• Aucun résultat trouvé

Solitary wave solutions to the Isobe–Kakinuma model for water waves

N/A
N/A
Protected

Academic year: 2022

Partager "Solitary wave solutions to the Isobe–Kakinuma model for water waves"

Copied!
25
0
0

Texte intégral

(1)

arXiv:1910.06481v1 [math.AP] 15 Oct 2019

Solitary wave solutions to the Isobe–Kakinuma model for water waves

Mathieu Colin and Tatsuo Iguchi Abstract

We consider the Isobe–Kakinuma model for two-dimensional water waves in the case of the flat bottom. The Isobe–Kakinuma model is a system of Euler–Lagrange equations for a Lagrangian approximating Luke’s Lagrangian for water waves. We show theoretically the existence of a family of small amplitude solitary wave solutions to the Isobe–Kakinuma model in the long wave regime. Numerical analysis for large amplitude solitary wave solutions is also provided and suggests the existence of a solitary wave of extreme form with a sharp crest.

1 Introduction

In this paper we consider the motion of two-dimensional water waves in the case of the flat bottom. The water wave problem is mathematically formulated as a free boundary problem for an irrotational flow of an inviscid and incompressible fluid under the gravitational field. Lettbe the time and (x, z) the spatial coordinates. We assume that the water surface and the bottom are represented as z= η(x, t) and z=−h, respectively. As was shown by J. C. Luke [13], the water wave problem has a variational structure. His Lagrangian density is of the form

(1.1) LLuke(Φ, η) =

Z η(x,t)

−h

tΦ(x, z, t) +1

2 (∂xΦ(x, z, t))2+ (∂zΦ(x, z, t))2 +gz

dz, where Φ = Φ(x, z, t) is the velocity potential and g is the gravitational constant. M. Isobe [7, 8] and T. Kakinuma [9, 10, 11] proposed a model for water waves as a system of Euler–

Lagrange equations for an approximate Lagrangian, which is derived from Luke’s Lagrangian by approximating the velocity potential Φ in the Lagrangian appropriately. In this paper, we adopt an approximation under the form

(1.2) Φ(x, z, t)≃

N

X

i=0

(z+h)piφi(x, t),

where p0, p1, . . . , pN are nonnegative integers satisfying 0 = p0 < p1 < · · · < pN. Then, the corresponding Isobe–Kakinuma model in a nondimensional form is written as

(1.3)

















Hpitη+

N

X

j=0

x

1

pi+pj+ 1Hpi+pj+1xφj

−δ−2 pipj

pi+pj−1Hpi+pj−1φj

= 0 for i= 0,1, . . . , N,

N

X

j=0

Hpjtφj +η+1 2

N

X

j=0

Hpjxφj

2

−2

N

X

j=0

pjHpj−1φj

2

= 0,

whereH(x, t) = 1 +η(x, t) is the normalized depth of the water andδa nondimensional param- eter defined by the ratio of the mean depth h to the typical wavelength λ. Here and in what

(2)

follows we use the notational convention 00 = 0. For the derivation and basic properties of this model, we refer to Y. Murakami and T. Iguchi [14] and R. Nemoto and T. Iguchi [15]. Moreover, it was shown by T. Iguchi [5, 6] that the Isobe–Kakinuma model (1.3) is a higher order shallow water approximation for the water wave problem in the strongly nonlinear regime. We note also that the Isobe–Kakinuma model (1.3) in the caseN = 0 is exactly the same as the shallow water equations. In the sequel, we always assumeN ≥1.

In this paper, we look for solitary wave solutions to this model under the form (1.4) η=η(x+ct), φjj(x+ct), j = 0,1, . . . , N,

wherec∈Ris an unknown constant phase speed. Plugging (1.4) into (1.3), we obtain a system of ordinary differential equations

(1.5)

















cHpiη+

N

X

j=0

1

pi+pj + 1Hpi+pj+1φj

−δ−2 pipj

pi+pj−1Hpi+pj−1φj

= 0 for i= 0,1, . . . , N, c

N

X

j=0

Hpjφj+η+1 2

N

X

j=0

Hpjφj

2

−2

N

X

j=0

pjHpj−1φj

2

= 0.

As expected, this system has a variational structure, that is, the solution of this system is obtained as a critical point of the functional

LIK0, . . . , φN, η) =cMIK0, . . . , φN, η) +EIK0, . . . , φN, η), where

MIK0, . . . , φN, η) = Z

R

η(x)∂x Φapp(x, η(x)) dx, EIK0, . . . , φN, η) = 1

2 Z

R

(Z η(x)

−1

(∂xΦapp(x, z))2+ (δ−1zΦapp(x, z))2

dz+η(x)2 )

dx, and Φapp is the approximate velocity potential defined by

Φapp(x, z) =

N

X

i=0

(z+ 1)piφi(x, t).

We note that MIK and EIK represent the momentum in the horizontal direction and the total energy of the water, respectively. Both of them are conserved quantities for the Isobe–Kakinuma model (1.3). In this paper we do not use this variational structure to construct solitary wave solutions to the Isobe–Kakinuma model, whereas we use a perturbation method with respect to the small nondimensional parameter δ in the long wave regime.

In order to give one of our main results in this paper concerning the existence of a family of solutions to (1.5), we introduce normsk · k and k · kk fork= 0,1,2, . . . ,by

(1.6) kuk= sup

x∈R

e|x||u(x)|, kukk =

k

X

j=0

ku(j)k,

(3)

whereu(j)is thej-th order derivative ofu. We also introduce the function spacesBek andBok as closed subspaces of all even and odd functions u∈Ck(R) satisfying kukk <+∞, respectively, equipped with the norm k · kk, and put Bα = ∩k=0Bαk for α = e, o. The following theorem guarantees the existence of small amplitude solitary wave solutions to the Isobe–Kakinuma model in the long wave regime.

Theorem 1.1. There exists a positive constant δ0 such that for any δ ∈ (0, δ0] the Isobe–

Kakinuma model (1.5) has a solution (cδ, ηδ, φδ0, . . . , φδN), which satisfies ηδ, φδ′0 ∈ Be and φδ1, . . . , φδN ∈Bo. Moreover, the solution satisfies cδ= 1 + 2γδ2 and

(kηδ−4γδ2sech2xkk+k(φδ0+ 4γδ2tanhx)kk ≤Ckδ4, kφδj −4γγjδ4sech2xtanhxkk≤Ckδ6 for j= 1, . . . , N

for any k ∈ N, where the constants γ, γ1, . . . , γN are determined through p1, . . . , pN and the constant Ck does not depend onδ but on k.

Remark 1.2. The constantsγ andγ= (γ1, . . . , γN)Tin the statement of Theorem 1.1 are given by γ=A−11 (1−a0) and γ = (1−a0)·A−11 (1−a0) with a N ×N matrix A1 and a vector a0 given by

(1.7) A1 =

pipj pi+pj−1

1≤i,j≤N

,

1 aT0 a0 A0

=

1 pi+pj+ 1

0≤i,j≤N

,

and1= (1, . . . ,1)T. The matrices A0 andA1 are positive, so that the constantγ is also positive.

We will use these notations throughout this paper.

Remark 1.3. This theorem ensures the existence of a family of solitary wave solutions to the Isobe–Kakinuma model traveling to the left. We can also show a similar existence theorem to solitary wave solutions to the model traveling to the right.

Remark 1.4. By neglecting the term of order O(δ4), the solitary wave solutions in the dimen- sional form are given by













c≃ 1 + 2ha√gh, η(x, t)≃asech2q a

4γh3(x+ct) , φ0(x, t)≃ −p

4γgah2tanhq a

4γh3(x+ct) , φj(x, t)≃0 for j= 1, . . . , N,

where g is the gravitational constant and ais the amplitude of the wave.

In this paper we also analyze numerically the existence of large amplitude solitary wave solutions to the Isobe–Kakinuma model in a special case where we choose the parameters as N = 1 and p1 = 2, that is, (7.1). We note that even in this simplest case the Isobe–Kakinuma model gives a better approximation than the well-known Green–Naghdi equations in the shallow water and strongly nonlinear regime. See T. Iguchi [5, 6]. Numerical analysis suggests that there exists a critical value ofδ given approximately by

(1.8) δc= 0.62633493

such that for any δ ∈ (0, δc), the Isobe–Kakinuma model (7.1) admits a smooth solitary wave solution and that this family of waves converges to a solitary one of extreme form with a shape

(4)

crest as δ ↑ δc. Moreover, the included angle θ of the crest in the physical space is given approximately by

(1.9) θ= 152.6.

See Figure 1.1.

θ

Figure 1.1: Solitary wave of extreme form with the included angle θ= 152.6.

Here, we mention the related results on the existence of the solitary wave solutions to the water wave problem. The existence of small amplitude solitary waves for the water wave problem was first given by K. O. Friedrichs and D. H. Hyers [3]. Then, C. J. Amick and J. F. Toland [2]

proved the existence of solitary wave solutions of all amplitudes from zero up to and including that of the solitary wave of greatest height. On the other hand, a periodic wave of permanent form is called Stokes’ wave. The existence of Stokes’s wave of extreme form as well as the sharp crest of the included angle 120 was predicted by G. G. Stokes [4], and then proved theoretically by C. J. Amick, L. E. Fraenkel, and J. F. Toland [1]. An extension of these existence theories to the water waves with vorticity was given by E. Varvaruca [16]. The author has proved the existence of the solitary wave as well as Stokes’ wave of greatest height with a shape crest of the same included angle 120as in the irrotational case. As for the model equations for water waves, it is well-known that the Korteweg–de Vries equation has solitary wave solutions of arbitrarily large amplitude and does not have any wave of extreme form. The Green–Naghdi equations are known as higher order shallow water approximate equations for water waves in the strongly nonlinear regime and have the same solitary wave solutions as those of the Korteweg–de Vries equation, but again does not have any solitary wave of extreme form. Compared to these models, the Isobe–Kakinuma model even in the simplest case catches the property on the existence of the solitary wave of extreme form, although the included angle of the crest is not 120. We also mention a result by D. Lannes and F. Marche [12], where it was shown that an extended Green–

Naghdi equations for the water waves with a constant vorticity has a solitary wave solution of extreme form.

The contents of this paper are as follows. In Section 2, we give conservation laws for the Isobe–Kakinuma model, which will be used in the numerical analysis in Section 7. In Section 3, by using formal asymptotic analysis we calculate the first order approximate term with respect toδ2 in the expansion (3.2) of the solitary wave solution to the Isobe–Kakinuma model (1.5). In Section 4, we reduce the problem by deriving equations for the remainder terms. In Section 5, we construct Green’s functions for the linearized equations together with their estimates. In Section 6, we first prove an existence theorem for the system of linear equations derived in Section 4 and then prove the existence of small amplitude solitary wave solutions to the Isobe–Kakinuma model. Finally, in Section 7, we analyze numerically large amplitude solitary wave solutions and calculate the solitary wave of extreme form together with the included angle.

(5)

Acknowledgement

T. I. was partially supported by JSPS KAKENHI Grant Number JP17K18742 and JP17H02856.

2 Conservation laws

As was explained in the previous section, for the Isobe–Kakinuma model (1.3), the mass, the momentum in the horizontal direction, and the total energy are conserved. In the numerical analysis which will be carried out in Section 7, we also need to know corresponding flux functions.

The following proposition gives such flux functions.

Proposition 2.1. Any regular solution (η, φ0, . . . , φN)to the Isobe–Kakinuma model (1.3)sat- isfies the conservation laws

(2.1) ∂tη+∂x

N

X

j=0

1

pj + 1Hpj+1xφj

= 0,

t

( η∂x

N

X

i=0

Hpiφi

!) +∂x

(

−η∂t N

X

i=0

Hpiφi

!

−1 2η2 (2.2)

+1 2

N

X

i,j=0

1

pi+pj + 1Hpi+pj+1(∂xφi)(∂xφj)−δ−2 pipj

pi+pj−1Hpi+pj−1φiφj

= 0, and

t

 1 2η2+1

2

N

X

i,j=0

1

pi+pj+ 1Hpi+pj+1(∂xφi)(∂xφj) +δ−2 pipj

pi+pj−1Hpi+pj−1φiφj

 (2.3)

−∂x

N

X

i,j

1

pi+pj+ 1Hpi+pj+1(∂xφi)(∂tφj)

= 0.

Proof. Conservation law of mass (2.1) is nothing but the first equation in (1.3) with i= 0. It

(6)

follows from the equations in (1.3) that

t (

η∂x

N

X

i=0

Hpiφi

!)

−∂x (

η∂t

N

X

i=0

Hpiφi

!)

= (∂tη)

N

X

i=0

Hpixφi−(∂xη)

N

X

i=0

Hpitφi

=

N

X

i=0

−

N

X

j=0

x

1

pi+pj+ 1Hpi+pj+1xφj

−δ−2 pipj

pi+pj −1Hpi+pj−1φj

∂xφi

+ (∂xη)

η+1 2

N

X

j=0

Hpjxφj

2

−2

N

X

j=0

pjHpj−1φj

2

=−

N

X

i,j=0

1

pi+pj + 1Hpi+pj+1(∂xφi)(∂x2φj)−δ−2 pipj

pi+pj−1Hpi+pj−1φjxφi

+ (∂xη)

 η−1

2

N

X

i,j=0

Hpi+pj(∂xφi)(∂xφj)−δ−2pipjHpi+pj−2φiφj

=∂x

 1 2η2−1

2

N

X

i,j=0

1

pi+pj+ 1Hpi+pj+1(∂xφi)(∂xφj)−δ−2 pipj

pi+pj −1Hpi+pj−1φiφj

 , which gives conservation law of momentum (2.2). We see also that

t

 1 2η2+ 1

2

N

X

i,j=0

1

pi+pj+ 1Hpi+pj+1(∂xφi)(∂xφj) +δ−2 pipj

pi+pj−1Hpi+pj−1φiφj

= (∂tη)

η+1 2

N

X

j=0

Hpjxφj

2

−2

N

X

j=0

pjHpj−1φj

2

+

N

X

i,j=0

1

pi+pj+ 1Hpi+pj+1(∂xφi)(∂txφj) +δ−2 pipj

pi+pj−1Hpi+pj−1φitφj

=−(∂tη)

N

X

j=0

Hpjtφj+∂x

N

X

i,j=0

1

pi+pj+ 1Hpi+pj+1(∂xφi)(∂tφj)

+

N

X

i,j=0

x

1

pi+pj+ 1Hpi+pj+1xφi

−δ−2 pipj

pi+pj−1Hpi+pj−1φi

tφj

=∂x

N

X

i,j=0

1

pi+pj+ 1Hpi+pj+1(∂xφi)(∂tφj)

 , which gives conservation law of energy (2.3).

These conservation laws provide directly the following proposition.

(7)

Proposition 2.2. For any regular solution (η, φ0, . . . , φN) to the Isobe–Kakinuma model (1.5) satisfying the condition at spatial infinity

(2.4) η(x), φ0(x), . . . , φN(x), φ1(x), . . . , φN(x)→0 as x→ ±∞, we have

(2.5) cη+

N

X

j=0

1

pj+ 1Hpj+1φj = 0 and

(2.6) 1

2−1 2

N

X

i,j=0

1

pi+pj+ 1Hpi+pj+1φiφj−δ−2 pipj

pi+pj−1Hpi+pj−1φiφj

= 0.

3 Approximate solutions

We first transform the Isobe–Kakinuma model (1.5) into an equivalent system. The first equation in (1.5) withi= 0 can be integrated under the condition at spatial infinity (2.4) so that we have (2.5), particularly,

cη=−

N

X

j=0

1

pj+ 1Hpj+1φj.

Plugging this into the first equation in (1.5) with i= 1, . . . , N to eliminate cη, we have Hpi

N

X

j=0

1

pj+ 1Hpj+1φj

=

N

X

j=0

1

pi+pj+ 1Hpi+pj+1φj

−δ−2 pipj

pi+pj−1Hpi+pj−1φj

, which are equivalent to

N

X

j=0

1

pj+ 1− 1 pi+pj+ 1

Hpj+2φ′′j−2

N

X

j=0

pipj

pi+pj −1Hpjφj = 0

fori= 1, . . . , N. Therefore, (1.5) under the condition (2.4) is transformed equivalently into

(3.1)



























 cη+

N

X

j=0

1

pj+ 1Hpj+1φj = 0,

N

X

j=0

1

pj+ 1 − 1 pi+pj+ 1

Hpj+2φ′′j−2

N

X

j=0

pipj

pi+pj−1Hpjφj = 0 for i= 1, . . . , N, c

N

X

j=0

Hpjφj +η+1 2

N

X

j=0

Hpjφj

2

−2

N

X

j=0

pjHpj−1φj

2

= 0, whereH = 1 +η.

(8)

To simplify the notation, we put φ= (φ1, . . . , φN)T. Suppose that (c, η, φ0,φ) is a solution to (3.1) and can be expanded with respect to δ2 as

(3.2)









c=c(0)2c(1)4c(2)+· · ·, η=δ2(0)2η(1)4η(2)+· · ·), φ020(0)2φ0(1)4φ0(2)+· · ·), φ=δ2(0)2φ(1)4φ(2)+· · ·).

Remark 3.1. In this expansion, we assumed a priori that η, φ0, and φ are of order O(δ2).

This is essentially the assumption that the solution is in the long wave regime.

Then, equating the coefficients of the lowest power of δ2 we have A1φ(0) =0 so that φ(0) =0, becauseA1 is nonsingular. Equating the coefficients of δ2 we have





c(0)η(0)0(0) = 0,

(1−a0′′0(0)+A1φ(1)=0, c(0)φ0(0)(0) = 0.

It follows from the first and the third equations of the above system that c2(0) = 1 in order to obtain a nontrivial solution. In the following we will consider the case c(0)= 1. Then, we have (3.3) φ0(0) =−η(0), φ(1) =−γφ′′0(0) =γη(0) ,

where γ is the constant vector defined in Remark 1.2. Next, equating the coefficients ofδ4 in the first and the third equations in (3.1) we have

(1)0(1)+c(1)η(0)(0)φ0(0)+a0·φ(1)= 0, φ0(1)(1)+c(1)φ0(0)+1·φ(1)+120(0))2 = 0, which together with (3.3) yield

(3.4) 2c(1)η(0)−3

2(0)−γη(0)′′ = 0, whereγ is the positive constant defined in Remark 1.2.

As is well known, (3.4) has a family of solutions of the form c(1) =α, η(0)(x) = 2αsech2 qα

x with a parameterα >0. Then, it follows from (3.3) that

φ0(0)(x) =−2√

2αγtanh qα

x

+ const., φ(1)(x) =−4αγqα

tanh qα

x

sech2 qα

x .

Therefore, it is natural to expect that (1.5) has a family of solutions of the form

(3.5)













c= 1 +αδ2+O(δ4), η(x) = 2αδ2sech2 qα

x

+O(δ4), φ0(x) =−2√

2αγδ2tanh q

α x

+O(δ4), φ(x) =−4αγq

α

δ4tanh q

α x

sech2 q

α x

+O(δ6) with parameters α >0 and 0< δ≪1.

(9)

4 Reduction of the problem

Without loss of generality, we can assume that α = 2γ. We denote the approximate solitary wave solution obtained in the previous section byη(0) andφ(0), that is,

(4.1) η(0)(x) = 4γsech2x, φ(0)(x) =−4γtanhx.

Then, we will look for solutions to the Isobe–Kakinuma model (1.5) in the form (4.2) c= 1 + 2γδ2, η=δ2η(0)4ζ, φ02φ(0)4ψ0, φ=δ4γη(0)6ψ.

Plugging these into the first and the third equations in (3.1) we obtain (4.3)

(ζ+ψ0+G12 (2γ−η(0))ζ+η(0)ψ0 +a0·ψ+G2

4F1 = 0, ζ+ψ0+G32 (2γ−η(0)0 +1·ψ+G4

4F2 = 0, where









G1 = 2γη(0)−η(0)21η′′(0), G22η(0)η(0)′′ ,

G3 =−2γη(0)+ 12η2(0)2η′′(0),

G42(2γ−η(0)(0)′′3η(0)η(0)′′ +12κ23(0))2, withκ1=a0·γ,κ2 =1·γ, and κ3 = (p1, . . . , pN)T·γ and

F1 =ζψ0(0)

N

X

j=1

ψj+

N

X

j=1

1

pj+ 1δ−4 Hpj+1−1−δ2(pj+ 1)η(0)

jη′′(0)2ψj), F2 =F3+ 2γF4+1

2(ψ02η(0)′′ )2 +1

−4

φ(0)202η(0)′′ ) +δ4F42

− φ(0)202η′′(0))2

+1 2δ−2

N

X

j=1

pjHpj−1jη(0)2ψj)

2

N

X

j=1

pjγjη(0)

2

 , withH = 1 +δ2η(0)4ζ and

F3(0)

N

X

j=1

pjψj +

N

X

j=1

δ−4(Hpj−1−δ2pjη(0))(γjη(0)′′2ψj),

F4 =

N

X

j=1

ψj+

N

X

j=1

δ−2(Hpj−1)(γjη′′(0)2ψj).

In the above calculations, we have used the identities

N

X

j=0

Hpjφj2φ(0)402η′′(0)) +δ6

κ3η(0)η′′(0)+

N

X

j=1

ψj

+δ8F3

2φ(0)402η′′(0)) +δ6F4.

(10)

Since η(0) satisfies (3.4) with c(1)= 2γ, we have G1 =G3, so that (4.3) is equivalent to (4.4)

((η(0)−2γ)ζ+ 2(γ−η(0)0 + (1−a0)·ψ =g02f0, ζ+ψ0=gN+12fN+1,

with g0 =G2−G4, gN+1 =G1,f0 =F1−F2, and fN+1 = (η(0)−2γ)ψ0 −1·ψ−G4−δ2F2. On the other hand, plugging (4.2) into the second equation in (3.1), we obtain

(4.5) (1−a0′′0−δ2(A0−1⊗a0′′+A1ψ=g+δ2f, where

g= 2(1−a0)−A1diag(p1, . . . , pN

η(0)η(0)+ (A0−1⊗a0)γη(0)′′′ , (4.6)

f = (1−a0) δ−4(H2−1−2δ2η(0)(0)−2(H2−1)ψ′′0 (4.7)

−(A0−1⊗a0−2 diag(Hp1+2−1, . . . , HpN+2−1)

(γη(0)′′′2ψ′′)

−A1

δ−4 diag(Hp1−1−δ2p1η(0), . . . , HpN −1−δ2pNη(0)) γη(0)−2 diag(Hp1−1, . . . , HpN−1)

ψ . To summarize, the Isobe–Kakinuma model (3.1) is reduced to

(4.8)





(0)−2γ)ζ+ 2(γ −η(0)0 + (1−a0)·ψ=g02f0, (1−a00′′−δ2(A0−1⊗a0′′+A1ψ=g+δ2f, ζ+ψ0 =gN+12fN+1

for the remainder terms (ζ, ψ0,ψ). We note thatg0 andgN+1are polynomials in (η(0), η(0), η(0)′′ ), that g is a polynomial in (η(0), η(0) , η′′′(0)), and that g0, gN+1 ∈ Be and g ∈ Bo. More- over, f0 and fN+1 are polynomials in (η(0), η(0) , η′′(0), ζ, ψ0,ψ,ψ), and f is a polynomial in (η(0), η(0), η(0)′′′ , ζ, ψ′′0,ψ, δ2ψ′′), whose coefficients are polynomials in δ.

In view of (4.8) we proceed to consider the following system of linear ordinary differential equations

(4.9)





(0)−2γ)ζ+ 2(γ−η(0)0+ (1−a0)·ψ =F0, (1−a00′′−δ2(A0−1⊗a0′′+A1ψ =F, ζ+ψ0 =FN+1

for unknowns (ζ, ψ0,ψ). We can rewrite the third equation in (4.9) as

(4.10) ζ =−ψ0+FN+1,

which enable us to remove the unknown variable ζ from the equations. It follows from the second equations in (4.9) that

ψ =−γψ0′′2A2ψ′′+A−11 F,

where A2 = A−11 (A0−1⊗a0). Plugging this and (4.10) into the first equation in (4.9), we obtain

(4.11) −γψ0′′′+ (4γ−3η(0)0 =F0+ (2γ−η(0))FN+1−γ·F−δ2β·ψ′′′,

(11)

whereβ=AT2(1−a0). This is the equation for ψ0. In order to derive equations to solveψ, we putχ=ψ′′0. Then, it follows from (4.9) that

(4.12)

((1−a0)·ψ =F0+ (2γ −η(0))FN+1+ (3η(0)−4γ)ψ0, (1−a0)χ−δ2(A0−1⊗a0′′+A1ψ=F,

This is the equations for (χ,ψ). Of course, χ is expected to be ψ′′0, but we do not know it a priori. However, we have the following lemma.

Lemma 4.1. If ζ, ψ0,ψ, and χ solve (4.10)–(4.12) and satisfies χ(0) = ψ0′′(0), then we have χ=ψ′′0. Particularly, ζ, ψ0, and ψ solve (4.9).

Proof. It follows from the second equation in (4.12) that ψ=−γχ+δ2A2ψ′′+A−11 F. Plugging this into the first equation in (4.12) we have

−γχ+ (4γ−3η(0)0 =F0+ (2γ −η(0))FN+1−γ·F−δ2β·ψ′′′. Comparing this with (4.11) we obtain (χ−ψ0′′) = 0, which gives the desired results.

5 Green’s functions

In view of (4.11), we first consider the solvability of the equation for an unknownu

(5.1) −γu′′+ (4γ −3η(0))u=f

under the condition u(x) → 0 as x → ±∞. We remind that η(0)(x) = 4γsech2x, so that this equation does not depend essentially on the positive constantγ. This equation has already been analyzed by K. O. Friedrichs and D. H. Hyers [3]. Let us recall briefly their result. In order to construct Green’s function, we consider the initial value problems with homogeneous equation:

(−γu′′1+ (4γ−3η(0))u1= 0, u1(0) = 0, u1(0) = 1,

(−γu′′2 + (4γ−3η(0))u2 = 0, u2(0) = 1, u2(0) = 0.

The solutions of these initial value problems have the form (u1(x) = sech2xtanhx,

u2(x) = 18(−6−cosh(2x) + 15sech2x−15xsech2xtanhx).

We note that these fundamental solutions satisfy

(5.2) |u(k)1 (x)| ≤Cke−2|x|, |u(k)2 (x)| ≤Cke2|x|

fork= 0,1,2, . . .. Since the Wronskian isu1(x)u2(x)−u1(x)u2(x) ≡1, the solution u of (5.1) can be written as

u(x) =γ−1

C1− Z x

0

u2(y)f(y)dy

u1(x) +γ−1

C2+ Z x

0

u1(y)f(y)dy

u2(x),

(12)

where C1 and C2 are arbitrary constants. In order that this solution satisfies the condition u(x)→0 as x→ ±∞, as necessary conditions, we obtain

C2+ Z

0

u1(y)f(y)dy=C2− Z 0

−∞

u1(y)f(y)dy = 0, so that we have a necessary condition

(5.3)

Z

−∞

u1(y)f(y)dy= 0

for the existence of the solution, and that the solution has the form u(x) =γ−1u1(x)

C1

Z x 0

u2(y)f(y)dy

−γ−1u2(x) Z

x

u1(y)f(y)dy.

Nonuniqueness of the solution comes from the translation invariance of the original equations.

Proposition 5.1. Let γ >0. For any f ∈Be, there exists a unique solution u∈Be to (5.1).

Moreover, for anyk= 0,1,2, . . ., the solution satisfies

kukk+2 ≤C0kfkk+Ckkfk0, where Ck is a positive constant depending on k while C0 does not.

Proof. Since f is even and u1 is odd, the necessary condition (5.3) for the existence of the solution is automatically satisfied. Moreover, we restricted ourselves to a class of solutions which are even so that the solution is unique and given by

u(x) =−γ−1u1(x) Z x

0

u2(y)f(y)dy−γ−1u2(x) Z

x

u1(y)f(y)dy.

Using this solution formula, we easily obtain kukk+2 ≤Ckkfkk.

We proceed to improve this estimate. Differentiating (5.1)j-times, we have

−γu(j)′′+ (4γ−3η(0))u(j)=f(j)+ 3[(dxd)j, η(0)]u

Applying the previous estimate withk= 0 tou(j)and adding them forj= 1,2, . . . , k, we obtain kukk+2 ≤ Ckfkk+CkkukWk,, which together with the interpolation inequality kukWk, ≤ ǫkukWk+2,∞ +CǫkukL ≤ǫkukk+2+Cǫkuk2 for any ǫ >0 yields the refined estimate.

In view of (4.12), we then consider the solvability of the system of ordinary differential equations with constant coefficients

(5.4)

((1−a0)·ϕ=f0,

(1−a00−δ2(A0−1⊗a0′′+A1ϕ=f,

for unknowns ϕ0 and ϕ = (ϕ1, . . . , ϕN)T while f0 and f = (f1, . . . , fN)T are given functions.

We procced to construct Green’s function to this system. By taking the Fourier transform of (5.4), we obtain

0 (1−a0)T

1−a0 (δξ)2(A0−1⊗a0) +A1 ˆ ϕ0

ˆ ϕ

= fˆ0

. Now, we need to show that the coefficient matrix is invertible. Put

(5.5) q(ξ2) =−det

0 (1−a0)T 1−a0 ξ2(A0−1⊗a0) +A1

. Then, we have the following lemma.

(13)

Lemma 5.2. q(ξ2) is a polynomial in ξ2 of degree N −1. Moreover, there exists a positive constant c0 such that for any ξ ∈R we have q(ξ2)≥c0>0.

Proof. It is easy to see that

q(ξ2) =−det

0 (1−a0)T 1−a0 ξ2(A0−a0⊗a0) +A1

.

Here, we will show that the symmetric matrix A0−a0⊗a0 is positive. In fact, for any φ = (φ1, . . . , φN)T∈RN, we see that

φ·(A0−a0⊗a0)φ=

N

X

i,j=1

1

pi+pj+ 1φiφjN

X

j=1

1 pj+ 1φj

2

= Z 1

0

N

X

j=1

φjzpj 2

dz− Z 1

0 N

X

j=1

φjzpjdz 2

≥0

by the Cauchy–Schwarz inequality. Moreover, the equality holds if and only if PN

j=1φjzpj is constant forz∈[0,1], that is, φ=0. This shows that A0−a0⊗a0 is positive.

Next, we note that for any invertible matrix A and any vector xwe have det

0 xT x A

=−(detA)x·A−1x.

This equality comes easily from the identity 1 −xTA−1

0 Id

0 xT x A

=

−x·A−1x 0T

x A

.

Since A1 is also positive, for each ξ ∈ R the matrix ξ2(A0 −a0 ⊗a0) +A1 is positive too.

Therefore, we have q(ξ2)>0, which yields the second assertion of the lemma.

It is easy to see that q(ξ2) is a polynomial in ξ2 of degree less than or equal to N −1 and that the coefficient of ξ2(N−1) is

0 (1−a0)T 1−a0 A0−a0⊗a0

,

which is positive due to the positivity A0−a0⊗a0. Therefore, we obtain the first assertion of the lemma.

We denote the inverse matrix by 0 (1−a0)T

1−a0 (δξ)2(A0−1⊗a0) +A1 −1

=

q((δξ)2) q((δξ)2)T q((δξ)2) Q((δξ)2)

,

whereq(ξ2) = (q12), . . . , qN2))T andQ(ξ2) = (qij2))1≤i,j≤N. Then, the solution (ϕ0,ϕ) to (5.4) can be expressed as

(5.6)

(ϕˆ0=q((δξ)2) ˆf0+q((δξ)2)·fˆ, ˆ

ϕ=q((δξ)2) ˆf0+Q((δξ)2)ˆf.

Références

Documents relatifs

Starting form the Zakharov/Craig-Sulem formulation of the gravity water waves equa- tions, we prove that one can define a pressure term and hence obtain a solution of the

Elle a néanmoins plusieurs inconvénients : elle se prête mal aux techniques d’analyse harmonique ; la compréhension des calculs nécessite quelques connaissances en

In Section 3, we prove the equivalence between the Water-Waves equations and the Euler equations, which completes the study of the weak but not strong convergence to zero in

It turns out that one may weaken these conditions: Wang proved in [50] a global existence result for the gravity water waves equation in one space dimension, and for infinite

Special attention has been given to the critical value of the Bond number B = 1/3 , where it is shown that the g Serre equations admit peaked solitary waves of elevation (peakons)

The paper of Craig et al [10] considers large periodic bottom variations, again for dimensions n = 1 + d (d = 1, 2), deriving model equations to quite high order of accuracy for

In this study we compute numerical traveling wave solutions to a compact version of the Zakharov equation for unidirectional deep-water waves recently derived by Dyachenko

From this derivation, we prove that solutions of the water waves equations, associated with well-prepared initial data, can be approximated in the shallow water regime, to within