

x

∈R

Q 0 (ǫζ, βb)) + ǫ 2 Q 1 (ǫζ, βb, v) sup

x

∈R

f (ǫζ)

1

,

(15)

sup

x

∈R

Q 0 (ǫζ, βb) + ǫ 2 Q 1 (ǫζ, βb, v)

f (ǫζ) ≤ sup

x

∈R

Q 0 (ǫζ, βb) + ǫ 2 Q 1 (ǫζ, βb, v)

x inf

∈R

f (ǫζ)

1

. where we recall that if (H1) is satisfied then, h 1 = 1 − ǫζ, h 2 = 1/δ + ǫζ − βb satisfy

x inf

∈R

h 1 ≥ h 01 , sup

x

∈R

| h 1 | ≤ 1 + 1/δ, inf

x

∈R

h 2 ≥ h 01 , sup

x

∈R

| h 2 | ≤ 1 + 1/δ.

Lemma 5.13 Let p = (µ, ǫ, δ, γ, β, bo) ∈ P CH , and let U = (ζ u , u)

∈ L

, b ∈ W 2,

satisfies (H1),(H2) and (H3). Then for any V, W ∈ X 0 , one has

Z[U ]V , W

≤ C V

X

0

W

X

0

, (5.14)

with C = C(M CH , h

01 1 , h

02 1 , ǫ U

L

, β b

W

2,

) .

Moreover, if U ∈ X s , b ∈ H s+2 , V ∈ X s

1 with s ≥ s 0 + 1, s 0 > 1/2, then one has

Λ s , Z[U ]

V , W

≤ C V

X

s−1

W

X

0

(5.15)

Λ s , Z

1 [U ]

V , Z[U ]W

≤ C V

H

s−1×

H

s−1

W

X

0

(5.16)

with C = C(M CH , h

01 1 , h

02 1 , ǫ U

X

s

, β b

H

s+2

) . Proof.

The proof of the Lemma 5.13 is the same as in [17, Lemma 6.4] adapted to our pseudo-symmetrizer.

5.2 Energy estimates

Our aim is to establish a priori energy estimates concerning our linear system. In order to be able to use the linear analysis to both the well-posedness and stability of the nonlinear system, we consider the following modified system

∂ t U + A 0 [U ]∂ x U + A 1 [U ]∂ x U + B [U ] = F;

U

|t=0

= U 0 . (5.17)

where we added a right-hand-side F , whose properties will be precised in the following Lemmas.

We begin by asserting a basic X 0 energy estimate, that we extend to X s space (s > 3/2) later on.

In the analysis below, the additional assumption ǫ ≤ M √ µ in p ∈ P CH is not used anymore (apart from the simplifications it offers when constructing system (5.1)).

Lemma 5.18 (X 0 energy estimate) Set (µ, ǫ, δ, γ, β, bo) ∈ P CH . Let T > 0 and U ∈ L

([0, T / max(ǫ, β)]; X 0 ) and U , ∂ x U ∈ L

([0, T / max(ǫ, β)] × R ) and b ∈ W 3,

such that ∂ t U ∈ L

([0, T / max(ǫ, β)] × R ) and

U , b satisfies (H1),(H2), and (H3) and U, U satisfy system (5.17), with a right hand side, F , such that F, Z[U ]U

≤ C F max(ǫ, β) U

2

X

0

+ f (t) U

X

0

, with C F a constant and f a positive integrable function on [0, T / max(ǫ, β)].

Then there exists λ, C 1 ≡ C(

∂ t U L

,

U L

,

∂ x U

L

, k b k W

3,

, C F ) such that

∀ t ∈ [0, T

max(ǫ, β) ], E 0 (U )(t) ≤ e max(ǫ,β)λt E 0 (U 0 ) + Z t

0

e max(ǫ,β)λ(t

t

)

f (t

) + max(ǫ, β)C 1

dt

,

(5.19)

The constants λ and C 1 are independent of p = (µ, ǫ, δ, γ, β, bo) ∈ P CH , but depend on M CH , h

01 1 , h

02 1 ,

and h

03 1 .

(16)

Proof.

Let us take the inner product of (5.17) by Z [U]U:

∂ t U, Z[U ]U

+ A 0 [U ]∂ x U, Z[U ]U

+ A 1 [U]∂ x U, Z[U ]U

+ B[U ], Z[U ]U

= F, Z[U ]U , From the symmetry property of Z[U ], and using the definition of E s (U), one deduces

1 2

d

dt E 0 (U ) 2 = 1 2 U,

∂ t , Z[U]

U

− Z[U ]A 0 [U ]∂ x U, U

− Z[U]A 1 [U]∂ x U, U

− B[U], Z[U ]U

+ F, Z[U ]U

. (5.20) Let us first estimate B[U ], Z[U ]U

. One has B[U ], Z[U ]U

= − g(ǫζ)vβ∂ x b, Q 0 (ǫζ, βb) + ǫ 2 Q 1 (ǫζ, βb, v)

f (ǫζ) ζ

+

ǫT

1 γβq 1 (ǫζ, βb)h 1 (h 1 + h 2 )v 2 ∂ x b (h 1 + γh 2 ) 3

, T[ǫζ, βb]v

≡ A 1 + A 2 .

| A 1 | ≤ βC U

L

, k b k W

2,

U

X

0

. In order to control A 2 , using the symmetry property of T[ǫζ], we write

ǫT

1 γβq 1 (ǫζ, βb)h 1 (h 1 + h 2 )v 2 ∂ x b (h 1 + γh 2 ) 3

, T[ǫζ, βb]v

= ǫ γβq 1 (ǫζ, βb)h 1 (h 1 + h 2 )v 2 ∂ x b (h 1 + γh 2 ) 3 , v

. Using the Cauchy-Schwartz inequality, one deduces

| A 2 | ≤ βC U

L

, k b k W

1,

U

X

0

. Altogether, one has

B[U ], Z[U ]U

≤ βC 1

U

X

0

≤ max(ǫ, β)C 1

U

X

0

. (5.21)

Now we have,

Z[U ]A 0 [U ] = ǫ Q

0

(ǫζ,βb)+ǫ

2

Q

1

(ǫζ,βb,v)

f(ǫζ ) f

(ǫζ)v Q 0 (ǫζ, βb) + ǫ 2 Q 1 (ǫζ, βb, v) Q 0 (ǫζ, βb) ǫQ[ǫζ, βb, v]

!

and

Z[U ]A 1 [U ] =

0 0 ǫ 2 Q 1 (ǫζ, βb, v) ǫςT[ǫζ, βb](v.)

. One has,

Z [U ]A 0 [U ]∂ x U, U

=

ǫ Q 0 (ǫζ, βb) + ǫ 2 Q 1 (ǫζ, βb, v)

f (ǫζ) f

(ǫζ)v∂ x ζ, ζ

+

Q 0 (ǫζ, βb)∂ x v, ζ +

ǫ 2 Q 1 (ǫζ, βb, v)∂ x v, ζ

+

Q 0 (ǫζ, βb)∂ x ζ, v +

ǫQ[ǫζ, βb, v]∂ x v, v and

Z[U]A 1 [U]∂ x U, U

=

ǫ 2 Q 1 (ǫζ, βb, v)∂ x ζ, v + ǫς

T[ǫζ, βb](v∂ x v), v

.

(17)

So that,

Z[U ]A 0 [U ]∂ x U, U

+ Z[U ]A 1 [U ]∂ x U, U

=

ǫ Q 0 (ǫζ, βb) + ǫ 2 Q 1 (ǫζ, βb, v)

f (ǫζ) f

(ǫζ)v∂ x ζ, ζ

+

Q 0 (ǫζ, βb)∂ x v, ζ +

ǫ 2 Q 1 (ǫζ, βb, v)∂ x v, ζ

+

Q 0 (ǫζ, βb)∂ x ζ, v +

ǫQ[ǫζ, βb, v]∂ x v, v

+

ǫ 2 Q 1 (ǫζ, βb, v)∂ x ζ, v + ǫς

T[ǫζ, βb](v∂ x v), v . One deduces that,

Z [U ]A 0 [U ]∂ x U, U

+ Z[U ]A 1 [U ]∂ x U, U

= − 1 2

ǫ∂ x

Q 0 (ǫζ, βb) + ǫ 2 Q 1 (ǫζ, βb, v)

f (ǫζ) f

(ǫζ)v ζ, ζ

∂ x Q 0 (ǫζ, βb) ζ, v

− ǫ 2

∂ x Q 1 (ǫζ, βb, v) ζ, v

+

ǫQ[ǫζ, βb, v]∂ x v, v + ǫς

T[ǫζ, βb](v∂ x v), v . One can easily remark that we didn’t use the smallness assumption of the Camassa-Holm regime ǫ = O ( √ µ) since we do not have anymore ∂ x v in the third term of the above identity.

One make use of the identity below, T[ǫζ, βb](v∂ x V ), V

= q 1 (ǫζ, βb)v∂ x V − µν∂ x (q 2 (ǫζ, βb)∂ x (v∂ x V )) , V

= − 1 2 ∂ x (q 1 (ǫζ, βb)v)V , V

+ µν q 2 (ǫζ, βb)∂ x (v∂ x V ) , ∂ x V

= − 1 2 ∂ x (q 1 (ǫζ, βb)v)V, V

+ µν q 2 (ǫζ, βb)(∂ x v)∂ x V, ∂ x V

− µν 1 2 ∂ x (q 2 (ǫζ, βb)v)∂ x V, ∂ x V . Using the same techniques as in [17, Lemma 6.5] we obtains,

Z [U ]A 0 [U ]∂ x U, U

+ Z[U ]A 1 [U ]∂ x U, U

≤ max(ǫ, β)C U

L

+

∂ x U L

+

b W

3,∞

U

2 X

0

. (5.22) The last term to estimate is U,

∂ t , Z[U ] U

. One has

U,

∂ t , Z[U ] U

≡ (v, h

∂ t , T i

v) + (ζ, h

∂ t , Q 0 (ǫζ, βb) + ǫ 2 Q 1 (ǫζ, βb, v) f (ǫζ)

i ζ)

=

v, ∂ t q 1 (ǫζ, βb) v

− µν

v, ∂ x (∂ t q 2 (ǫζ, βb))(∂ x v) +

ζ, ∂ t

Q 0 (ǫζ, βb) + ǫ 2 Q 1 (ǫζ, βb, v) f (ǫζ)

ζ .

From Cauchy-Schwarz inequality and since ζ and b satisfies (H1), one deduces

1 2 U,

∂ t , Z[U ]

U ≤ ǫC(

∂ t U L

,

U L

)

U

2 X

0

≤ max(ǫ, β)C(

∂ t U L

,

U L

)

U

2 X

0

.

(5.23) One can now conclude with the proof of the X 0 energy estimate. Plugging (5.21), (5.22) and (5.23) into (5.20), and making use of the assumption of the Lemma on F. This yields

1 2

d

dt E 0 (U ) 2 ≤ max(ǫ, β) C 1 E 0 (U ) 2 +

f (t) + max(ǫ, β)C 1

E 0 (U),

(18)

where C 1 ≡ C(

∂ t U L

,

U L

,

∂ x U L

,

b

W

3,

, C F ). Consequently d

dt E 0 (U ) ≤ max(ǫ, β)C 1 E 0 (U ) +

f (t) + max(ǫ, β)C 1

. Making use of the usual trick, we compute for any λ ∈ R ,

e max(ǫ,β)λt ∂ t (e

max(ǫ,β)λt E 0 (U )) = − max(ǫ, β)λE 0 (U ) + d dt E 0 (U ).

Thanks to the above inequality, one can choose λ = C 1 , so that for all t ∈ [0, max(ǫ,β) T ], one deduces d

dt (e

max(ǫ,β)λt E 0 (U )) ≤

f (t) + max(ǫ, β)C 1

e

max(ǫ,β)λt .

Integrating this differential inequality yields

∀ t ∈ [0, T

max(ǫ, β) ], E 0 (U)(t) ≤ e max(ǫ,β)λt E 0 (U 0 ) + Z t

0

e max(ǫ,β)λ(t

t

)

f (t

) + max(ǫ, β)C 1

dt

. (5.24)

This proves the energy estimate (5.19).

Let us now turn to the a priori energy estimate in “large” X s norm.

Lemma 5.25 ( X s energy estimate) Set (µ, ǫ, δ, γ, β, bo) ∈ P CH , and s ≥ s 0 + 1, s 0 > 1/2. Let U = (ζ, v)

and U = (ζ, v)

be such that U, U ∈ L

([0, T / max(ǫ, β)]; X s ), ∂ t U ∈ L

([0, T / max(ǫ, β)] × R ), b ∈ H s+2 and U satisfies (H1),(H2), and (H3) uniformly on [0, T / max(ǫ, β)], and such that sys- tem (5.17) holds with a right hand side, F , with

Λ s F, Z[U ]Λ s U

≤ C F max(ǫ, β) U

2

X

s

+ f (t) U

X

s

, where C F is a constant and f is an integrable function on [0, T / max(ǫ, β)].

Then there exists λ, C 2 = C(

U X

Ts

,

b

H

s+2

, C F ) such that the following energy estimate holds:

E s (U )(t) ≤ e max(ǫ,β)λt E s (U 0 ) + Z t

0

e max(ǫ,β)λ(t

t

) f (t

) + max(ǫ, β)C 2

dt

, (5.26) The constants λ and C 2 are independent of p = (µ, ǫ, δ, γ, β, bo) ∈ P CH , but depend on M CH , h

01 1 , h

02 1 , and h

03 1 .

Remark 5.27 In this Lemma, and in the proof below, the norm U

X

Ts

is to be understood as essential sup:

k U k X

Ts

≡ ess sup

t

[0,T /ǫ] | U (t, · ) | X

s

+ ess sup

t

[0,T /ǫ],x

∈R

| ∂ t U (t, x) | . Proof.

Let us multiply the system (5.17) on the right by Λ s Z[U ]Λ s U , and integrate by parts. One obtains Λ s ∂ t U, Z[U ]Λ s U

+ Λ s A 0 [U ]∂ x U, Z[U ]Λ s U

+ Λ s A 1 [U ]∂ x U, Z[U ]Λ s U

+ Λ s B[U ], Z[U]Λ s U

= Λ s F, Z[U ]Λ s U

, (5.28) from which we deduce, using the symmetry property of Z [U ], as well as the definition of E s (U ):

1 2

d

dt E s (U ) 2 = 1 2 Λ s U,

∂ t , Z[U ] Λ s U

− Z[U ]A 0 [U ]∂ x Λ s U, Λ s U

− Z[U ]A 1 [U ]∂ x Λ s U, Λ s U

Λ s , A 0 [U ] + A 1 [U ]

∂ x U, Z [U]Λ s U

− Λ s B[U ], Z[U ]Λ s U

+ Λ s F, Z[U ]Λ s U

. (5.29)

Références

Documents relatifs

The dead-water effect is then explained in that way: the ship, traveling in a density-stratified water, generates internal waves of elevation in its trail, and therefore suffers from

However, we point out that these latter provide an overall good agreement with the reference solution, as shown in Table 1 where the relative error on the wave amplitude at t = 25 T

Recently, Duchˆene, Israwi and Talhouk derived in [12], an asymptotic model for the propagation of one-dimensional internal waves at the interface between two layers of

Key words : Green–Naghdi model, Nonlinear shallow water, Dispersive waves, Nonlinear interactions, Improved dispersion, Splitting method, Hybrid method, Finite volume,

In the case of water and ammonia in these interiors, an exotic phase, called superionic ice, has been highlighted in numerical simulations, and various work have provided

Les personnes en situation de handicap sont placées dans différents types d’hébergement selon leur degré de déficience intellectuelle, mais les places

The ideal of the first settlers, namely Quakers and then Puritans, was spiritual; they rejected the supremacy of the Church of England, which they believed,

Our first aim is to find the paths followed by any operation through the binary tree, in order to obtain the access triggering rates, thanks to Equation 1. Binary trees are more